MangrovesEcologyBiodiversityandManagement Springer2021

Download as pdf or txt
Download as pdf or txt
You are on page 1of 551

Rajesh P.

Rastogi
Mahendra Phulwaria
Dharmendra K. Gupta Editors

Mangroves:
Ecology,
Biodiversity and
Management
Mangroves: Ecology, Biodiversity and
Management
Rajesh P. Rastogi • Mahendra Phulwaria •
Dharmendra K. Gupta
Editors

Mangroves: Ecology,
Biodiversity and
Management
Editors
Rajesh P. Rastogi Mahendra Phulwaria
Research in Environment Division CS-I (Mangroves & Coral Reef) Division
Ministry of Environment, Forest and Ministry of Environment, Forest and Climate
Climate Change Change
New Delhi, India New Delhi, India

Dharmendra K. Gupta
Ministry of Environment, Forest and
Climate Change
New Delhi, India

ISBN 978-981-16-2493-3 ISBN 978-981-16-2494-0 (eBook)


https://doi.org/10.1007/978-981-16-2494-0

# The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Singapore
Pte Ltd. 2021
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by
similar or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface

Mangroves are one of the most productive and biologically important blue-carbon
ecosystems across the coastal intertidal zone, on the planet Earth. It develops very
differently in various ecological settings; however, the expansion of mangroves is
commonly found along the tropical and subtropical coastlines around the world. In
the current scenario of serious environmental issues of global warming and climate
change, mangrove forests may play a vital role in mitigating the rising greenhouse
gas emissions due to their extreme capacity in carbon sequestration per unit area. It
has been estimated that mangrove forests can store about four to five times more
carbon than terrestrial rainforests. Moreover, mangroves play a key role in mitigating
global climate change and defend against extreme natural disasters, such as
onslaught of cyclones, floods, winds and tidal surges, to the human settlements
along the coastal ecosystems.
Various flora and fauna inhabiting mangrove ecosystems have developed several
discrete adaptations to withstand daily fluctuating environment with changes in
salinity, water and oxygen content with the rise and fall of the sea-tide. Moreover,
mangroves are rich in biological diversity of different taxonomic groups with great
ecological and commercial importance. Mangroves yield large amounts of oysters,
fish, prawns, crabs, fuel-wood, timbers, tannins and other natural products.
Owing to the utmost ecological as well as the economic importance of
mangroves, their restoration and proper management are crucial. However, increas-
ing anthropogenic activities and global climate change have raised worldwide
concerns towards the conservation, survival and productivity of mangrove
ecosystems. Therefore, effective conservation strategies and solid management
action plan should be implemented at the global level for their sustainable use,
ecological balance and development as a unique ecosystem intended for livelihood
and healthy environment.
The proposed book will emphasize the emerging information on ecology of
mangroves, with a special reference to their biodiversity and management.
This book has attempted to span the depth of mangrove’s ecology starting with
more general information such as types, importance and biogeography of the
mangrove ecosystem in Chaps. 1 and 2. Mangrove is an important habitat for
fauna, in relation to food availability, and as a nursery for several species during
different stages of their life cycles, which has been discussed in Chap. 3 as feeding

v
vi Preface

and breeding grounds of mangroves. Chapter 4 focuses on diverse environmental


factors influencing the mangrove ecosystems. Chapter 5 highlights the energy flux in
a mangrove ecosystem, which will help to understand the energy balance in a
mangrove ecosystem. The information regarding nitrogen and phosphorus budget
in mangrove ecosystems by exploring the storage, transformation and fluxes in
mangrove sediment, biomass, atmosphere and tidal waters has been discussed in
Chap. 6. Determination of carbon sinks and estimation of blue-carbon stock of
mangrove ecosystem have been reviewed in Chaps. 7 and 8. Chapters 9 and 10
address the responses of mangroves to climate change in the Anthropocene and deep
insight of mangroves in combating climate change, respectively. Chapter 11
describes the role of mangroves in pollution abatement. Chapter 12 presents the
measurement and modelling of above-ground root systems as attributes of flow and
wave attenuation function of mangroves. Chapter 13 explains the roles of mangrove
as natural barrier to environmental risks and coastal protection. Chapters 14 and 15
all deal with structure and diversity of polychaetes in mangroves of the Indian coast,
as well as structure, composition and diversity of plants in mangrove ecosystems in
different locations of the world, respectively. Chapter 16 deals with the patters of
livelihood of forest dependent dwellers in relation to the exploitation of resources at
the fringe of Indian Sundarbans. The role of mangroves in sustainable tourism
development and conserving it for future generations has been mentioned in
Chap. 17. In Chap. 18, the role of mangroves in supporting the aquaculture
industries has been discussed. Chapters 19 and 20 highlight the ecological valuation
and ecosystems of mangroves along with the development of effective management
action plans to promote the sustainability of mangrove forest reserves, respectively.
Chapter 21 presents an overview of how the RS and GIS technologies are evolving
in the context of their use for scientific and quantitative studies on mangroves.
I believe that this book will be helpful to a great extent for the academicians and
researchers in the field of mangroves. Undoubtedly, the contents incorporated in this
book can be used as a textbook by undergraduate and postgraduate students, teachers
and researchers in the fields of mangroves ecology, biodiversity and management.
I thank Ms. Aakanksha Tyagi, Senior editor (Books), Springer Nature for her
assistance in seeing it through to completion. I am sincerely grateful to the entire
team of Springer for the coordination, support and implementation of this book
project. Last but not least, I express my sincere gratitude to all the authors for their
kind collaboration and scientific contributions towards the completion of this book,
successfully.

New Delhi, India Rajesh P. Rastogi


New Delhi, India Mahendra Phulwaria
New Delhi, India Dharmendra K. Gupta
March 2021
Contents

1 Mangroves: Types and Importance . . . . . . . . . . . . . . . . . . . . . . . . . 1


K. Kathiresan
2 Biogeography of the Mangrove Ecosystem: Floristics, Population
Structure, and Conservation Strategies . . . . . . . . . . . . . . . . . . . . . . 33
P. Ragavan, K. Kathiresan, Sanjeev Kumar, B. Nagarajan,
R. S. C. Jayaraj, P. M. Mohan, V. Sachithanandam, T. Mageswaran,
and T. S. Rana
3 Mangroves as Feeding and Breeding Grounds . . . . . . . . . . . . . . . . 63
D. Arceo-Carranza, X. Chiappa-Carrara, R. Chávez López,
and C. Yáñez Arenas
4 Factors Influencing Mangrove Ecosystems . . . . . . . . . . . . . . . . . . . 97
Joanna C. Ellison
5 Energy Flux in Mangrove Ecosystems . . . . . . . . . . . . . . . . . . . . . . . 117
Engku Azlin Rahayu Engku Ariff,
Ahmad Faris Seman Kamarulzaman, and Mohd Nazip Suratman
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem . . . . . . . 127
Raghab Ray, Sandip Kumar Mukhopadhyay, and Tapan Kumar Jana
7 Mangroves as a Carbon Sink/Stocks . . . . . . . . . . . . . . . . . . . . . . . . 157
Tengku Mohd Zarawie Tengku Hashim and Mohd Nazip Suratman
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its
Dynamics in Relation to Hydrogeomorphic Settings and Land
Use-land Cover . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
Karuna Rao, Prabhat Ranjan, and AL. Ramanathan
9 Responses of Mangrove Ecosystems to Climate Change in the
Anthropocene . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Daniel M. Alongi
10 Roles of Mangroves in Combating the Climate Change . . . . . . . . . . 225
Anupam Kumari and Mangal S. Rathore

vii
viii Contents

11 Role of Mangroves in Pollution Abatement . . . . . . . . . . . . . . . . . . . 257


Arumugam Sundaramanickam, Ajith Nithin,
and Thangavel Balasubramanian
12 Measurement and Modeling of Above-Ground Root Systems as
Attributes of Flow and Wave Attenuation Function of
Mangroves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
Masaya Yoshikai, Takashi Nakamura, Rempei Suwa, Rene Rollon,
and Kazuo Nadaoka
13 Mangrove as a Natural Barrier to Environmental Risks and
Coastal Protection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
Nazlin Asari, Mohd Nazip Suratman, Nurul Atiqah Mohd Ayob,
and Nur Hasmiza Abdul Hamid
14 Diversity and Community Structure of Polychaetes in Mangroves
of Indian Coast . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
P. Murugesan and T. Balasubramanian
15 Structure and Diversity of Plants in Mangrove Ecosystems . . . . . . . 361
Nurun Nadhirah Md Isa and Mohd Nazip Suratman
16 Livelihood of Forest Dependent Dwellers in Relation to the
Exploitation of Resources at the Fringe of Indian Sundarban . . . . . 371
Chandan Surabhi Das
17 The Roles of Mangroves in Sustainable Tourism Development . . . . 401
Yarina Ahmad and Mohd Nazip Suratman
18 Aquaculture in Mangroves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419
Tengku Mohd Zarawie Tengku Hashim,
Engku Azlin Rahayu Engku Ariff, and Mohd Nazip Suratman
19 Ecological Valuation and Ecosystem Services of Mangroves . . . . . . 439
Hong Tinh Pham, Thi Hong Hanh Nguyen, and Sy Tuan Mai
20 Management Action Plans for Development of Mangrove Forest
Reserves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 455
Mohamad Danial Md Sabri, Mohd Nazip Suratman,
and Nur Hajar Zamah Shari
21 Geospatial Tools for Mapping and Monitoring Coastal
Mangroves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475
L. Gnanappazham, Kumar Arun Prasad, and V. K. Dadhwal
Editors and Contributors

About the Editors

Rajesh P. Rastogi is Scientist, at Ministry of Environment, Forest and Climate


Change (MoEF&CC), New Delhi, India. He obtained his graduate and postgraduate
degree from Patna University, Bihar and completed PhD in botany at Banaras Hindu
University, Varanasi, India. He is the recipient of some national/international awards
and fellowships. He was a visiting scientist at Friedrich Alexander University,
Nuremberg, Germany and served as a Visiting Professor of biochemistry at
Chulalongkorn University, Thailand. He has published several research/review
papers in journals/books of international repute and also edited some books. His
field of research includes freshwater/marine algae and cyanobacterial ecology, plant-
physiology, photochemistry and photobiology and stress biology.

Mahendra Phulwaria is Scientist at Ministry of Environment, Forest and Climate


Change (MoEF&CC), New Delhi, India. He received his graduation degree from
Maharshi Dayanand Saraswati University and did postgraduation and PhD in botany
from Jai Narain Vyas University, Rajasthan, India. His main field of research
includes plant tissue culture (organogenesis, synthetic seed technology, in vitro
conservation and micropropagation of woody and medicinal plants) and molecular
biology (genetic diversity analysis using DNA barcoding). He was awarded with
SERB Young Scientist by Department of Science and Technology, New Delhi and
Post-Doctoral Fellowship by UGC, New Delhi. He is a Member of International
Association of Plant Biotechnology, India. Chapter.

Dharmendra K. Gupta is Scientist ‘F’ and Director (S) of Hazardous Substance


Management Division in the Ministry of Environment, Forest and Climate Change,
Government of India at New Delhi, India and has already published more than
115 refereed research papers/review articles in peer-reviewed international journals
and also has edited 17 books of Springer. His field of research includes abiotic stress
by radionuclides/heavy metals and xenobiotics in plants, antioxidative system in
plants and environmental pollution (radionuclides/heavy metals) remediation

ix
x Editors and Contributors

through plants (phytoremediation). According to Google Scholar, his total citation of


research papers is more than 5000 and h-index is 36.

Contributors

Nur Hasmiza Abdul Hamid Faculty of Applied Sciences, Universiti Teknologi


MARA (UiTM), Shah Alam, Malaysia
Institute for Biodiversity and Sustainable Development, Universiti Teknologi
MARA (UiTM), Shah Alam, Malaysia
Yarina Ahmad Faculty of Administrative Science and Policy Studies, Universiti
Teknologi MARA (UiTM), Shah Alam, Malaysia
Institute for Biodiversity and Sustainable Development, Universiti Teknologi
MARA (UiTM), Shah Alam, Malaysia
Daniel M. Alongi Tropical Coastal & Mangrove Consultants, Pakenham, Australia
D. Arceo-Carranza Facultad de Ciencias, Unidad Multidisciplinaria de Docencia
e Investigación Sisal, Universidad Nacional Autónoma de México, Sisal, Yucatán,
Mexico
C. Yáñez Arenas Facultad de Ciencias, Unidad Multidisciplinaria de Docencia
e Investigación Sisal, Sede Parque CientÚfico, Universidad Nacional Autónoma de
México, Mérida, Mexico
Nazlin Asari Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM),
Shah Alam, Malaysia
Institute for Biodiversity and Sustainable Development, Universiti Teknologi
MARA (UiTM), Shah Alam, Malaysia
Thangavel Balasubramanian Faculty of Marine Sciences, CAS in Marine Biol-
ogy, Annamalai University, Parangipettai, India
X. Chiappa-Carrara Facultad de Ciencias, Unidad Multidisciplinaria de Docencia
e Investigación Sisal, Universidad Nacional Autónoma de México, Sisal, Yucatán,
Mexico
Escuela Nacional de Estudios Superiores Mérida, Universidad Nacional Autónoma
de México, Mérida, Yucatán, Mexico
V. K. Dadhwal Indian Institute of Space Science and Technology (IIST),
Trivandrum, India
Chandan Surabhi Das Department of Geography, Barasat Government College,
Kolkata, India
Joanna C. Ellison School of Geography, Planning and Spatial Sciences, Univer-
sity of Tasmania, Launceston, TAS, Australia
Editors and Contributors xi

Engku Azlin Rahayu Engku Ariff Faculty of Applied Sciences, Universiti


Teknologi MARA (UiTM) Pahang Campus, Bandar Tun Razak Jengka, Malaysia
L. Gnanappazham Indian Institute of Space Science and Technology (IIST),
Trivandrum, India
Tapan Kumar Jana Department of Marine Science, University of Calcutta,
Kolkata, India
R. S. C. Jayaraj Rain Forest Research Institute, Jorhat, India
K. Kathiresan Centre of Advanced Study in Marine Biology, Annamalai Univer-
sity, Parangipettai, India
Sanjeev Kumar Physical Research Laboratory, Ahmedabad, India
Anupam Kumari Academy of Scientific and Innovative Research, (AcSIR),
Ghaziabad, India
Division of Applied Phycology and Biotechnology, CSIR-Central Salt and Marine
Chemicals Research Institute (CSIR-CSMCRI), Bhavnagar, Gujarat, India
R. Chávez López Facultad de Estudios Superiores Iztacala, Universidad Nacional
Autónoma de México, Mexico City, Mexico
T. Mageswaran National Centre for Sustainable Coastal Management, MoEF &
CC, Chennai, India
Sy Tuan Mai Mangrove Ecosystem Research Center, Hanoi National University of
Education, Hanoi, Vietnam
Nurun Nadhirah Md Isa Faculty of Applied Sciences, Universiti Teknologi
MARA (UiTM) Pahang Branch, Jengka, Pahang, Malaysia
Mohamad Danial Md Sabri Forestry and Environment Division, Forest Research
Institute Malaysia, Kepong, Selangor, Malaysia
P. M. Mohan Department of Ocean Studies and Marine Biology, Pondicherry
University, Port Blair, India
Nurul Atiqah Mohd Ayob Faculty of Applied Sciences, Universiti Teknologi
MARA (UiTM), Shah Alam, Malaysia
Sandip Kumar Mukhopadhyay Department of Marine Science, University of
Calcutta, Kolkata, India
P. Murugesan Faculty of Marine Sciences, Centre of Advanced Study in Marine
Biology, Annamalai University, Parangipettai, India
Kazuo Nadaoka School of Environment and Society, Tokyo Institute of Technol-
ogy, Tokyo, Japan
B. Nagarajan Institute of Forest Genetics and Tree Breeding, Coimbatore, India
xii Editors and Contributors

Takashi Nakamura School of Environment and Society, Tokyo Institute of Tech-


nology, Tokyo, Japan
Thi Hong Hanh Nguyen Faculty of Environment, Hanoi University of Natural
Resources and Environment, Hanoi, Vietnam
Ajith Nithin Faculty of Marine Sciences, CAS in Marine Biology, Annamalai
University, Parangipettai, India
Hong Tinh Pham Faculty of Environment, Hanoi University of Natural Resources
and Environment, Hanoi, Vietnam
Kumar Arun Prasad Department of Geography, Central University of Tamil
Nadu, Thiruvarur, India
P. Ragavan Physical Research Laboratory, Ahmedabad, India
AL. Ramanathan School of Environmental Sciences, Jawaharlal Nehru Univer-
sity, New Delhi, India
T. S. Rana CSIR-National Botanical Research Institute, Rana Paratap Marg,
Lucknow, India
Prabhat Ranjan Central Pollution Control Board, New Delhi, India
Karuna Rao School of Environmental Sciences, Jawaharlal Nehru University,
New Delhi, India
Mangal S. Rathore Academy of Scientific and Innovative Research, (AcSIR),
Ghaziabad, India
Division of Applied Phycology and Biotechnology, CSIR-Central Salt and Marine
Chemicals Research Institute (CSIR-CSMCRI), Bhavnagar, Gujarat, India
Raghab Ray Atmosphere and Ocean Research Institute, The University of Tokyo,
Kashiwa, Japan
Rene Rollon Institute of Environmental Science and Meteorology, College of
Science, University of the Philippines, Quezon City, Philippines
V. Sachithanandam National Centre for Sustainable Coastal Management, MoEF
& CC, Chennai, India
Ahmad Faris Seman Kamarulzaman Faculty of Applied Sciences, Universiti
Teknologi MARA (UiTM) Pahang Campus, Bandar Tun Razak Jengka, Malaysia
Nur Hajar Zamah Shari Forestry and Environment Division, Forest Research
Institute Malaysia, Kepong, Selangor, Malaysia
Arumugam Sundaramanickam Faculty of Marine Sciences, CAS in Marine
Biology, Annamalai University, Parangipettai, India
Mohd Nazip Suratman Faculty of Applied Sciences and Institute for Biodiversity
and Sustainable Development, UiTM, Shah Alam, Malaysia
Editors and Contributors xiii

Rempei Suwa Forestry Division, Japan International Research Center for Agricul-
tural Sciences (JIRCAS), Tsukuba, Japan
Tengku Mohd Zarawie Tengku Hashim Faculty of Applied Sciences and Insti-
tute for Biodiversity and Sustainable Development, UiTM, Shah Alam, Malaysia
Masaya Yoshikai School of Environment and Society, Tokyo Institute of Tech-
nology, Tokyo, Japan
Mangroves: Types and Importance
1
K. Kathiresan

Abstract

Mangroves are of great ecological significance and economic importance. They


are of different types—deltaic, estuarine, lagoon, and fringe mangroves—based
on coastal location. The mangroves are of six functional types—fringe, riverine,
basin, over-wash, scrub, and hammock. They are also of three types—river-
dominated, tide-dominated, and interior mangroves—based on tidal range and
sedimentation. In addition, there are six broad types of mangroves: large deltaic
systems, tidal plains, composite plains, fringing barriers with lagoons, drowned
bedrock valleys, and coral coasts. Mangroves are ecologically significant in
protecting the coast from solar UV-B radiation, ‘greenhouse’ gases, cyclones,
floods, sea level rise, wave action, and coastal soil erosion. They act as nutrient
sinks, sediment traps, and nutrient source to support the food web in other coastal
ecosystems. The mangroves are the most efficient in carbon sequestration and
climate change mitigation. They provide feeding, breeding, and nursery grounds
for many food fishes and wildlife animals. They protect other marine systems
such as islands, coral reefs, seaweeds, and seagrass meadows. Mangroves are
economically valuable in supplying the forestry and fishery products and also in
serving as sites for developing a burgeoning eco-tourism. The mangroves are of
great bioprospecting potential as a source of salt-tolerant genes, chemicals, and
valuable products that can be used in medical, industrial, agricultural, and food
sectors.

K. Kathiresan (*)
Centre of Advanced Study in Marine Biology, Annamalai University, Parangipettai, Tamil Nadu,
India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 1


Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_1
2 K. Kathiresan

Keywords

Mangroves · Ecological role · Economic value · Bioprospecting · Coastal


ecosystems

1.1 Introduction

Mangroves are highly productive and biologically diverse coastal habitats of tropical
and subtropical regions of the world. They are the one place on earth where land,
freshwater, and ocean mix. They are found in intertidal areas along sheltered shores and
estuaries. The mangroves have diversified habitats, such as water bodies, forests, litter
forest floor, mudflats and adjacent coral reefs and seagrass meadows; and this habitat
diversity supports a wide variety of organisms. The mangrove habitats are chiefly
colonized by flowering trees and shrubs, remarkably adapted to harsh coastal
conditions, such as seawater, periodic inundation and exposure, wind, waves, strong
currents, and anaerobic soil. There are no other groups of plants in the entire plant
kingdom with such highly developed adaptations. The ‘standing crop’ of mangroves is
greater than any other aquatic ecosystems on the earth (Duke et al. 1998; Kathiresan and
Bingham 2001; Kathiresan and Qasim 2005; Spalding et al. 2010; Tomlinson 2016).
Mangroves occur in low-lying coastal plains where the topography is smooth, but
not steep and the tidal amplitude is large. They have luxuriant growth in the alluvial
soil substrates with fine-textured loose mud or silt, rich in humus. The mangrove
plants find difficult to colonize the coastal zone with waves of high energy, and
hence they normally establish themselves in sheltered shorelines (Kathiresan and
Bingham 2001; Kathiresan and Qasim 2005).
Mangroves are the only tall tree forest system, located between land and sea in
tropical and subtropical coasts. They are a rare forest type in the world with 80 tree
and shrub species, occupying 13.8 million hectares in 118 countries and territories
(Giri et al. 2011; Duke 2017). They are largely restricted to the latitudes between 32
N and 38 S. Growth and biomass production of the mangroves decrease with
increasing latitudes and they are the highest around the equator region. The
mangroves are often called as ‘tidal forests’, ‘coastal woodlands’, ‘oceanic rain-
forests’ or ‘blue carbon forest’. Unfortunately, long-term survival of the mangroves
is at a great risk, and the ecosystem services offered by them may totally be lost in
the world within the next 100 years (Duke et al. 2007). The present chapter deals
with different types of mangroves, ecological significance, and economic impor-
tance, for better conservation and management.

1.2 Types of Mangroves

Mangroves and their environment are strongly interacting with each other. The
mangroves influence chemical and physical conditions of their environment,
which in turn influence growth and productivity of the mangroves. They are found
1 Mangroves: Types and Importance 3

Fig. 1.1 Tropical coastal settings of mangrove forests

in a variety of tropical coastal settings such as the deltas, estuarine deltas, estuaries,
estuarine lagoons, lagoons, and coastal fringes (Fig. 1.1) (Kjefve 1990). The deltas
are formed by active deposition of sediment at the river mouth, which is colonized
by mangroves. Estuaries are the sites of sediment deposition, which is colonized by
mangroves. Coastal lagoons are formed behind sand spits and barrier islands; these
sites are of less wave action and colonized by mangroves. Mangroves do occur on
sediment substrates along the coastal fringes where wave energy is low.
There are six functional types of mangrove forests as shown in Fig. 1.2, namely,
over-wash, fringe, riverine, basin, scrub (dwarf), and hammock forests (Lugo and
Snedaker 1974; Woodroffe 1992). The last three types are the modified forms of the
first three types. The six types can be summarized as follows:

1. Over-wash mangrove forests: These are small mangrove islands, frequently


formed by tidal washings.
2. Fringe mangrove forests: These occur along the borders of protected shorelines
and islands, influenced by daily tidal range. They are sensitive to erosion and long
exposure to turbulent waves and tides.
3. Riverine mangrove forests: These are luxuriant patches of mangroves existing
along rivers and creeks, which get flooded daily by the tides. Such forests are
influenced by large amounts of freshwater and nutrients and thus making the
system highly productive with tall trees.
4. Basin mangrove forests: These are stunted mangroves located along the interior
side of the swamps and in drainage depressions.
4 K. Kathiresan

Fig. 1.2 Six functional types of mangrove forests

5. Hammock mangrove forests: These are similar to the basin type but occurring in
more elevated sites than the four types given-above.
6. Scrub mangroves: These are dwarf mangrove forests occurring along flat coastal
fringes.

The above classification is not providing any information on the physical pro-
cesses that take place in mangrove forests. Considering the physical processes,
another classification has been proposed with three types of mangrove forests:
(1) river-dominated, (2) tide-dominated, and (3) interior mangrove forests (that
have less influence of river/or tides) (Fig. 1.3) (Woodroffe 1992). The river-
dominated mangroves have a strong out-welling, whereas the tide-dominated
mangroves have bidirectional flux, and while the mangroves that are located in
interior region are typical sinks for sediment nutrients.
There are six broad classes of mangrove settings, based on the tidal range and
sedimentation:

1. Large deltaic systems (occurring in low tidal range and substrate with very fine
sediments) (e.g. mangroves of Borneo, Sundarbans).
2. Tidal plains (where large mudflats are formed by alluvial sediments, reworked
by tides, and then the mudflats are colonized by mangroves).
3. Composite plains (influenced by both tidal and alluvial conditions, e.g. lagoons
formed behind wave-built barriers where mangroves grow).
4. Fringing barriers with lagoons (high wave energy conditions with sediments of
fine sand and mud) (e.g. mangroves of the Philippines)
1 Mangroves: Types and Importance 5

Fig. 1.3 Three Major types


of mangrove forests

5. Drowned bedrock valleys (e.g. mangroves of Northern Vietnam or Eastern


Malaysia)
6. Coral coasts (mangroves growing at the bottom of coral sand or in platform
reefs) (e.g. mangroves of Indonesia, and Singapore).

Based on global distribution, the mangroves are of two types: ‘Old world
mangroves’ and ‘New world mangroves’. The ‘Old world mangroves’ is the place
of origin for mangroves in the Eastern hemisphere, whereas the ‘New world
mangroves’ is the place of relatively recent origin for mangroves in the Western
hemisphere. The Eastern hemisphere is Indo-West Pacific region that includes East
Africa, Indo-Malaysia, and Australasia. The Western hemisphere is Atlantic East
Pacific region that includes West America, East America, and West Africa. Thus,
there are six geographic regions of global mangroves in the world. The Eastern
hemisphere has 57% of global mangrove area, while Western hemisphere has 43%.
The Eastern hemisphere is rich in biodiversity with 63 mangrove species, while the
Western hemisphere is poor in biodiversity with only 19 species. Mangroves have
broader ranges along the warmer eastern coastlines of the Americas and Africa than
along the cooler western coastlines (Fig. 1.4). This difference in distribution is
attributed to the presence of warm and cold oceanic currents (Duke 1992).
Global mangrove habitats are inhabited by 80 tree and shrub species including
69 species and 11 hybrids, belonging to 32 genera under 17 families worldwide.
Except one genus all the other genera are flowering plants (Tomlinson 2016; Duke
et al. 1998; Duke 2017). Mangrove diversity is the highest around South East Asia in
the old world mangroves. Some mangrove genera are specific to some regions:
Pelliciera, Conocarpus, and Laguncularia are present only in the new world,
whereas Osbornia and Camptostemon exist only in the old world. The mangrove
fern Acrostichum aureum is the only species, common to both the new and old
worlds. In the world, there are two places, richest in mangrove genetic diversity, and
these are the Bhitarkanika (Odisha) in India and the Baimuru in Papua New Guinea.
6 K. Kathiresan

Fig. 1.4 Global distribution of mangroves with six geographic regions

Among the countries, Indonesia has the largest extent of mangrove cover.
Sundarbans is the largest single block of mangrove forest in the world, and the
second largest ones are in Niger Delta, northern Brazil, and southern Papua.

1.3 Mangrove Ecosystem Services

1.3.1 Protecting from Intense Sunlight and UV-Radiation

Mangroves are able to grow under intense sunlight and solar UV-B radiation.
Avicennia marina grows under hot and arid conditions with high sunlight, while
Rhizophora species are quite tolerant to solar UV-B radiation. The mangroves are
rich in phenolic compounds that dissipate the excessive sunlight. The flavonoids
accumulated in the mangrove leaves serve as UV-screen compounds. Hence, the
mangroves are free from the deleterious effects of heat energy and UV-B radiation
under-canopy environment (Moorthy and Kathiresan 1997a, b).

1.3.2 Reducing ‘Greenhouse Gas’ and Carbon Sequestration

The coastal vegetated habitats such as mangroves, salt marshes, and seagrasses are
considered to be ‘blue carbon ecosystems’. They are amongst the most significant
carbon reservoir in the biosphere, and they play a great role in oceanic carbon cycle.
The annual carbon burial by coastal vegetated habitats is 180 times greater than that
in the deep sea sediments. The vegetated coastal habitats cover only less than 0.2%
of the seafloor, but they contribute about 50% of the global burial of organic carbon
in marine sediments (Duarte et al. 2005).
1 Mangroves: Types and Importance 7

Among blue carbon ecosystems, the mangroves are efficient in removing atmo-
spheric CO2 through photosynthesis, and storing carbon in their biomass and soil
substrates. By area, mangrove forests constitute only 0.1% of the world’s plant
biomes and only 0.7% of the global coastal zone, yet they contribute to the global
carbon in a larger way. Mangrove net primary production equals 10% of total net
primary production in the coastal zone (Alongi 2007).
Mangroves are the only blue carbon forest of the world. The mangrove forests
account for only 1% (13.5 Gt.year 1) of carbon sequestration by the world’s forests,
but they account for 14% of carbon sequestration by the global ocean (Alongi 2012).
They supply more than 10% of the organic carbon, essential to the world oceans
(Dittmar et al. 2006). Hence, the mangroves reduce the problems of ‘greenhouse
gases’ and global warming (Kathiresan and Bingham 2001).
Mangrove forests are globally significant carbon sinks, storing carbon in a range
of 455–856 mega-gram in one hectare of forest and 6.2–11.7 peta-gram in the world
(Alongi 2020; Kauffman et al. 2020; Ouyang and Lee 2020). They are 10 times
greater in carbon sequestration and four times efficient in carbon stocks than other
tropical forests (McLeod et al. 2011). The mangroves also have larger carbon stock
in tropics (895  90 MgC ha 1) than sub-tropics (547  66 MgC ha 1) (Sanders
et al. 2016). The mangrove forests hold higher carbon stock in Asia-Pacific
(1094 MgC ha 1) than other regions: Latin America (939 MgC ha 1), West Central
Africa (799 MgC ha 1), and Arabian/Oman Gulf (217 MgC ha 1) (Kauffman and
Bhomia 2017).
Mangroves are highly productive, storing large amounts of carbon in their soil
system over a very long period of time due to high sedimentation rates and anoxic
soils (Donato et al. 2011; Atwood et al. 2017; Alongi 2018) in contrast to other forest
soils that store carbon only for a short time. The mangrove soil is reported to account
for 76.5% of total carbon in ecosystem (Alongi 2014, 2020) and to store carbon in a
range of 2.6–6.4 PgC in the top one metre of soil in the global mangroves (Jardine
and Siikamaki 2014; Atwood et al. 2017; Sanderman et al. 2018). The mangrove soil
is more efficient in carbon burial by 2.4-folds than salt marshes, by 5.2-folds than
seagrasses, and by four-fold than in tropical forests (Duarte et al. 2005). In addition,
mangrove root biomass is higher than other forest types. The biomass invested in
mangrove roots is 40% of shoot, in contrast to 25% in upland forests. This higher
root biomass helps to ensure stability in the soft substrates of mangrove
environment.
Mangroves have high carbon sequestration, which is estimated to be 14.2 TgC.
year 1 and the value per unit area is 1.71  0.17 MgC ha 1 year 1 for global
mangroves (Alongi 2018). The Indian Sundarbans has the capacity to sequester a
total of 2.79 TgC in its natural forest area of 4264 km2 (Ray and Jana 2017). The
carbon sequestration is high at early stage of the forest. A 20-year old plantation of
mangroves stores 11.6 kg m 2 of carbon with C burial rate of 580 g m 2 year 1 in
Japan (Fujimoto 2000). In Malaysia, a 20 year old stand of Rhizophora apiculata
mangrove forest is reported to store 7.14 MgC ha 1 year 1. The rate of carbon
sequestered in mangrove mud is estimated to be 1.5 MgC ha 1 year 1. Each hectare
of mangrove sediment contains 700 Mg carbon per metre depth (Ong 1993; Ong
8 K. Kathiresan

et al. 1995). However, species-wise variations do occur. Avicennia marina performs


better to display 75% higher rate of carbon sequestration than that in Rhizophora
mucronata (Kathiresan et al. 2013a). Total carbon is 98.2% higher in natural
mangroves and 41.8% in planted mangroves than that in non-mangrove soil
(Kathiresan et al. 2014).
Mangrove loss disturbs the carbon stocks resulting in emission of greenhouse gas
(Alongi 2012). A loss of about 35% of the world’s mangroves has resulted in a net
loss of 3.8  1014g C stored as mangrove biomass (Cebrain 2002). Mangrove
deforestation in the world generates emissions of 0.02–0.12 pico grams of carbon
per year. This is as much as around 10% of emissions from global deforestation of all
forests (Spalding et al. 2010). Carbon storage is reduced in disturbed mangroves than
that in intact mangroves: by four folds lower in degraded or deforested forests, three
folds lower in the forests, impacted by domestic or aquaculture effluents, and two
folds lower in the forests, affected by storms and flood (Perez et al. 2018). Thus
failing to preserve mangrove forests can cause considerable carbon emissions and
thus global warming. Therefore, mangrove restoration can be a novel counter-
measure for global warming issue.

1.3.3 Protecting from Cyclone and Storms

Mangrove forests save the coastal people against cyclones and storms. The ‘super-
cyclone’ that occurred on the 29th October 1999 with a wind speed of 310 km/h along
the Odisha coast in India caused heavy damage in the coastal area that was devoid of
mangroves. But, there was practically no damage occurred in the coastal areas with
dense mangrove forest. This super-cyclone killed over 10,000 people and caused a
heavy loss of livestock and property. This loss would have been avoided, had the
mangrove forests been intact. The protection value of one hectare of intact mangroves
is reported to be 8700 US dollars, as against the value of 5000 US dollars fetched for
one hectare of cleared land. Yet another example is the cyclone ‘Nargis’ that hit the
coast of Myanmar on the third May, 2008 killed over 30,000 people and heavy loss of
properties only in the areas that were devoid of mangroves. The beneficial effect of
mangroves was also recorded during the cyclones especially ‘Aila’—2009,
‘Ockhi’—2017, ‘Gaja’—2018 along the east coast of India. The mangroves act as
a defence force against natural calamity and hence protecting mangroves as storm
buffers generates more value to society.

1.3.4 Protecting from Giant Waves

Mangrove forests protect the coast against strong wave actions. This was evident
during the 26th December 2004 tsunami that occurred in the Indian Ocean area,
which killed 3 million people in Asian and African countries and caused a loss of
6 billion US dollars in 13 countries. However, mangroves mitigated the deleterious
impact of the tsunami waves and protected the shoreline against damage (Kathiresan
1 Mangroves: Types and Importance 9

Fig. 1.5 The 2004 tsunami broken a long boat jetty in to pieces, while mangroves intact without
damage in Parangipettai, south east India

and Rajendran 2005; Danielsen et al. 2005). However, the 2004 tsunami affected
mangroves in some places, especially in Andaman and Nicobar Islands. In
Andaman, the tsunami caused considerable change in the mangrove stands of the
islands; where Avicennia marina and Sonneratia alba were not generally affected,
while Rhizophora species got affected due to continuous submergence by seawater
as a result of the tsunami waves. In South Andaman, 30–80% of mangrove stands
got affected due to natural elevation of land; however, in middle Andaman and North
Andaman, mangroves were not affected (Dam Roy and Krishnan 2005;
Ramachandran et al. 2005).
The role of mangroves in reducing the sea-waves is well-known. A hydraulic
experiment has proved that mangroves are more effective for reduction of wave
damages than concrete seawall structures such as wave dissipating block, breakwater
rock, and houses (Harada et al. 2002) (Fig. 1.5). Another study has proved that six-
year-old mangrove forest of 1.5 km width reduces the sea-waves by 20 times, from
1 m high waves at the open sea to 0.05 m at the coast (Mazda et al. 1997). The
reduction of wave amplitude and energy by tree vegetation has also been proved by
measurements of wave forces and modeling of fluid dynamics (Massel et al. 1999).
According to an analytical model, 30 trees from 10 m2 in a 100 m wide belt can
reduce the maximum tsunami flow pressure by more than 90%, if the wave height is
less than 4–5 m (Hiraishi and Harada 2003). As per our observation, the mangroves
can provide protection against tsunami in the situation, where the height of man-
grove forest (with >25 trees/10 m2) is higher than the tsunami wave height
(Kathiresan and Rajendran 2005). Therefore, conserving or restoring mangroves
will save the coast from future events of natural disasters such as tsunami.
10 K. Kathiresan

1.3.5 Controlling Flood Damage

Mangroves protect coastal systems against floods that are often caused by tidal
waves or heavy rainfall in association with storms. The mangroves are able to
control the flood due to the root system, which has a larger spread out area and
also ability to promote sedimentation.
Global mangroves provide flood protection benefits exceeding 65 billion US
dollars per year. If mangroves were lost, annually 15 million more people would
be affected by flood across the world. The greatest economic benefits are recorded
with USA, China, India, and Mexico. The economic benefits in terms of people
protected are found in Vietnam, India, and Bangladesh. Many 20-km coastal
stretches receive more than 250 million US dollars annually in flood protection
benefits from mangroves (Pelayo et al. 2020). The mangroves protect more than
150,000 people from flooding every year in Abidjan and Lagos in West Africa,
Mumbai and Karachi in South Asia, Wenzhou in East Asia, and Cebu and Denpasar
in South east Asia. The mangroves provide annually over 500 million US dollars in
avoiding the property loss in cities such as Miami in the USA and Cancun in Mexico.
However, the mangroves are beneficial not only to urban areas but also to less
populated coastal floodplains (Pelayo et al. 2020).
Besides flood control, the mangroves do prevent the entry of seawater inland and
protect the underground water systems, which are a source of drinking water supply
to coastal population (Kathiresan 2018). In addition, the mangroves reduce the
salinity of the groundwater. This is evident by the fact that there is a very sharp
decline in salt concentrations of groundwater at the interface between salt flats and
mangroves (Ridd and Sam 1996).

1.3.6 Preventing Coastal Soil Erosion

Mangroves reduce the wave action and prevent the coastal erosion. The reduction of
waves increases with density of vegetation. In a tall mangrove forest, the rate of
wave reduction is as large as 20% in a distance of 100 metre (Mazda et al. 1997).
Mangrove forests are ‘live seawall’-like natural formations and are very cost-
effective as compared to the concrete seawall and other structures for coastal
protection against erosion (Harada et al. 2002). The mangrove forest with 100 m
width is proved to protect the sea dyke, lying behind the forest, for more than
50 years, in contrast to rock fencing that protects the sea dyke for only 5 years.
This is because of the fact that the rock fencing is not long resistant to wave damage
as compared to mangrove forest, as proved in the Red River Delta, Vietnam. The
cost of mangrove planting is 1.1 million US dollars, but it has helped to avoid the
maintenance cost of 7.3 million US dollars per year for the sea dyke. However,
mangrove deforestation causes coastal erosion, as proved in the Gulf of Kachchh and
other regions (World Disaster Report 2002).
1 Mangroves: Types and Importance 11

1.3.7 Trapping Coastal Sediments

Mangroves are capable of trapping sediment, and thus acting as sinks for the
suspended sediments (Woodroffe 1992; Wolanski et al. 1992; Wolanski 1995;
Furukawa et al. 1997). The mangroves catch sediments by their complex aerial
root systems and thus they function as land expanders. Generally annual sedimenta-
tion rate ranges between 1 and 8 mm, in the mangrove areas (Bird and Barson 1977).
A contrary view is also proposed that the mangrove forests are the result, and not the
cause of sedimentation in coastal areas, but they accelerate the sedimentation
process. This depends largely on the exchange process taking place between
mangroves and the adjoining coastal areas (Woodroffe 1992).
The mechanism of sediment trapping in mangrove habitats is shown in Fig. 1.6
(Furukawa et al. 1997; Kathiresan 2003). The mangroves inhibit tidal flows, due to
the friction force that is provided by the trees with complex aerial root structures. The
soil particles are carried in suspension by the incoming tide, whereas the soil
particles are left behind in the mangrove swamps by the outgoing tides. Thus, the
particles settle down in the forests during the low tide, when the turbulence gets
reduced and the water velocity becomes low and sluggish to carry the soil particles
back to the sea. In contrast, the particles are held in suspension during the high tide,
when the turbulence is high.
Density of mangrove species and their complexity of root systems are the most
important factors that determine the sedimentation process (Kathiresan 2003). The
sedimentary process also varies in different types of mangrove forests: riverine,
basin, and fringe types. The process falls in decreasing order: River-
ine > basin > fringe (Ewel et al. 1998). The river-dominated system receives

Mangroves

Turbulence (?) Friction (?)

High tide Low tide

Increase (?) Decrease (?) Reduces tidal


Turbulence Turbulence flow

High tide Low tide

Maintaining Settlement of
sediment in suspended
suspension sediment

Fig. 1.6 Mechanism of sedimentation as induced by mangroves


12 K. Kathiresan

sediment supply from the upland areas. The tide-dominated fringe system contains
abundant sediment from the sea, but the sedimentation gets disturbed by the tides.
The interior basin mangroves are the sinks for sediments (Woodroffe 1992).

1.3.8 Deepening of Creeks

Water circulation through mangrove forests is important for keeping creeks deeper.
The water movement in tidal creek is different from surrounding mangrove swamps.
This is because of the fact that the tidal creek is deep, while the mangrove swamp is a
shallow system, colonized with vegetation. The cause of water movement is the tide.
The flow of water during the low tide is much greater than that of the high tide. For
example, the riverine mangroves produce asymmetrical tidal currents which are
stronger at the low tide than at the high tide (Medeiros and Kjerfve 1993). The
fast low tide tends to flush out the material from the mangrove swamp area and
maintains the depth of creeks. When the area of forest swamp is reduced, the speed
of the low tide is reduced, and the creeks get clogged up. This is commonly observed
in some Southeast Asian countries where deforestation of mangroves has reduced
the navigability of the canals and river mouths (Wolanski et al. 1992).

1.3.9 Trapping & Recycling of Nutrients

Mangrove soil serves as a ‘sink’ for retaining nutrients. This depends on the soil
characteristics and water flow patterns of the mangrove habitats. Mangrove soil,
algae, microbes, and physical processes absorb large amounts of organic and heavy
metal pollutants (Wong et al. 1995). Therefore, mangroves especially Avicennia
marina are capable of surviving in the areas that are dumped with heavy organic
wastes and toxic heavy metals. Ammonium nitrogen is rapidly assimilated by
bacteria and benthic algae present in the mangrove soil and hence, export of
ammonium nitrogen is largely prevented. The loss of nitrogen and phosphorus is
also significantly prevented due to reduced flow of water in the severely damaged
mangrove areas, as reported in the North Queensland (Kaly et al. 1997).
Mangrove soil is highly efficient in absorbing and holding heavy metals such as
Fe, Zn, Cr, Pb, Cd, Mn, and Cu, thereby preventing the metal pollution in coastal
environments. In the mangrove ecosystem, heavy metals are mostly trapped down in
the soil, but only <1% of all these metals are present in mangrove vegetation (Silva
et al. 1990). This trend of higher levels of heavy metals in soil than vegetation is due
to (1) low availability of metals to the plants from sediments, (2) exclusion of the
metals by the mangrove plant itself, and (3) physiological adaptations that prevent
accumulation of metals inside the plants. Oxygen exuded by the underground roots
promotes the formation of iron plaques that adhere to the root surfaces and prevent
the trace metals from entering the root cells. The metals are precipitated in the form
of stable metal sulphides under anoxic conditions and the sulphides are buried deep
into the mangrove soil. This process decreases the bioavailability of trace metals to
1 Mangroves: Types and Importance 13

the plants from the mangrove soil. Mercury does not form sulphides, and hence it is
immobilized in organic complexes in mangrove soil. Disturbances may cause the
mangrove soil to lose its metal binding capacity, resulting in the mobilization of
metals. The degrading mangroves then shift their site from ‘sink’ to ‘source’ of
heavy metals (Lacerda 1998; Kathiresan and Bingham 2001).
In addition to retaining nutrients, the mangrove systems also help in recycling of
nutrients especially carbon, nitrogen, and sulphur and making the nutrients available
in assimilable forms to other organisms. It is noteworthy that mangrove ecosystem is
the only biotic system that most efficiently recycles sulphur.

1.3.10 Litter Decomposition & Nutrient Enrichment

Mangroves produce large amounts of litter in the form of falling leaves, branches,
and other debris. These litters are subjected to leaching of nutrients and microbial
colonization, which produce high levels of dissolved organic matter and the
recycling of nutrients both in the mangrove and in adjacent habitats. The nutrients
can potentially enrich the coastal sea and, ultimately supporting the fishery resources
(Lee et al. 1990; Benner et al. 1990; Chale 1993).
Microbes play a vital role in decomposition of mangrove litter and recycling of
nutrients. During early decomposition process, potassium and carbohydrates are
quickly leached out in a very short time. Tannins, in contrast, leach out very slowly
and the high level of tannin reduces the colonization of bacterial populations in the
initial period of decomposition. However, tannins are degraded by fungi which are
the earlier colonizers on the mangrove litter. Once the tannins are leached out and
fungally decomposed, the bacterial populations rapidly increase (Steinke et al. 1990,
1993; Rajendran 1997; Rajendran and Kathiresan 1999). The N2-fixing azotobacters
are one of the important groups in decomposing litter (Chale 1993; Wafar et al.
1997), and their activities increase the content of protein nitrogen by 2–3 times in the
litter after 1 month of decomposition (Fig. 1.7; Rajendran 1997). This protein-rich
detritus in turn, attracts shrimp, crabs, and fish.

1.3.11 Supporting Fishes and Wildlife

Mangroves are important for fish production by serving as nursery, feeding, and
breeding grounds for crabs, prawns, mollusks, and finfish. The calm waters provide
habitat for fishes, while the mangrove aerial roots, tree trunks and forest floor support
oysters, snails, barnacles, crabs, and other invertebrates. The muddy or sandy
sediments of mangroves are inhabited by a variety of epibenthic, infaunal, and
meiofaunal invertebrates. Nearly 80% of the fish catches are directly or indirectly
dependent on mangrove and other coastal ecosystems worldwide (Kjerfve and
Macintosh 1997; Kathiresan 2000; Kathiresan and Rajendran 2002). The mangroves
also support a variety of wildlife such as the Bengal tiger, dolphins, monitor water
lizard, estuarine crocodile, deer, sea turtles, monkeys, wild pigs, snakes, fishing cats,
14 K. Kathiresan

Fig. 1.7 Changes in the nitrogen–fixing azotobacter counts (1  104 g 1 leaf tissue), the total
nitrogen content (% of leaf tissue) and the juvenile prawns (no. haul 1) along with decomposing
senescent leaves of Avicennia marina

insects, and birds. Mangrove leaves, flowers, and fruits are fed in fresh condition by
insects, mollusks, crabs, and mammals. Ants protect mangrove trees from insect
grazers. A variety of species including bees, bats, and birds such as humming birds,
sun-birds, and honey eaters facilitate pollination. Many bird species use mangroves
as nesting or roosting grounds, including terrestrial and marine species (Kathiresan
and Bingham 2001; Kathiresan and Qasim 2005).

1.3.12 Supporting Coastal Food Web

Mangrove habitats contribute to complex food webs and energy transfers between
terrestrial and marine systems (Fig. 1.8). The mangroves produce decomposing
1 Mangroves: Types and Importance 15

Mangrove
Litter Fall

Leaching of Crab
Nutrients Consumption

Decomposition Faeces/Death

Land Dissolved Nutrient Retention Microbial Particulate matter Land


Run-Off Nutrients Biomass & Sediment Run-Off
Nutrient Regeneration

Phyto- Sea Benthic Small


: Epiphytes Invertebrates
plankton Grass Algae

Detritus

Larvae & Juveniles


of Fin & Shell Fishes

Immigration Emigration of
of Larvae Sub-Adults

Phytoplankton Adult prawns/


Based Material Fishes in Offshore

Fig. 1.8 A food web in a mangrove and coastal offshore ecosystems

organic matter, known as ‘detritus’ which serves as nutritious food to innumerable


finfish and shellfish especially juveniles. These detritus–feeding fishes are preyed
upon by larger carnivorous organisms. In addition to detritus, the mangrove litter
decomposition provides essential nutrients to boost the growth of microorganisms
and phytoplankton which are fed by zooplankton and fishes in mangroves and
offshore regions (Marshall 1994; Robertson and Alongi 1995; Alongi et al. 1992;
Twilley et al. 1992; Hemminga et al. 1995; Alongi 1998).
Mangroves are detritus-based system. The juvenile fishes feed directly on man-
grove detritus, small detritivorous invertebrates and on benthic microalgae growing
in the mangrove system. The sub-adult fishes migrate from mangroves into sea that
is predominant with plankton-based materials, while the juveniles immigrate from
sea into the mangrove system that is predominant with detritus-based materials. Thus
16 K. Kathiresan

the life cycle of fishes is completed with the help of mangroves and other coastal
systems.
Mangrove plants are the key primary producers, in addition to algae, benthic
algae, and phytoplankton in the mangrove forests. These forests also receive consid-
erable quantities of external organic material that are transported from upstream or
offshore ecosystems. About 10% of net productivity of mangroves is incorporated
within local sediments, while 50–60% is consumed or decomposed, and 30% is
exported to offshore regions. The nutrient export also supports productivity in
adjacent waters including benthic and pelagic systems of other coastal ecosystems
such as seagrasses and coral reefs (Kathiresan and Qasim 2005).

1.3.13 Protecting Other Coastal Systems

Mangroves provide protection to other coastal systems such as coral reefs,


seagrasses, and seaweeds. The mangroves are preventing the coastal soil erosion,
supplying the clean water after trapping soil particles, and providing the nutrient-rich
water through litter decomposition. However, when the mangroves are removed, the
loose sediment makes water turbid and not allowing required light for primary
production, thereby destroying the associated systems (Fig. 1.9).

Fig. 1.9 Influence of intact or


deforested mangroves on
seaweed, coral and seagrass
ecosystems
1 Mangroves: Types and Importance 17

1.4 Uses of Mangroves

Mangroves are of great economic value, providing the ecosystem services worth
of at least 1.6 billion US dollars per year, and supporting the coastal livelihoods
worldwide (Costanza et al. 1997). Ecosystem economic value is estimated at 91000
US dollars for one hectare of mangrove forest, which is greater than other marine
ecosystems: seagrasses, deep sea, coastal plankton, and tidal marsh (Macreadie et al.
2019). The mangroves are known to support over 70 direct human activities, ranging
from fuel-wood collection to fisheries (Dixon 1989; Lucy 2006). The economic
value is placed in a range of 2000–9990 US dollars per hectare per year, and this is
much greater than that of coral reefs, continental shelves or the open sea (Costanza
et al. 1997; Spalding et al. 2010). One hectare of mangroves is estimated to store
794 tons of CO2 equivalent for the carbon credit value of Rs. 168 per ton of CO2
storage, to protect 243 people from flood, to support the commercial fish stock of
25.3 million individuals, to yield the annual fish catch of 5.7 tons with economic
gain of Rs. 4.3 lakh (Kathiresan and Rajendran 2002; Thomas and Mark 2018).

1.4.1 Firewood and Wood Products

Mangrove twigs are used for making charcoal and firewood, due to high calorific
value. Charcoal is an energy product largely derived from Rhizophora species in
Thailand, Malaysia, Vietnam, and Indonesia. One ton of mangrove firewood is
equivalent to 5 tons of coal, and it burns producing high heat without generating
smoke. In Thailand, about 90% of the felled timber is used for charcoal production.
From Thailand and Indonesia, mangrove charcoal is exported to Singapore,
Malaysia, Hong Kong, and other Asian countries (Kathiresan 2015).
Mangrove wood is termite-resistant and durable due to its high content of tannin,
and hence the mangroves are used as timber. Pneumatophores are used to make
bottle stoppers and floats. Mangrove palm (Nypa) leaves are used to thatch roofs,
mats, and baskets (Fig. 1.10). Its sap of young inflorescence is used for sugar
production, alcohol distillation, and vinegar production. Its soft endosperm of fruits
is edible and widely used in Thailand, Indonesia, and Philippines.
Mangrove wood chips are a priced commodity for export in Indonesia and East
Malaysia. The stalks and fibres are processed into cellulose, paper, and artificial silk
(rayon) and supplied from Indonesia to Japan and Taiwan, for cellulose industrial
use. Japan has established paper mills and chipboard factories in Kalmantan and
Sumatra of Indonesia (Kathiresan 2015).
Tannin is extracted from the bark of mangroves belonging to the family
Rhizophoraceae and is used for tanning leather and fishing nets in India, Malaysia,
and Pakistan. Extracts of Bruguiera parviflora, B. gymnorrhiza, Rhizophora
apiculata, R. mucronata, and Ceriops tagal are used for tanning fishing nets and
fish traps in northern Australia.
18 K. Kathiresan

Fig. 1.10 Nypa fruticans, an economically valuable and freshwater loving mangrove palm

1.4.2 Honey Collection

Mangrove forests are a rich source of honey, facilitating apiculture activities. Indian
Sundarbans provide employment to 2000 people engaged in extracting 111 tons of
honey annually and this accounts for about 90% of honey production among the
mangroves of India (Krishnamurthy 1990). In Bangladesh, 185 tons of honey and
44.4 tons of wax are harvested each year in the western part of the mangrove forest
(Siddiqi 1997). A bulk of honey comes from Ceriops species and Excoecaria
agallocha, while the honey of best quality is produced from mangroves such as
Aegialitis rotundifolia and Cynometra ramiflora.

1.4.3 Mangrove Foliage as Fodder

Mangroves especially Avicennia are nutritive fodders to buffaloes, sheep, goats, and
camels (Fig. 1.11). It is believed that the cattle feeding on mangroves yield highly
nutritious milk. Camel grazing is very common in India, Pakistan, Persian Gulf
region, and Indonesia (Qasim 1998). Over 16,000 camels are herded into the
mangroves of Indus delta of Pakistan (Vannucci 2002). The camel herding is one
of the activities practiced by the pastoral communities known as ‘Maldharis’ in
Gujarat, India. The maldharis are shifting along with their camels to far away areas in
search of fodder for their cattle. The degradation and non-access of mangroves has
critically impacted their livelihoods.
1 Mangroves: Types and Importance 19

Fig. 1.11 Goat grazing on the leaves of Avicennia marina

1.4.4 Fisheries and Livelihood

Mangroves are known for fish production, facilitating the fishing and aquaculture
activities. Upto 80% of global fish catch is dependent on mangroves, thereby
ensuring the food security of coastal people (Ellison 2008). About 40,000 fishers
get an annual yield of about 540 million seeds of Penaeus monodon for aquaculture,
in the Sundarbans mangroves of West Bengal (Chaudhuri and Choudhury 1994).
One hectare of mangroves can yield 767 kg of wild fish and crustaceans which can
provide the annual revenue of about 11,300 US dollars, which is on par with the
most profitable intensive shrimp farming (Primavera 1991). The mangrove-rich area
provides about 70-times more catches of fish and economic gain than the mangrove-
poor area does (Kathiresan and Rajendran 2002). Shells of mangrove mollusks are
also collected for the manufacture of lime. According to coastal people, ‘No
mangroves, no prawns; no mangroves, no crabs; no forest on land, either no fish
or only fewer fish in the sea’.

1.4.5 ECO-Tourism Development

Mangroves are the attractive sites for developing a burgeoning eco-tourism espe-
cially in Southeast Asian countries. This is due to rich natural treasures associated
with the mangroves. In Malaysia, night tourism is well-developed in mangroves by
using fire-flies. In India, mangroves are increasingly attractive due to the presence of
(1) world’s largest nesting site for the Olive Ridley turtle in Gahirmatha coast of
Odisha; (2) seagrass meadows associated with the seacow (Dugong); (3) coral reefs
associated with the most beautiful ornamental fishes; (4) intertidal mudflats teeming
20 K. Kathiresan

with millions of the migratory and residential birds; and, (5) dense mangrove forest
colonized with Royal Bengal tigers.

1.4.6 Environmental Risk Reduction

The potential of mangroves in removal of atmospheric CO2 is remarkable in


mitigating the impacts of global warming and climate change. For example, Indian
mangrove forests can remove 96 million tons of atmospheric CO2 everyday which is
equivalent to the carbon credit value of 386 million US dollars in the international
market (Kathiresan 2018). The mangroves save coastal human lives, and prevent
heavy economic loss of properties during extreme weather conditions and natural
calamities such as flood, storm surges, cyclone, and tsunami (Kathiresan and
Rajendran 2005). One hectare of land with mangroves has 2 times higher cyclone-
protection cost than the selling price of mangrove ‘cleared’ land (Das 2004).
Mangroves protect groundwater aquifers from seepage of seawater, thereby ensuring
the water security of coastal population. The mangroves remove the solid and
wastewater pollution, and the heavy metals are buried deep into the mangrove soil
thereby avoiding their toxic pollution effects (Kathiresan 2018).

1.4.7 Traditional Medicinal Value

Mangrove extracts have been traditionally used as medicine. Bruguiera species are
used for reducing blood pressures and Excoecaria agallocha for the treatment of
leprosy and epilepsy. Roots and stems of Derris trifoliata are used as insecticides
and for narcotising fishes, whereas Acanthus ilicifolius is used in the treatment of
rheumatic disorders. Seeds of Xylocarpus species have antidiarrhoeal properties,
while barks of Rhizophora species have astringent, antidiarrhoea, and antiemetic
activities. Avicennia species have tonic effect, whereas Ceriops species produce
hemostatic activity. Tender leaves of Acrostichum are used as a vegetable. A
beverage is prepared from the fruits of Sonneratia species. The traditional knowl-
edge deserves scientific validation. Several mangroves are in clinical trials towards
drug developments. One such is Rhizophora racemosa in the treatment of type
2 diabetes (Tsabang et al. 2016).

1.4.8 Herbal Tea from Mangroves

The mangrove plants are rich in polyphenols that are essential ingredients for
tea-making. Catechin is a predominant group of polyphenols involved in enzymatic
oxidation by polyphenol oxidase during fermentation. This process results in the
synthesis of two major compounds of tea, namely theaflavins (flavouring chemicals)
and thearubigins (nerve stimulant chemicals). Our laboratory has developed a
protocol for making the tea from the mangrove plant, Ceriops decandra, with the
1 Mangroves: Types and Importance 21

beverage qualities similar to the conventional tea (Kathiresan 1995; Kathiresan and
Pandian 1991, 1993). Further, polyphenols such as tannins are commercially impor-
tant plant products. Gallotannins can be used in leather, medical, pharmaceutical,
food, and beverage industries (Veera Ravi and Kathiresan 1990).

1.4.9 Antimicrobial and Antioxidant Activities

Mangrove extracts can inhibit the bacterial growth. The lignin extracted from
Ceriops decandra is reported to significantly protect the mice from lethal infection
of E. coli. This antibacterial activity is due to the antioxidant property of the lignins
(Sakagami et al. 1998). In general, mangroves display a strong antioxidant activity
(Table 1.1) due to the presence of high amounts of phenolics. The antioxidants
remove radical oxygen that otherwise damages cellular biomolecules during disease
incidence.

1.4.10 Anticancer Activity of Mangroves

Mangroves are traditionally used to treat the cancers and tumours (Kathiresan et al.
2006b). Earlier reports have revealed the anticancer activities of Avicennia africana,
A. nitida, Bruguiera exeristata, and B. parviflora (Bandaranayake 1998, 2002). We
have proved that black tea from Ceriops decandra prevents the incidence of
chemically induced oral cancer in the animal model. The carcinogenic effect of
DMBA is proved in animals by well-developed squamous cell carcinoma, along
with well-defined epithelial and keratin pearls in the connective tissue with cellular
pleomorphism. However, the tumour-bearing animals when treated with mangrove
tea, exhibit no tumour but only hyperplasia. In addition, the hair loss is shown to be
caused in the animals-bearing tumour whereas the hair loss is prevented in the
animals treated with mangrove tea (Sithranga Boopathy and Kathiresan 2010;
Sithranga Boopathy et al. 2011a, b). The cancer is a serious issue with growing
threat of global warming. In this regard, we have experimentally proved that the oral
cancer increases with increasing temperature (Kathiresan and Sithrangaboopathy
2008). We have found that the callus extract of Excoecaria agallocha and Acanthus

Table 1.1 Antioxidant property of mangroves


Mangrove extract % of DPPH radical scavenging
Acanthus ilicifolius (leaf) 99.25
Excoecaria agallocha (leaf) 98.59
Avicennia marina (leaf) 75.82
Avicennia marina (flower) 57.97
Rhizophora annamalayana (leaf) 16.57
Rhizophora mucronata (leaf) 74.8
Rhizophora mucronata (flower) 45.97
22 K. Kathiresan

ilicifolius suppresses the lung cancer induced by a chemical carcinogen (benzo


(a)pyrene) in albino mice, and the plant completely cures the cancer within
16 weeks after the treatment (Singh and Kathiresan 2013, 2014).

1.4.11 Anti-Viral Activities

Polyphenols have the property of precipitating protein. Hence, the polyphenol-rich


mangroves are believed to inactivate viral proteins and pathogenic viral activity. Our
laboratory has reported that mangroves inhibit the viruses that cause human and
animal diseases such as human immunodeficiency virus (HIV), hepatitis-B-virus,
Newcastle disease virus, vaccinia virus, Semiliki forest virus, and encephalomyo-
carditis virus (Premanathan et al. 1992, 1993, 1994a, b, 1995, 1996). The members
of the family Rhizophoraceae are most inhibitory to the viruses tested (Premanathan
et al. 1992). Bioactive compounds that are responsible for anti-HIV are lignins and
acid sugars (glucose, galactose, uronic acid, arabinose, and galactosamine). The
sugar molecules protect the host cells from the virus-induced cytopathogenicity
and they block the viral antigens and their expression, thereby completely inhibiting
viral binding to the cells (Premanathan et al. 1999).
Mangroves contain limonoids with anti-viral activities. Seeds of Xylocarpus
moluccensis possess anti-viral activities against pandemic influenza A virus
(Li et al. 2015).The bioactive limonoid is krishnolide-A, effective against the
human immunodeficiency virus (HIV) (Zhang et al. 2017) and influenza A virus
(Ren et al. 2018). Thus the mangroves are promising for the treatment of incurable
viral diseases.

1.4.12 Anti-Diabetic Activity

A diabetic is abnormal condition of a high level of post-prandial glucose, insufficient


insulin, and low insulin action in humans. We have proved that the mangrove extract
of Ceriops decandra at a concentration of 120 mg/kg exhibits promising anti-
diabetic activity on par with the standard drug, glibenclamide (Nabeel et al. 2010).
This deserves further clinical studies and drug development.
Other mangroves reported to have anti-diabetic activity are Sonneratia alba,
Rhizophora mangle, R. apiculata, R. mucronata, Avicennia marina, A. officinalis,
Lumnitzera littorea, L. racemosa, Bruguiera cylindrica, Kandelia candel,
Xylocarpus moluccensis, Nypa fruticans, and Aegiceras corniculatum. Bioactive
compounds responsible for the anti-diabetic activity are complex polysaccharides,
epicatechin, flavonoids, terpenoids, saponin, tannins, quercetin, stigmasterol, and
corosolic acid (Gajula et al. 2020).
1 Mangroves: Types and Importance 23

1.4.13 Anticoagulant Activity of Mangroves

Mangrove extract prolongs the time taken for blood clotting due to the presence of
anticoagulant polysaccharides. The activity increases with increasing concentrations
of the extract from 100 to 1000 g/mL and also with increasing level of sulphate in the
extracts. We have reported the highest anticoagulant activity in Avicennia marina,
followed by Rhizophora species, Excoecaria agallocha, and Aegiceras
corniculatum (Kathiresan et al. 2006a).

1.4.14 Neuro-Protective Activity

Mangroves are efficient in inhibiting the acetyl cholinesterase enzyme (AChE),


resulting in neuro-protection. This enzyme is responsible for damaging the central
nervous system thereby causing disorders such as Alzheimer disease. The
mangroves with anti-AChE are Rhizophora lamarckii (leaf), Avicennia officinalis
(leaf), Ceriops tagal (leaf, root), Sonneratia alba (leaf, bark), Nypa fruiticans, and
R. mucronata (leaf) (Ravikiran et al. 2020).

1.4.15 Bioactive Nanoparticle Synthesis

Our laboratory has reported for the first time the ability of mangroves in synthesizing
the nanoparticles. Among the mangrove extracts, Xylocarpus mekongensis is effi-
cient, with the highest production of silver nanoparticles and others (gold, calcium,
copper, magnesium, zinc, and lead). Gallic acid is mainly responsible for the
nanoparticle synthesis, whereas glucose is required for stabilizing the nanoparticles
synthesized (Asmathunisha and Kathiresan 2013, 2018; Asmathunisha 2013). It is
also proved that the callus tissue of Sesuvium portulacastrum produces silver
nanoparticles better than intact plant tissue does (Asmathunisha et al. 2010).
The nanoparticles possess different biological activities. The silver nanoparticles,
synthesized by leaf extract of Rhizophora mucronata Avicennia marina, and Acan-
thus ilicifolius exhibit mosquito larvicidal activity against Aedes aegypti and Culex
quinquefasciatus (Gnanadesigan et al. 2012). The silver nanoparticles, synthesized
by coastal stand of Prosopis chilensis have strong antibacterial activity in controlling
the vibriosis, a common shrimp disease (Kathiresan et al. 2013b) and inhibiting the
human pathogenic bacteria (Asmathunisha et al. 2010). The silver nanoparticles are
reportedly promising in stabilizing the cotton fabrics and making them odour
resistant, preserving the apple fruits, purifying the drinking water from microbial
contaminants, detoxifying the carcinogenic ethidium bromide, and in controlling the
cancer cells (e.g. Kathiresan 2020).
Antimicrobial activity is well-known for the silver nanoparticles, synthesized by
mangroves (Rhizophora mucronata, Ceriops decandra, C. tagal, Excoecaria
agallocha, Heritiera fomes, and Sonneratia apetala). The silver nanoparticles inhibit
bacteria through damaging the structure and function of bacterial cell wall (Das and
24 K. Kathiresan

Thatoi 2020). Antioxidant activity is reported for the silver/zinc nanoparticles,


synthesized by mangroves (Avicennia officinalis, Excoecaria agallocha, Heritiera
fomes, Xylocarpus granatum, and Sonneratia apetala). This activity is due to free
radical scavenging (Das and Thatoi 2020). Anti-diabetic activity is also reported for
the silver/zinc nanoparticles, synthesized by mangroves (Avicennia officinalis,
Heritiera fomes, Xylocarpus granatum, and Sonneratia apetala). This activity is
due to inhibition of α-amylase and α-glucosidase (Das and Thatoi 2020). Anti-
inflammatory activity is also reported for the silver nanoparticles, synthesized by
mangroves (Avicennia officinalis, Heritiera fomes, Xylocarpus granatum, and
Sonneratia apetala). This activity is due to inhibition of protein denaturation (Das
and Thatoi 2020).

1.4.16 Mosquito Repellents and Larvicides

Mangrove extracts are effective in killing larvae of mosquito vectors for dengue
fever, filariasis, and malarial diseases, transmitted by Aedes aegypti, Culex
tritaeniorhynchus, and Anopheles stephensi, respectively (Thangam and Kathiresan
1988, 1989, 1991, 1992a, b, 1993a, 1994). We have shown that the extract from stilt
root of Rhizophora apiculata is much effective for mosquito larvicidal activity at
low concentrations (17–26 ppm). This activity is due to pyrethrum derivative as a
bioactive compound (Thangam and Kathiresan 1997). The mangrove extract
exhibits repellent activity against Aedes aegypti when it is applied as a paste on
human skin, and the mangrove extract also shows smoke repellency and produces
lethal effect on Culex quinquefasciatus and Aedes aegypti (Thangam and Kathiresan
1992a, b, 1993b).

1.4.17 Lead Molecules for Drug Development

Our computer-based drug discovery has identified lead molecules from mangroves to
develop drugs. The compounds are scalaradial for cervical cancer (Senthilraja and
Kathiresan 2011), triterpenoid for anti-breast cancer (Senthilraja et al. 2011),
dinitrophenylhydrazone for anti-oral cancer, heptadecanoic acid for anti-skin cancer
(Saravanakumar et al. 2012), stigmasterol as antimalarial (Senthilraja et al. 2012),
and against white spot syndrome virus (Sahu et al. 2012). These compounds are
being tested towards the development of drugs.

1.4.18 Valuable Genes from Mangroves

Mangroves are rich in genes that confer resistance to environmental stressors.


M.S. Swaminathan Research Foundation has isolated a salt-tolerant gene
(AM 244) from the mangrove species Avicennia marina, and introduced it into a
1 Mangroves: Types and Importance 25

paddy crop (Pusa Basmati and IR64) via Agrobacterium. The salt-tolerant paddy
variety may be cultivated along the coast using saltwater in the future.

1.5 Concluding Remarks

Mangrove habitats are of different types that vary in productivity and ecosystem
services. Based on coastal location, they are of three types: deltaic, estuarine, lagoon,
and fringe. Functionally, they are of six types: fringe, riverine, basin, over-wash,
scrub, and hammock. According to tidal range and sedimentation, the mangroves are
of three types: river-dominated, tide-dominated, and interior. In addition, there are of
six broad types of mangroves: large delta, tidal plains, composite plains, fringing
barriers with lagoons, drowned bedrock valleys, and coral coasts.
Mangroves are promising for bioprospecting due to the presence of structurally
new chemicals to overcome harsh coastal conditions, and valuable genes for
tolerating wind, salinity, and flood (Kathiresan and Ravikumar 2010; Kathiresan
2010, 2015). Bioprospecting the mangroves is likely to develop patents, which can
provide revenue and employment opportunities. The economic benefits are many
folds greater than the cost of plantation or conservation. There is a greater potential
for raising mangroves as ‘cash crops’ for their possible applications in food, medical,
agricultural and industrial sectors in the future.

References
Alongi DM (1998) Coastal Ecosystems Processes. CRC Press, Boca Raton, FL, USA.
Alongi DM (2007) Mangrove forests of Papua, In: Marshall AJ, Beehler B.M, (eds.) Ecology of
Papua, Part 1, Periplus Editions (HK) Ltd., pp. 824–857.
Alongi DM (2012) Carbon sequestering in mangrove forests. Carbon Manage 3: 313–322
Alongi DM (2014) Carbon cycling and storage in mangrove forests. Ann Rev Mar Sci 6: 195–219.
Alongi DM (2018) (ed.) Mangrove Forests: In blue Carbon, Springer, Switzerland, pp. 23–36.
Alongi DM (2020) Global Significance of Mangrove Blue Carbon in Climate Change Mitigation.
Sci 2(3): 57. https://doi.org/10.3390/sci2030057.
Alongi DM, Boto KG, Robertson AI (1992) In: Robertson AI, Alongi DM, (eds.) Tropical
Mangrove Ecosystems, American Geophysical Union, Washington, D.C, pp. 251–292.
Asmathunisha N (2013) Mangrove-Based Nanoparticles: Synthesis, Characterization
andApplications (Ph.D. thesis). CAS in Marine Biology, Annamalai University, India, 261 pp.
Asmathunisha N, Kathiresan K (2013) A review on biosynthesis of nanoparticles by marine
organisms. Colloids Surf B: Biointerfaces 103: 283–287.
Asmathunisha N, Kathiresan K (2018) Seasonal variation changes in biosynthesis of nanoparticles
by mangroves. Inter J Pharmacol Biosci 9B: 302–310.
Asmathunisha N, Kathiresan K, Anburaj R, Nabeel MA (2010) Synthesis of antimicrobial silver
nanoparticles by callus and leaf extracts from saltmarsh plant, Sesuvium portulacastrum
L. Colloids Surf B: Biointerfaces 79: 488–493.
Atwood TB, Connolly RM, Almahasheer H, Carnell PE, Duarte CM, Lewis CJE, Irigoien X,
Kelleway JJ, Lavery PS, Macreadie PI, Serrano O (2017) Global patterns in mangrove soil
carbon stocks and losses. Nat Clim Change 7: 523.
Bandaranayake WM (1998) Traditional and medicinal uses of mangroves. Mangroves Salt Marshes
2: 133–148.
26 K. Kathiresan

Bandaranayake WM (2002) Bioactivities, bioactive compounds and chemical constituents of


mangrove plants. Wetl Ecol Manage 10: 421–452.
Benner R, Hatcher PG, Hedges JI (1990) Early diagenesis of mangrove leaves in a tropical estuary;
bulk chemical characterization using solid-state (super 13) C NMR and elemental analyses.
Geochim Cosmochim Acta 54 (7): 2003–2013.
Bird ECF, Barson MM (1977) Measurement of physiographic changes on mangrove-fringed
estuaries coastlines. Mar Res Indonesia 18: 73–80.
Cebrain J (2002) Variability and control of carbon consumption, export, and accumulation in
marine communities. Limnol Ocean 47: 11–22.
Chale FMM (1993) Degradation of mangrove leaf litter under aerobic conditions. Hydrobiologia
257 (3): 177–183.
Chaudhuri AB, Choudhury A (1994) Mangroves of the Sundarbans Vol. I: India. IUCN, Bangkok,
Thailand, 247 pp.
Costanza R, D’Arge R, De Groot R, Farber S, Grasso M, Hannon B, Linnberg K, Naeema S,
O’Neill RV, Parvelo J, Raskin RG, Sutton P, Van den Belt M (1997) The value of the world’s
ecosystem services and natural capital. Nature 387: 253–260.
Dam Roy S, Krishnan P (2005) Mangrove stands of Andaman vis-à-vis tsunami. Curr Sci 89:
1800–1804.
Danielsen F, Sorensen MK, Olwig MF, Selvam V, Parish F, Burgess ND, Hiralshi T, Karunagaran
VM, Rasmussen MS, Hansen LB, Quarto A, Suryadiputra N (2005) The Asian Tsunami: a
protective role for coastal vegetation. Science 310: 643.
Das, Saudamini (2004) Mangroves: A natural defense against cyclones: An investigation from
Orissa, India. SANDEE Policy Brief, No. 24–07.
Das SK, Thatoi H (2020) Mangrove plant mediated green synthesis of nanoparticles and their
pharmaceutical applications: an overview In: Patra JK, Mishra RR, Thatoi H (eds.) Biotechno-
logical Utilization of Mangrove Resources Academic Press—an imprint of Elsevier.
Pp. 355–369.
Dittmar T, Hertkorn N, Kattner G, Lara RJ (2006) Mangroves, a major source of dissolved organic
carbon to the oceans. Global Biogeochem Cycles 20: 1 GB101210, doi: https://doi.org/10.1029/
2005GB002570.
Dixon JA (1989) The value of mangrove ecosystems. Tropical Coastal Area Management News-
letter 4: 5–8.
Donato DC, Kauffman JB, Murdiyarso D, Kurnianto S, Stidham M, Kanninen M (2011)
Mangroves among the most carbon-rich forests in the tropics. Nat Geosci 4: 293–297.
Duarte CM, Borum J, Short FT, Walker DI (2005) Seagrass Ecosystems: Their Global Status and
Prospects. In: Polunin NVC (ed) Aquatic Ecosystems: Trends and Global Prospects, Cambridge
Univ. Press.
Duke NC (1992) Mangrove floristics and biogeography. In: Robertson AI, Alongi DM (eds.)
Coastal and Estuarine Studies: Tropical Mangrove Ecosystems. American Geophysical
Union, Washington DC, USA: 63–100.
Duke NC (2017) Mangrove Floristics and Biogeography Revisited: Further Deductions from
Biodiversity Hot Spots, Ancestral Discontinuities, and Common Evolutionary Processes
Chapter 2 V.H. Rivera-Monroy et al. (eds.), Mangrove Ecosystems: A Global Biogeographic
Perspective, Springer International Publishing AG pp. 17–53. https://doi.org/10.1007/978-3-
319-62206-4_2.
Duke NC, Ball MC, Ellison JC (1998) Factors influencing biodiversity and distributional gradients
in mangroves. Glob Ecol Biogeogr Lett 7: 27–47.
Duke NC, Meynecke J-O, Dittmann S, Ellison AM, Anger K, Berger U, Cannicci S, Diele K, Ewel
KC, Field CD, Koedam N, Lee SY, Marchand C, Nordhaus I, Dahdouh-Guebas F (2007) A
world without mangroves? Science 317: 41–42.
Ellison AM (2008) Managing mangroves with benthic biodiversity in mind: moving beyond roving
banditry. J Sea Res 59: 2–15.
1 Mangroves: Types and Importance 27

Ewel, K, Twilley R, Ong J (1998) Different kinds of mangrove forests provide different goods and
services. Glob Ecol Biogeogr Lett 7: 83–94.
Fujimoto K (2000) Below-ground carbon sequestration of mangrove forests in the Asia-Pacific
region. Proceedings of Asia-Pacific Cooperation on Research for Conservation of Mangroves,
Okinawa, Japan, pp. 87–96.
Furukawa K, Wolanski E, Mueller H (1997) Currents and sediment transport in mangrove forests.
Estuar Coast Shelf Sci 44 (3): 301–310.
Gajula H, Kumar V, Vijendra PD, Rajashekar J, Sannabommaji T, Basappa G (2020) Secondary
metabolites from mangrove plants and their biological activities. In: Biotechnological Utiliza-
tion of Mangrove Resources (Eds. J.K. Patra, R.R. Mishra & H. Thatoi). Academic Press—an
imprint of Elsevier. pp. 117–134.
Giri C, Ochieng E, Tieszen LL, Zhu Z, Singh A, Loveland T, Masekand J, Duke N (2011) Status
and distribution of mangrove forests of the world using earth observation satellite data. Global
Ecol Biogeogr 20: 154–159.
Gnanadesigan M, Anand M, Ravikumar S, Maruthupandy M, Syed Ali M, Vijayakumar V, et al.,
(2012) Antibacterial potential of biosynthesized silver nanoparticles using Avicennia marina
mangrove plant. Appl Nanosci 2: 143–147.
Harada K, Imamura F, Hiraishi T (2002) Experimental study on the effect in reducing Tsunami by
the coastal permeable structures. Final Proc. Int. Offshore Polar Eng. Conf., USA, pp. 652–658.
Hiraishi T, Harada K (2003) Green belt Tsunami Prevention in South-Pacific Region. Greenbelt
tsunami prevention in South Pacific region. Report of the port and airport research institute 42:
1–23.
Hemminga MA, Gwada P, Slim FJ, De Koeyer P, Kazungu J (1995) Leaf production and nutrient
contents of the seagrass Thalassodendron ciliatum in the proximity of a mangrove forest (Gazi
Bay, Kenya). Aquat Bot 50 (2): 159–170.
Jardine SL, Siikamaki JV (2014) A global predictive model of carbon in mangrove soils. Environ
Res Lett 9: 104013.
Kaly UL, Eugelink G, Robertson AI (1997) Soil conditions in damaged North Queensland
mangroves. Estuaries 20(2): 291–300.
Kathiresan K (1995) Studies on tea from mangrove leaves. Environ Ecol 13(2): 321–323
Kathiresan K (2000) A review of studies on Pichavaram mangrove, southeast India. Hydrobiologia
430: 185–205.
Kathiresan K (2003) How do mangrove forests induce sedimentation?. Revista de Biol Trop 51(2):
355–360.
Kathiresan K (2010) Importance of mangrove forests of India. J Coast Environ 1(1): 11–26.
Kathiresan K (2015) Bioprospecting potential of mangrove resources. In: Kathiresan K, Ajmal
Khan S (eds.) International Training Course on ‘Mangrove Biodiversity and Ecosystems’Course
Manual. Annamalai University (Centre of Advanced Study in Marine Biology), Parangipettai,
India, 753 pp.
Kathiresan K (2018) Mangrove Forests of India. Curr Sci 114: 976–981.
Kathiresan K (2020) Bioprospecting potential of mangrove resources. In: Patra JK, Mishra RR,
Thatoi H (eds.) Biotechnological Utilization of Mangrove Resources. Academic Press—an
imprint of Elsevier. Pp.225–242.
Kathiresan K, Ravindran VS, Muruganandham A (2006a) Mangrove extracts prevent the blood
coagulate. Indian J Biotechnol 5: 252–254.
Kathiresan K, Sithranga BN, Kavitha S (2006b) Coastal vegetation an under explored source of
anticancer drugs. Indian J Nat Prod Rad 5(2): 115–119.
Kathiresan K, Bingham BL (2001) Biology of mangroves and mangrove ecosystems. Adv Mar Biol
40: 81–251.
Kathiresan K, Pandian M (1991) Effect of UV on quality of black tea from Ceriops decandra. Sci
Cult 57 (3&4): 93–95.
Kathiresan K, Pandian M (1993) Effect of UV on black tea constituents of mangrove leaves. Sci
Cult 59 (7–10): 61–63.
28 K. Kathiresan

Kathiresan K, Qasim SZ (2005) Biodiversity of Mangrove Ecosystems. Hindustan Publishing


Corporation, New Delhi, 251 pp.
Kathiresan K, Ravikumar S (2010) Marine pharmacology: an overview. Mar Pharmacol 1: 1–37.
Kathiresan K, Rajendran N (2002) Fishery resources and economic gain in three mangrove areas on
the southeast coast of India. Fish Manage Ecol 49(5): 277–283.
Kathiresan K, Rajendran N (2005) Coastal mangrove forests mitigated tsunami. Estuar coast Shelf
Sci 65: 601–606.
Kathiresan K, Sithrangaboopathy N (2008) Temperature effect on chemical induced carcinogenesis
in hamster cheek pouch. Environ Toxic Pharmacol 26(2): 147–149.
Kathiresan K, Anburaj R, Gomathi V, Saravanakumar K (2013a) Carbon sequestration potential of
Rhizophora mucronata and Avicennia marina as influenced by age, season, growth and
sediment characteristics in southeast coast of India. J Coastal Conserv 17: 397–408.
Kathiresan K, Gomathi V, Anburaj R, Saravanakumar K (2014) Impact of mangrove vegetation on
seasonal carbon burial and other sediment characteristics in the Vellar-Coleroon estuary, India. J
For Res 25: 787–794.
Kathiresan K, Nabeel MA, Gayathridevi M, Asmathunisha N, Gopalakrishnan A (2013b) Synthesis
of silver nanoparticles by coastal plant Prosopis chilensis (L.) and their efficacy in controlling
vibriosis in shrimp Penaeus monodon. Appl Nanosci 3: 65–73.
Kauffman JB, Bhomia RK (2017) Ecosystem carbon stocks of mangroves across broad environ-
mental gradients in West Central Africa: Global and regional comparisons. PLoS ONE
12, 30187749.
Kauffman JB, Adame MF, Arifanti VB, Schile-Beers LM, Bernardino AF, Bhomia RK, Donato
DC, Feller IC, Ferreira TO, Jesus Garcia MC, MacKenzie RA, Megonigal JP, Murdiyarso D,
Simpson L, Hernandez Trejo H (2020). Total ecosystem carbon stocks of mangroves across
broad global environmental and physical gradients. Ecol Monogr e01405. https://doi.org/10.
1002/ecm.1405.
Kjefve B (1990) Manual for investigation of hydrological processes in mangrove ecosystems.
UNESCO/UNDP Regional Project, 79 pp.
Kjerfve B, Macintosh DJ (1997) In: Kjerfve B, Lacerda LD, Diop, EHM (eds.) Mangrove
Ecosystem Studies in Latin America and Africa. UNESCO, Paris, pp. 1–7.
Krishnamurthy K (1990) The apiary of mangroves. In: Whigham DF, Dykyjoya D, Hejny S, (eds.)
Wetland Ecology and management: Case studies pp. 135–140. Kluwer Academic press,
Netherlands.
Lacerda LD (1998) Trace metals biogeochemistry and diffuse pollution in mangrove ecosystems.
ISME Mangrove Ecosystems Occasional Papers 2: 1–61.
Lee HL, Seleena P, Winn Z (1990) Bacillus thuringiensis serotype H-14 isolated from mangrove
swamp soil in Malaysia. Mosquito Borne Diseases Bulletin 7(4): 134–135.
Li W, Jiang Z, Shen L, Pedpradab P, Bruhn T, Wu J, et al. (2015) Antiviral limonoids including
khayanolides from the Trang mangrove plant Xylocarpus moluccensis. J Nat Prod 78 (7):
1570–1578.
Lucy E (2006) Counting mangrove ecosystems as economic components of Asia’s coastal infra-
structure. Proceedings of International Conference and Exhibition on Mangroves of Indian and
Western Pacific Oceans (ICEMAN 2006), Aug. 21–24, 2006 Kuala Lumpur. Pp. 1–14
Lugo AE, Snedaker SC (1974) The ecology of mangroves. Ann Rev Ecol Syst 5 : 39–64.
Macreadie PI, Anton A, Raven JA, Beaumont N, Connolly RM, Friess DA, Kelleway JJ,
Kennedy H, Kuwae T, Lavery PS, Lovelock CE, Smale DA, Apostolaki ET, Atwood TB,
Baldock J, Bianchi TS, Chmura GL, Eyre BD, Fourqurean JW, Hall-Spencer JM, Huxham M,
Hendriks IE, Krause-Jensen D, Laffoley D, Luisetti T, Marbà N, Masque P, McGlathery KJ,
Megonigal JP, Murdiyarso D, Russell BD, Santos R, Serrano O, Silliman BR, Watanabe K,
Duarte CM, 2019. The future of Blue Carbon science. Nat Commun 10: 3998. doi: https://doi.
org/10.1038/s41467-019-11693-w.
Marshall N (1994) Mangrove conservation in relation to overall environmental considerations.
Hydrobiologia 285 (1–3): 303–309.
1 Mangroves: Types and Importance 29

Massel SR, Furukawa K, Brinkman RM (1999) Surface wave propagation in mangrove forests.
Fluid Dyn Res 24 (4): 219–249.
Mazda Y, Magi M, Kogo M, Hong PN (1997) Mangrove on coastal protection from waves in the
Tong King Delta, Vietnam. Mangr Salt Marsh 1: 127–135.
Mcleod E, Chmura GL, Bouillon S, Salm R, Bjork M, Duarte M, Lovelock CE, Schlesinger HW,
Silliman BR (2011) A blueprint for blue carbon: toward an improved understanding of the role
of vegetated coastal habitats in sequestering CO2. Ecol Environ 9 (10): 552–560.
Medeiros CQ, Kjerfve B (1993) Hydrology of a tropical estuarine system: Itamaraca, Brazil. Estuar
Coast Shelf Sci 36: 495–515.
Moorthy P, Kathiresan K (1997a) Photosynthetic pigments in tropical mangroves: Impacts of
seasonal flux on UV-B radiation and other environmental attributes. Bot Mar 40: 341–349.
Moorthy P, Kathiresan K (1997b) Influence of UV-B radiation on photosynthetic and biochemical
characteristics of a mangrove Rhizophora apiculata Blume. Photosynthetica 34(3): 465–471.
Nabeel MA, Kathiresan K, Manivannan S (2010) Antidiabetic activity of the mangrove species
Ceriops decandra in alloxan-induced diabetic rate. J Diabetes 2: 97–103.
Ong JE (1993) Mangroves—a carbon source and sink. Chemosphere 27: 1097–1107.
Ong JE, Gong WK, Clough BF (1995) Structure and productivity of a 20-year-old stand of
Rhizophora apiculata Bl. mangrove forest. J Biogeogr 22: 417–424.
Ouyang X, Lee SY (2020) Improved estimates on global carbon stock and carbon pools in tidal
wetlands. Nat Commun 11: 317.
Pelayo M, Losada IJ, Torres-Ortega S, Narayan S, Beck MW (2020) The global flood protection
benefits of mangroves. Sci Rep 10: 1–1.
Perez A, Libardoni BG, Sanders CJ (2018) Factors influencing organic carbon accumulation in
mangrove ecosystems. Biol Lett 14: 20180237.
Premanathan M, Chandra K, Bajpai SK, Kathiresan K (1992) A survey of some Indian marine
plants of antiviral activity. Bot Mar 35: 321–324.
Premanathan M, Kathiresan K, Chandra K, Bajpai SK (1993) Antiviral activity of marine plants
against New castle disease virus. Trop Biomed 10: 31–33.
Premanathan M, Kathiresan K, Chandra K (1994a) Antiviral activity of marine and coastal plants
from India. Int J Pharmacogn 32: 330–336.
Premanathan M, Kathiresan K, Chandra K (1994b) In vitro anti-vaccinia virus activity of some
marine plants. Indian J Med Res 99: 236–238.
Premanathan M, Kathiresan K, Chandra K (1995) Antiviral evaluation of some marine plants
against Semliki forest virus. Int J Pharmacogn 33 (1): 75–77.
Premanathan M, Nakashima H, Kathiresan K, Rajendran N, Yamamoto N (1996) In vitro anti
human immunodeficiency virus activity of mangrove plants. Indian J Med Res 130: 276–279.
Premanathan M, Arakaki R, Izumi H, Kathiresan K, Nakano M, Yamamoto N (1999) Antiviral
properties of a mangrove plant, Rhizophora apiculata Blume, against human immunodeficiency
virus. Antivir Res 44 (2): 113–122.
Primavera JH (1991) Marine shrimp culture: Principles and Practices 23: 701–728.
Qasim SZ (1998) Mangroves, In: Glimpses of the Indian Ocean, (University Press, Hyderabad), pp
123–129.
Rajendran N (1997) Studies on mangrove—associated prawn seed resources of the Pichavaram,
Southeast coast of India. Ph.D thesis. Annamalai University, India.
Rajendran N, Kathiresan K (1999) Biochemical changes in decomposing leaves of mangroves.
Chem Ecol 17: 91–102.
Ramachandran S, Anitha S, Balamurugan V, Dharanirajan K, Vendhan KE, Divien MIP, Senthil
Vel A, Sjjahad Hussain I, Udayaraj A (2005) Ecological impact of tsunami on Nicobar Islands
(Camorta, Katchal, Nancowry and Trinkat). Curr Sci 89: 195–200.
Ray R, Jana TK (2017) Carbon sequestration by mangrove forest: One approach for managing
carbon dioxide emission from coal-based power plant. Atmospheric environment 171: 149–154.
Ravikiran T, Anand S, Ramachandregowda S, Kariyappa AS, Dundaiah B, Gopinath MM (2020)
Neuroprotective effects of mangrove plants. In: Patra JK, Mishra RR, Thatoi H (eds.)
30 K. Kathiresan

Biotechnological Utilization of Mangrove. Resources Academic Press—an imprint of Elsevier.


pp. 261–273.
Ren JL, Zou XP, Li WS, Shen L, Wu J (2018). Limonoids containing a C1-O-C29 moiety: isolation,
structural modification and antiviral activity. Mar Drugs 16(11): 434.
Ridd PV, Sam R (1996) Profiling groundwater salt concentrations in mangrove swamps and tropical
salt flats. Estuar Coast Shelf Sci 43(5): 627–635.
Robertson AI, Alongi DM (1995) Role of riverine mangrove forests in organic carbon export to the
tropical coastal ocean; a preliminary mass balance for the Fly Delta (Papua New Guinea).
Geo-Marine Letters 15(3–4): 134–139.
Sakagami H, Kasdhimata M, Toguchi M, Satoh K, Odanaka Y, Ida Y, et al. (1998) Radical
modulation activity of lignins from a mangrove plant Ceriops decandra (Griff.) Ding. Hou. In
Vivo 12: 327–332.
Sahu SK, Kathiresan K, Singh R, Senthilraja P (2012) Molecular docking analyses of Avicennia
marina derived phytochemicals against white spot syndrome virus (WSSV) envelope protein-
VP28. Bioinformation 8 (18): 897–900.
Sanders CJ, Mather DT, Tait DR, Williams D, Holloway C, Sippo JZ, Santos IR (2016) Are global
mangrove carbon stocks driven by rainfall? J Geophys Res Biogeosci 121: 2600–2609.
Sanderman J, Hengl T, Fiske G, Solvik K, Adame MF, Benson L, et al. (2018) A global map of
mangrove forest soil carbon at 30 m spatial resolution. Environ Res Lett 13: 055002.
Saravanakumar K, Sunil KS, Kathiresan K (2012) In-silico studies on fungal metabolites against
breast cancer protein (BRCA1). Asian J Trop Biomed B488 (3): 1–3.
Senthilraja P, Kathiresan K (2011) Computational selection of compounds derived from mangrove
ecosystem for anti-cervical cancer activity. J Recent Sci Res 2 (4): 93–98.
Senthilraja P, Kathiresan K, Saravanakumar K (2011) Comparative analysis of bioethanol produc-
tion by different strains of immobilized marine yeast. J Yeast Fungal Res 2 (8): 113–116.
Senthilraja P, Sahu SK, Kathiresan K (2012) Potential of mangrove derived compounds against
dihydrofolate reductase: an in-silico docking study. J Comput Biol Bioinforma Res 4: 23–27.
Singh CR, Kathiresan K (2013) Anticancer efficiency of root tissue and root callus of Acanthus
ilicifolius L., on Benzo pyrene induced pulmonary carcinoma in Musmusculaus. World J Pharm
Pharm Sci 2: 5271–5283.
Singh CR, Kathiresan K (2014) In vitro cytotoxicity effect of mangroves against non-small cell lung
carcinoma A549 and NCI-H522. World J Pharm Pharm Sci 3: 1067–1078.
Boopathy NS, Kathiresan K (2010) Effect of mangrove species Ceriops decandra on hair loss in the
hamster rats induced with oral cancer. Zool Surv India pp. 447–454.
Boopathy NS, Kathiresan K, Manivannan S, Jeon YJ (2011a) Effect of mangrove tea extract from
Ceriops decandra (Griff.) Ding Hou. on salivary bacterial flora of DMBA induced Hamster
buccal pouch carcinoma. Indian J Microbiol 51(3): 338–344.
Boopathy NS, Kathiresan K, Jeon YJ (2011b) Effect of mangrove tea extract from Ceriops
decandra (Griff.) on haematology and biochemical changes in DMBA induced Hamster buccal
pouch carcinogenesis. J Environ Toxicol Pharmacol 32: 193–200.
Siddiqi NA (1997) Management of Resources in the Sunderbans Mangroves of Bangladesh.
International News letter of coastal Management—Intercoast Network. Special edition 1:
22–23.
Silva CAR, Lacerda LD, Rezende CE (1990) Heavy metal reservoirs in a red mangrove forest.
Biotropica 22: 339–345.
Spalding M, Kainuma M, Collins L (2010) World Atlas of Mangroves. Earthscan, London,
Washington DC. Pp. 319.
Steinke TD, Barnabas AD, Somaru R (1990) Structural changes and associated microbial activity
accompanying decomposition of mangrove leaves in Mgeni estuary (South Africa). South Afr J
Bot 56(1): 39–48.
Steinke TD, Holland AJ, Singh Y (1993) Leaching losses during decomposition of mangrove leaf
litter. South Afr J Bot 59: 21–25.
1 Mangroves: Types and Importance 31

Tomlinson PB (2016) The botany of mangroves. Second Edition. Cambridge University Press,
Cambridge, p 418.
Thangam TS, Kathiresan K (1988) Toxic effect of mangrove plant extracts on mosquito larvae
Anopheles stephensi L. Curr Sci 47 (16): 914–915.
Thangam TS, Kathiresan K (1989) Larvicidal effect of marine plant extracts on mosquito Culex
tritaeniorhynchus. J Mar Biol Assoc India 31 (1–2): 306–307.
Thangam T, Kathiresan K (1991) Mosquito larvicidal activity of marine plants with synthetic
insecticides. Bot Mar 34: 537–539.
Thangam TS, Kathiresan K (1992a) Smoke repellency and killing effect of marine plants against
Culex quinquefasciatus. Trop Biomed 9: 35–38.
Thangam TS, Kathiresan K (1992b) Mosquito larvicidal activity of mangrove plant extract against
Aedesa egypti. Int Pest Control 34 (4): 116–119.
Thangam TS, Kathiresan K (1993a) Mosquito larvicidal activity of mangrove plants extract against
Aedes aegypti and Culex quinquefasciatus. Int Pest Control 35: 94–95.
Thangam TS, Kathiresan K (1993b) Repellency of marine plant extracts against Aedes aegypti. Int J
Pharmacogn 31: 321–323.
Thangam TS, Kathiresan K (1994) Mosquito larvicidal activity of Rhizophora apiculata. Int J
Pharmacogn 32: 33–36.
Thangam TS, Kathiresan K (1997) Mosquito larvicidal activity of mangrove plant extracts and
synergistic activity of R. apiculata with pyrethrum against Culex quinquefasciatus. Int J
Pharmacogn 35 (1): 69–71.
Thomas W, Mark S (2018) Technical report on Mangrove Restoration potential IUCN, University
of Cambridge, the Nature Conservancy, p. 33.
Tsabang N, Ngah N, Estella FT, Agbor GA (2016) Herbal medicine and treatment of diabetes in
Africa: case study in Cameroon. Diabetes Case Rep 1 (112): 2.
Twilley RR, Chen R, Hargis T (1992) Carbon sinks in mangroves and their implication to carbon
budget of tropical ecosystems. Water Air and Soil Pollut 64: 265–288.
Vannucci M (2002) Indo-west pacific mangroves. In: “Mangrove ecosystems” (Lacerda L.D ed.)
pp. 123-215. Springer-Verlag, Berlin.
Veera Ravi A, Kathiresan K (1990) Seasonal variations in gallotannin from mangroves. Indian J
Mar Sci 19: 224–225.
Wafar S, Untawale AG, Wafar M (1997) Litter fall and energy flux in a mangrove ecosystem.
Estuar Coast Shelf Sci 44: 111–124.
Wolanski E (1995) Transport of sediment in mangrove swamps. Hydrobiologia 295: 31–42.
Wolanski E, Mazda Y, Ridd P (1992) Mangrove hydrodynamics. In: Robertson AI, Alongi DM
(eds.) Coastal and Estuarine Studies: Tropical Mangrove Ecosystems. American Geophysical
Union, Washington, DC., USA, pp. 43–62.
Wong YS, Lam CY, Che SH, Li XR, Tam NFY (1995) Effect of wastewater discharge on nutrient
contamination of mangrove soil and plants. Hydrobiologia 295: 243–254.
Woodroffe C (1992) Mangrove sediments and geomorphology. In : Robertson AI, Alongi DM
(eds.) Coastal and Estuarine Studies: Tropical Mangrove Ecosystem. American Geophysical
Union, Washington DC., USA, pp. 7–41.
World Disasters Report (2002) Focus on Reducing Risk. Geneva: IFRC.
Zhang Q, Satyanandamurty T, Shen L, Wu J (2017) Krishnolides A-D: New 2-Ketokhayanolides
from the Krishna mangrove, Xylocarpus moluccensis. Mar Drugs 15 (11): 333.
Biogeography of the Mangrove Ecosystem:
Floristics, Population Structure, 2
and Conservation Strategies

P. Ragavan, K. Kathiresan, Sanjeev Kumar, B. Nagarajan,


R. S. C. Jayaraj, P. M. Mohan, V. Sachithanandam, T. Mageswaran,
and T. S. Rana

Abstract

Biogeography refers to reconstruction in the patterns of distribution of biological


diversity and to identify the processes that are responsible for those distributions
over the time. Despite the better understanding of global patterns of mangrove
distribution using improved exploration techniques, satellite cartography, or the
use of geographic information systems (GIS), the causal factors and processes
underlying such patterns are still debated. The biogeography of mangroves has
been widely discussed based on two alternative biogeographic processes, viz.,
dispersal and vicariance. However, in recent decades, human developmental
activities, climate change and extreme natural events have affected the distribu-
tion patterns of species in both the terrestrial and marine environments.

P. Ragavan (*) · S. Kumar


Physical Research Laboratory, Ahmedabad, India
K. Kathiresan
Centre of Advanced Study in Marine Biology, Faculty of Marine Science, UGC-BSR Faculty
Fellow, Annamalai University, Parangipettai, Tamil Nadu, India
B. Nagarajan
Institute of Forest Genetics and Tree Breeding, Coimbatore, Tamil Nadu, India
R. S. C. Jayaraj
Rain Forest Research Institute, Jorhat, Assam, India
P. M. Mohan
Department of Ocean Studies and Marine Biology, Pondicherry University, Port Blair, Andaman
and Nicobar Islands, India
V. Sachithanandam · T. Mageswaran
National Centre for Sustainable Coastal Management, MoEF & CC, Chennai, India
T. S. Rana
CSIR-National Botanical Research Institute, Rana Paratap Marg, Lucknow, India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 33


Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_2
34 P. Ragavan et al.

Expansion of mangroves towards the poles and recent massive mangrove dieback
in the Gulf of Carpentaria, Australia, for instance, is considered an impact of
climate change and extreme events, respectively, whereas degradation of ecolog-
ical health of existing mangroves are attributed to unsustainable human develop-
mental activities. In addition, changes in distributional limits of the mangroves
species, genetic discontinuity of population of widely distributed mangroves
resulted from the recent molecular floristic studies, and recognition of the role
of Pleistocene sea-level fluctuations, ocean currents, geomorphology, nutrient
cycling, and hydrology of estuary in shaping the present distribution and popula-
tion structure of mangrove species warrant a further comprehensive account of
the biogeography of mangroves. Understanding the changes in plant species
composition, distribution and the underlying processes are imperative for conser-
vation and management of threatened mangrove habitats, where plants are the
founder species which create habitats, modulate ecosystem functions, and support
entire ecological communities. Considering these facts, the biogeography of the
global mangroves has been examined in detail, and threats to the mangroves of
various regions are also briefly reviewed. In addition, given the ecological and
economic values of mangroves, the effectiveness of existing conservation
measures are evaluated, and the safeguards needed to maintain the ecological
health of the existing mangroves is highlighted.

Keywords

Biogeography · Conservation · Mangroves · Population Structure

2.1 Introduction

Mangroves are phylogenetically unrelated group of plants that evolved mechanisms


to adapt, disperse, and establish in intertidal areas of tropical and subtropical coast
around the world (Tomlinson 2016). Despite the pan-tropical distribution, the
mangroves are species-poor in comparison with other tropical ecosystems (Duke
2017). However, it supports rich biological diversity and offers valuable ecological
services to mankind like sediment filter, carbon sink, breeding ground for coastal
fauna and coastal defence against storm surges and tsunami (Lee et al. 2014). In the
past, it has been presumed that the water buyout propagules and seeds of the
mangrove species facilitate the long distance dispersal, therefore the resultant popu-
lation across the region are homogenous (Duke et al. 1998a, b; Maguire et al. 2000).
In contrast, recent studies on phylogeography of widely distributed mangrove
species revealed the existences of strong genetic differentiation among the popula-
tion (Guo et al. 2018a). High genetic differentiation indicates the low gene flow
between the populations and suggesting the restricted seed dispersal. Recent studies
have provided sufficient evidences that the homogenization driven by long distance
dispersal is weak due to historical barriers (continental drift and vicariant) and
contemporary barriers (vast oceanic surface distance, sea surface currents, and
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 35

geomorphology) (Guo et al. 2018a; Wee et al. 2020). In addition, in the last few
decades, many mangrove forests are rapidly getting small, fragmented, and lost due
to human developmental activities, climate change, and extreme natural events
(Feller et al. 2017; Thomas et al. 2017; Bryan-Brown et al. 2020). Population size
reduction and fragmentation cause a loss of genetic diversity and lead to the loss of
adaptability to environmental changes and subsequently accelerate the extinction
risk. Hence, understanding the changes in plant species composition, distribution
pattern, and population genetic structure of mangrove species is imperative for
conservation and management. Albeit, the biogeography of mangrove has been
dealt by several workers earlier as well, recently Saenger et al. (2019) has provided
a detailed account on biogeography of IWP mangroves considering the taxonomical
studies carried out in the last two decades (Sheue et al. 2003; Sheue et al. 2005;
Sheue et al. 2009a, b, 2010; Duke 2010; Cooper et al. 2016; Ragavan et al. 2016;
Ono et al. 2016). Furthermore, in recent times, several studies have been carried out
to understand the phylogeography of mangrove species based on molecular data
(Wee et al. 2015; Yang et al. 2016; Guo et al. 2016; Yang et al. 2017; Guo et al.
2018a, b, c). Phylogeography is a fundamental component of a core modern
biogeography (Lomolino et al. 2006) as well as an ever-expanding bridge between
biogeography and related disciplines (Riddle et al. 2008). Hence, the present text
discussed biogeography of global mangrove species with respect to their
phylogeographical pattern, and also highlighted the role of the Indian subcontinent
in diversification of Indo-West Pacific (IWP) mangroves. In addition, given the
ecological and economic values of mangroves, the effectiveness of existing conser-
vation measures are evaluated, and betterment needed to safeguard the ecological
health of the existing mangroves is highlighted.

2.2 Mangrove Floristics of the World and Distribution Pattern

Globally, the mangrove realm is largely confined to sheltered tropical and subtropi-
cal coastlines within the latitudes of around 32 N and 38 S and covers an area of
137,600 km2 spanning in 118 countries and territories (Bunting et al. 2018; Fig. 2.1).
Mangroves are sensitive to freezing and chilling temperatures; therefore, they are
most common in the tropics and subtropics (Duke 2017). The defined threshold limit
of mangroves is mean winter sea surface temperature of at least 20  C (Duke et al.
1998a). However, the mangroves are now expanding into the temperate regions of
multiple continents due to the reduced frequency and intensity of extreme freeze
events caused by climate change (Saintilan et al. 2014). About 96% of global
mangroves are confined between the Tropic of cancer (23.4 N) and Tropic of
Capricorn (23.4 S) (Bunting et al. 2018). Longitudinally there are two main centres
of mangroves, viz., Atlantic East Pacific (AEP) and Indo-West Pacific (IWP).
Mangroves of IWP represent 56% of global mangroves whereas AEP mangroves
represent 44% (Bunting et al. 2018). South-east Asia alone represents 32% of global
mangroves. Detailed global mangroves statistics are given in Table 2.1.
36

Fig. 2.1 Map showing distribution and species richness of global mangroves (reproduced based on Duke 2017)
P. Ragavan et al.
2

Table 2.1 Statistics of global mangroves (net loss and gain, protected, restorable and degraded areas)
Proportion Extent of Area of
Area of of original highly degraded Proportion
mangrove Proportion Area mangrove restorable mangrove of
Area Area protected of Area restorable areas mangrove areas mangrove
(km2) (km2) Loss Gain (km2) in protected (km2) in restorable areas (km2) in degraded
Region in 1996 in 2016 (km2) (km2) 2016 (%) 2016 (%) (km2) 2016 (%)
Australia 10,332 10,037 370 74 4553 45.4 350.9 3.3 328.6 54.6 0.5
& new
Zealand
East & 7630 7329 424 122 3112 42.5 412.0 5.3 407.0 133.0 1.6
southern
Africa
East Asia 159 159 12 13 21 12.9 7.0 4.0 6.5 2.6 1.8
Middle 334 319 19 4 100 31.3 11.4 3.3 7.9 2.7 0.8
East
North & 22,702 21,072 2196 566 12,411 58.9 2277.2 9.6 1636.3 140.2 0.7
Central
America
& the
Caribbean
Pacific 6410 6327 146 63 563 8.9 166.6 2.6 147.1 5.0 0.1
Islands
South 19,632 19,063 1106 537 13,649 71.6 1068.2 5.2 794.9 92.6 0.5
America
South 8701 8492 435 226 5428 63.9 352.7 3.9 279.7 32.4 0.4
Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . .

Asia
Southeast 46,789 44,060 3308 579 8769 19.9 3037.1 6.4 2591.2 847.0 1.9
Asia
(continued)
37
38

Table 2.1 (continued)


Proportion Extent of Area of
Area of of original highly degraded Proportion
mangrove Proportion Area mangrove restorable mangrove of
Area Area protected of Area restorable areas mangrove areas mangrove
(km2) (km2) Loss Gain (km2) in protected (km2) in restorable areas (km2) in degraded
Region in 1996 in 2016 (km2) (km2) 2016 (%) 2016 (%) (km2) 2016 (%)
West & 20,107 19,857 422 171 5317 26.8 437.1 2.1 430.5 78.5 0.4
Central
Africa
Total 142,795 136,714 8437 2356 53,923 39.4 8120.0 5.5 6629.9 1388.6 1.0
Source: Worthington and Spalding (2019)
P. Ragavan et al.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 39

At present, the updated list of true mangrove species of the world consists of
85 species (including 14 natural hybrids) in 30 genera from 17 families worldwide
(Tables 2.2 and 2.3). Despite similar areas (Saenger et al. 1983; Ellison et al. 1999)
and essentially identical physical conditions of the mangrove environments (Chap-
man 1976; Duke 1992), the IWP region harbors 72 mangrove species (including
13 natural hybrids), whereas the AEP region has 15 species, which include one
natural hybrid (Tables 2.2 and 2.3). Pelliciera benthamii is a new mangrove species
discovered from AEP, recently (Duke 2020). Natural hybrids are not discussed here
because they are usually of limited distribution, occurring within the range of the
putative parental species, and they do not offer a great deal of biogeographic
information. Only two species (Acrostichum aureum, Rhizophora samoensis) and
three genera (Acrostichum, Avicennia, and Rhizophora) have a common distribution
in both the IWP and the AEP. Except for the family Tetrameristaceae (¼
Pelliceriaceae), all the mangrove families of AEP are represented in IWP. Difference
in species richness and distinct species composition between the IWP and AEP is
attributed to complete isolation caused by the closure of Tethys seaway and regional
differences in the rate of origin of new mangrove lineages (Duke 2017). Indo-West
Pacific has a continuous presence of large areas of the continental shelf with islands
scattered among shallow tropical seas resulted by a more complex geological history
of tectonic movements, and these geomorphic features are favourable for a high rate
of origin of new mangrove lineages in IWP compared to the AEP.
Mangroves of IWP are often divided into three sub-regions, viz., East Africa,
Indo-Malesia, and Australasia. East African mangroves represent 10 mangroves
species, whereas Indo-Malesia and Australasia are represented by 55 and 43 species,
respectively. Species present in East African bioregions are widely distributed across
IWP. However, significant distributional discontinuity exists between Indo-Malesia
and Australasia. Both the regions share 39 mangrove species, but 16 mangrove
species are exclusive to Indo-Malesia, these are Acanthus volubilis, Acanthus
xiamenensis, Phoenix paludosa, Excoecaria indica, Sonneratia apetala, Sonneratia
griffithii, Brownlowia tersa, Camptostemon philippinensis, Heritiera fomes, Aglaia
cucullata, Aegiceras floridum, Aegialitis rotundifolia, Ceriops decandra, Ceriops
zippeliana, Kandelia candel, and Kandelia obovata. Whereas, only 4 mangrove
species, viz., Avicennia integra, Diospyros littorea, Excoecaria ovalis, and
Rhizophora samoensis are restricted to Australasia region. Within Indo-Malesia,
distributional discontinuity also exists between South Asia (also known as Indian
Subcontinent) and South-east Asia. All the 38 mangrove species present in South
Asian mangroves are common between both the regions, whereas 17 species are
restricted to South-east Asia, and these are Acanthus xiamenensis, Avicennia
rumphiana, Sonneratia griffithii, Sonneratia lanceolata, Sonneratia ovata,
Brownlowia argentea, Camptostemon philippinensis, Camptostemon schultzii,
Aegiceras floridum, Osbornia octodonta, Aegialitis annulata, Bruguiera exaristata,
Ceriops australis, Ceriops pseudodecandra, Ceriops zippeliana, Kandelia obovata,
and Rhizophora stylosa. Six species are endemic to Southeast Asia, and they are
Acanthus xiamenensis, Sonneratia griffithii, Camptostemon philippinensis,
Aegiceras floridum, Ceriops zippeliana, and Kandelia obovata. Ten species are
40

Table 2.2 Mangrove floristics of the World (* indicates presence of mangrove species)
Indo-West Pacific Atlantic East Pacific
East Indo- West East West
Family Genus Species Africa Malesia Australasia America America Africa
1. Acanthaceae 1. Acanthus 1. Acanthus ebracteatus * *
2. Acanthus ilicifolius * *
3. Acanthus volubilis *
4. Acanthus xiamenensis *
2. Avicennia 5. Avicennia alba * *
6. Avicennia bicolora * *
7. Avicennia germinansa * * *
8. Avicennia schauerianaa * *
9. Avicennia integra *
10. Avicennia marina * * *
11. Avicennia officinalis * *
12. Avicennia rumphiana * *
2. Arecaceae 3. Nypa 13. Nypa fruticans * *
4. Phoenix 14. Phoenix paludosa *
3. Bignoniaceae 5. 15. Dolichandrone * *
Dolichandrone spathacea
6. Tabebuia 16. Tabebuia palustrisa * *
4. Combretaceae 7. Conocarpus 17. Conocarpus erectusa * * *
8. 18. Laguncularia racemosaa * * *
Laguncularia
9. Lumnitzera 19. Lumnitzera littorea * *
20. Lumnitzera racemosa * * *
5. Ebenaceae 10. Diospyros 21. Diospyros littorea *
6. Euphorbiaceae 11. Excoecaria 22. Excoecaria agallocha * *
P. Ragavan et al.
2

23. Excoecaria indica *


24. Excoecaria ovalis *
7. Fabaceae 12 Cynometra 25. Cynometra iripa * *
13.Mora 26. Mora oleiferaa * *
8. Lythraceae 14. Pemphis 27. Pemphis acidula * * *
15. Sonneratia 28. Sonneratia alba * * *
29. Sonneratia apetala *
30.Sonneratia caseolaris * *
31. Sonneratia griffithii *
32. Sonneratia lanceolata * *
33. Sonneratia ovata * *
9. Malvaceae 16. Brownlowia 34. Brownlowia argentea * *
35. Brownlowia tersa *
17. 36. Camptostemon *
Camptostemon philippinensis
37. Camptostemon schultzii * *
18. Heritiera 38. Heritiera fomes *
39. Heritiera littoralis * * *
10. Meliaceae 19. Aglaia 40. Aglaia cucullata *
20. Xylocarpus 41. Xylocarpus granatum * * *
42. Xylocarpus moluccensis * *
11 Primulaceae 21 Aegiceras 43. Aegiceras corniculatum * *
44. Aegiceras floridum *
Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . .

12. Myrtaceae 22. Osbornia 45. Osbornia octodonta * *


13. 23.Aegialitis 46. Aegialitis annulata * *
Plumbaginaceae
47. Aegialitis rotundifolia *
41

(continued)
42

Table 2.2 (continued)


Indo-West Pacific Atlantic East Pacific
East Indo- West East West
Family Genus Species Africa Malesia Australasia America America Africa
14.Pteridaceae 24. 48. Acrostichum aureumb * * * * * *
Acrostichum
49. Acrostichum * * *
danaeifoliuma
50. Acrostichum speciosum * *
15.. 25. Bruguiera 51. Bruguiera cylindrica * *
Rhizophoraceae
52. Bruguiera exaristata * *
53. Bruguiera gymnorhiza * * *
54. Bruguiera parviflora * *
55. Bruguiera sexangula * *
26. Ceriops 56. Ceriops australis * *
57. Ceriops decandra *
58. Ceriops * *
pseudodecandra
59. Ceriops tagal * * *
60. Ceriops zippeliana *
27. Kandelia 61. Kandelia candel *
62. Kandelia obovata *
28. Rhizophora 63. Rhizophora apiculata * *
64. Rhizophora manglea * * *
65. Rhizophora samoensisb * * *
66. Rhizophora mucronata * * *
67. Rhizophora racemosaa * * *
P. Ragavan et al.
2

68. Rhizophora stylosa * *


16. Rubiaceae 29. 69. Scyphiphora * *
Scyphiphora hydrophylacea
17. 30. Pelliciera 70. Pelliciera rhizophoraea * *
Tetrameristaceae
71. Pelliciera benthamiia * *
Total 10 55 43 14 14 7
a
Present in Atlantic East Pacific mangroves
b
Present in both Atlantic East Pacific and Indo-West Pacific mangroves
Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . .
43
44 P. Ragavan et al.

Table 2.3 Natural hybrids of global mangroves


References
Natural hybrids (Parental Species)
1. Lumnitzera  rosea (L. littorea  L. racemosa) Tomlinson et al. (1978)
2. Sonneratia  gulngai (S. alba  S. caseolaris) Duke (1984)
3. Sonneratia  hainanensis (S. alba  S. ovata) Ko (1985), Wang et al. (1999)
4. Sonneratia  urama (S. alba  S. lanceolata) Duke (1994)
5. Sonneratia  zhongcairongii (Sonneratia alba  Xie et al. (2020), Zhong et al.
S. apetala) (2020)
6. Bruguiera  rhynchopetala (B. gymnorhiza  Ge (2001) and Sun and Lo
B. sexangula) (2011)
7. Bruguiera  hainesii (B. gymnorhiza  B. cylindrica Ono et al.(2016)
8. Bruguiera  dungarra (B. exaristata and Duke and Kudo (2018)
B. gymnorhiza)
9. aRhizophora  harrisonii (R. mangle  R. racemosa) Tomlinson and Womersley
(1976)
10. Rhizophora  lamarckii (R. apiculata  R. stylosa) Duke and Bunt (1979)
11. Rhizophora  annamalayana (R. apiculata  Kathiresan (1995)
R. mucronata)
12. Rhizophora  mohanii (R. mucronata  R. stylosa) Ragavan et al. (2015)
13. Rhizophora  selala (R. stylosa  R. mangle) Tomlinson (1978)
14. Rhizophora  tomlinsonii (R. apiculata  R. mangle) Duke (2010)
Unnamed natural hybrids
1. Avicennia marina  A. rumphiana Huang et al. (2014)
2. A. germinans  A. bicolor Mori et al. (2015)
3. Sonneratia alba  S. griffithii Qiu et al. (2008)
4. Acrostichum aureum  A. speciosum Tsai et al. (2012)
5. Ceriops tagal  C. australis Zhang et al. (2013)
a
Hybrid confined to the Atlantic East Pacific only

not found in Australasia but are shared between South-east Asia and South Asia, and
these are Acanthus volubilis, Phoenix paludosa, Excoecaria indica, Sonneratia
apetala, Brownlowia tersa, Heritiera fomes, Aglaia cucullata, Aegialitis
rotundifolia, Ceriops decandra, and Kandelia candel. Mangroves of AEP are
divided into three sub-regions, viz., West America, East America, and West Africa.
All the known AEP mangrove species are present in West and East American
sub-regions, whereas only 8 mangrove species are present in the West African
region. In general, longitudinally mangrove species richness increases towards
east, whereas latitudinally the species richness increases towards equator.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 45

2.3 Phylogeography of Extant Mangrove Species

Recent phylogeographical studies revealed the low variation within populations, but
high differentiation among populations of widely distributed mangrove species (Yan
et al. 2016; Guo et al. 2018a), which is in contrast to the expectation that long-lived
woody species maintain more variation within species and within populations but
have less variation among populations (Hamrick et al. 1992). The major cause of the
low genetic diversity within populations and within species, especially in marginal
populations of the species’ distribution ranges are repeated founder effects or
bottleneck events associated with geo-climatic history in IWP region and genetic
drift in small and geographically isolated populations (Yan et al. 2016). Further-
more, these studies reveal the role of barriers of gene flow, viz., geographic distance,
glacial vicariance (land masses), and ocean currents in shaping the distribution
pattern of extant mangrove species. The genetic differentiation with respect to
geographical distance is minimal in most of the mangrove species due to their
potential of long distance dispersal, whereas the levels of gene flow between
populations in most of the mangrove species are determined by land barriers, vast
oceanic surface distance, and directionality and dynamic interactions of wind and
current, acting alone or in synergy (Lo et al. 2014; Wee et al. 2014; Wee et al. 2020).
In the Indo-West Pacific (IWP) region, genetic differentiation between
populations has been reported in many mangrove species on both sides of the
Malay Peninsula (¼Sunda shelf) and of the Wallaces. The Malay Peninsula is a
land barrier associated with the genetic break between the Indian Ocean region
(referring to the coasts located on the west of the Malay Peninsula) and Southeast
Asian region (referring to the coasts located on the east of the Malay Peninsula). This
distinctive genetic break across the Malay Peninsula has been observed in many
mangrove species, including Lumnitzera racemosa (Su et al. 2006; Li et al. 2016),
Lumnitzera littorea (Su et al. 2007), Ceriops tagal (Ge and Sun 2001; Liao et al.
2007), Ceriops decandra (Tan et al. 2005; Huang et al. 2008), Sonneratia alba (Wee
et al. 2017; Yang et al. 2017), Bruguiera gymnorhiza (Minobe et al. 2010; Urashi
et al. 2013; Ono 2016), Rhizophora apiculata (Inomata et al. 2009; Wee et al. 2014;
Yan et al. 2016), Xylocarpus granatum (Tomizawa et al. 2017; Guo et al. 2018c),
Excoecaria agallocha (Zhang et al. 2008; Guo et al. 2018b), Xylocarpus
moluccensis (Guo et al. 2018c), Acanthus ilicifolius (Guo et al. 2020) and Heritiera
littoralis (Banerjee et al. 2020). During the glacial periods, falls in sea levels caused
the Malay Peninsula, Sumatra, and Java to become connected and form the Sunda
Land, which halted the exchange of seawater between the Indian Ocean and the
South China Sea (Wyrtki 1961; Wang et al. 1995). This glacial isolation resulted in
genetic divergence between populations from these two oceans. Despite the promi-
nent gene flow barrier, recent studies identified the Malay peninsula as a corridor for
genetic exchange between the oceanic regions, especially during interglacial periods
of the Pleistocene when the land shelves submerged (Wee et al. 2014; Yang et al.
2017). A study on the contemporary gene flow across the Malay Peninsula has
revealed that species with higher dispersal potential (Bruguiera gymnorhiza and
Rhizophora mucronata) exhibit much higher proportion of recent inter-population
46 P. Ragavan et al.

migration along the Malacca Strait than the species with lower dispersal potential
(Avicennia alba and Sonneratia alba) (Wee et al. 2020). Thus, the genetic differen-
tiation in Rhizophora mucronata is found only at the edge of the Andaman Sea and
the Strait of Malacca (Wee et al. 2014).
The genetic break across Wallaces has been observed in Ceriops tagal (Huang
et al. 2008), Lumnitzera racemosa (Li et al. 2016; Su et al. 2006), L. littorea (Su et al.
2007), Rhizophora stylosa (Wee et al. 2014; Yan et al. 2017), Rhizophora apiculata
(Guo et al., 2016; Yan et al. 2016), Sonneratia caseolaris (Yang et al. 2016), S. alba
(Yang et al. 2017), Xylocarpus granatum (Tomizawa et al. 2017; Guo et al. 2018c),
and Heritiera littoralis (Banerjee et al. 2020). Plate motions of the Indo-Australian
Archipelago, the emergence of the Sahul shelf during glacial time and Ocean
currents have been identified as barriers for gene flow between Southeast Asia and
Australasia/Oceania. However, it is seldom considered to be the strongest genetic
break except for R. apiculata and R. stylosa. Genetic differentiation is distinct
between South-East Asian (Japan, Vietnam, Philippine, and Indonesia) and
Oceanian (Fiji, Vanuatu, and New Caledonia) in the populations of Rhizophora
stylosa (Wee et al. 2015), and Rhizophora apiculata (Guo et al. 2016). As the
locations of these genetic breaks correspond to the boundaries of oceanic currents,
Wee et al. (2015) suggested that oceanic circulation patterns might have acted as
“cryptic barriers”. In contrast, Banerjee et al. (2020) and Guo et al. (2018c) reported
only a weak genetic differentiation between South-east Asian and Australasia
populations of Heritiera littoralis and Xylocarpus moluccensis, respectively, and
noted the existences of considerable gene flow between these two regions being
mediated by the Indonesian throughflow which moves from the Celebes Sea through
the Makassar Strait to the Timor Sea.
In addition, several studies have also reported the genetic structure within each
oceanic region. For instance, within Oceania, a strong genetic differentiation is
found in the populations of Sonneratia alba and Rhizophora apiculata between
Australia and South-West Pacific Islands (Wee et al. 2017; Guo et al. 2016). Within
South China Sea, the genetic differentiation is recorded between mainland and
Island coast populations of Sonneratia alba (Wee et al. 2017) and Northern and
Southern SCS populations of Heritiera littoralis (Banerjee et al. 2020). Similarly,
within the Indian Ocean, genetic differentiation occurs between the Arabian Sea and
Bay of Bengal populations of Bruguiera gymnorhiza (Urashi et al. 2013) and
between African and Asian populations of B. gymnorhiza (Ono 2016). Oceanic
currents and the vastness of the Ocean/sea are often attributed to the genetic
differentiation within the oceanic region. In Atlantic East Pacific (AEP) genetic
differentiation is observed between populations of the Atlantic and Pacific Oceans in
all studied mangrove species and it is attributed to vicariance following the final
closure of the Central American Isthmus (Takayama et al. 2013; Cerón-Souza et al.
2015).
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 47

2.4 Origin and Diversification of Mangroves Based


on the West Tethyan Hotspot

Mangroves evolved from terrestrial rather than marine habitats. This is proved by the
presence of mangrove pollen fossils in India below marine foraminiferan
assemblages (lower deposits of estuarine environments) suggesting the mangrove
evolution from a non-marine habitat and adapted to an estuarine habitat (Srivastava
and Binda 1991). There are two hypotheses available for the origin of mangroves,
viz., centre-of-origin and vicariance hypothesis. According to the centre-of-origin
hypothesis (van Steenis 1962; Chapman 1975, 1976), the entire mangrove taxa first
appeared in the Indo-West Pacific and subsequently dispersed to other regions.
According to the vicariance hypothesis (McCoy and Heck 1976; Mepham 1983),
all mangroves originated around the Tethys Sea, and continental drift then isolated
the flora in different regions of the earth. In addition, Specht (1981) proposed an
Australasian centre-of-origin hypothesis mainly based on pollen observations made
by Churchill (1973). However, vicariance hypothesis has been widely accepted by
many subsequent researchers (Tomlinson 1986; Duke 1995; Duke et al. 1998a;
Ellison et al. 1999; Kathiresan and Bingham 2001). In the past, three possible routes
have been suggested to explain the modern distribution of mangroves, viz., (1) East-
ward dispersal—initial diversification in IWP and dispersal eastward across Pacific
into AEP, (2) Westward dispersal—initial diversification in IWP and dispersal
westward through Tethys seaways into the Atlantic, (3) Connection between the
IWP and AEP regions around the southern tip of African continent (Ellison et al.
1999; Srivastava and Prasad 2019). All the three routes cannot be dismissed,
although there is no adequate fossil evidence to support it. Further, all these
hypothetical routes emphasize the initial colonization occurred primarily in South-
east Asia/Malaysia and dispersed to other regions; poor dispersal abilities and the
closure of the Tethys connection to the Atlantic is often attributed to the restriction of
most mangrove taxa to the IWP. Due to the presence of high species richness and the
prevalence of equable climatic conditions since the end of the Cretaceous, the South-
East Asia is believed to be the place of origin for angiosperms which first acquired
the mangrove habitat, from where most contemporary mangrove genera originated.
However, the oldest South-East Asian mangrove fossils are often less older than
equivalent ones from distant sites (Mepham 1983). For instance, fossil records of
Rhizophora and Sonneratia are known from the Paleocene of India and France, but it
appears in Borneo during Middle Eocene and Lower Miocene, respectively (Ellison
et al. 1999). Thus a consistent hypothesis accounting for the distribution of modern
mangrove species is far from complete.
The combined evidence from fossil and molecular studies revealed that the
biodiversity hotspots occurred at different places through time and there have been
at least three marine biodiversity hotspots during the past 50 million years (Renema
et al. 2008). Recent paleobiogeographical studies based on the re-examination of
fossil evidences revealed that mangroves were at their prime during the geological
period of Eocene rather than in the Late Cretaceous (Srivastava and Prasad 2019).
Particularly, the records of mangrove fossils of recognized mangrove genera like
48 P. Ragavan et al.

Nypa, Acrostichum, Rhizophora, Avicennia, Pelliciera, Sonneratia, Bruguiera,


Ceriops, Heritiera, and Aegiceras were present around the Tethyan shoreline during
Eocene at the convergence zone between the African and European continents
(Ellison et al. 1999; Srivastava and Prasad 2019). This supports the existences of
Eocene West Tethyan and Arabian hotspots prior to Southeast Asian hotspot. Thus,
it is apparent that rich biodiversity of South-east Asia is not a unique feature, but is
the latest manifestation of a pattern that was present in the past (Renema et al. 2008).
Further, it is also plausible that the initial colonization of mangroves with high
species richness occurred in west tethyan hotspot and dispersed west to the
Americas, as well as south and east, along the coasts of Asia, East Africa, Malagasy,
and Greater India. Subsequent continental drift resulted current distribution pattern
of mangroves.

2.4.1 Founder Effect Speciation- Causes for Genotypically


and Phenotypically Distinct AEP Population

Birth and death of successive hotspots are attributed to environmental changes


caused by the geological events such as Plate tectonics (Renema et al. 2008).
Duke (2017) recognized the role of four geological events in shaping the current
distribution pattern of mangroves, and they are (1) the separation of Africa and South
America (~100 mya) and progressive opening of the South Atlantic Ocean,(2) the
closure of the Tethys Sea between Africa and Eurasia 25–35 mya, (3) the separation
of India (70–75 mya) and Australia (~50 mya) with their subsequent collisions, and
(4) the opening of the North Atlantic (from ~60 mya onwards). Of these, except
event 3, others are responsible for the formation of genotypically and phenotypically
distinct AEP population. Collision of Africa and Eurasia had resulted in contraction
of West tethyan hotspot and subsequently terminated the tethyan seaway. As a result,
the populations of AEP remain isolated from parental populations and experience the
founder event along the coast of West Africa and South and North America.
Subsequent widening of the Atlantic, lack of gene flow and regional climate change
are attributed to the strong genetic drift in the founder population leading to the
evolution of genotypically and phenotypically distinct AEP population through
founder effect speciation. Lack of speciation in AEP species can be attributed to
the prevalence of dry tropical climate of the New World tropics since Miocene
(Mepham 1983). Wet climates support terrestrial-mangrove transition, but a marked
drying trend in the AEP region as evident from the extinction of Nypa. The range
contraction of Pelliciera since the Miocene also suggested the lack of suitable
environmental conditions to support the terrestrial-mangrove transition. Whereas
in IWP, the continuous presence of extensive wet habitat since the end of the
Cretaceous favours the terrestrial-mangrove transition. Thus, diversity continues to
increase in IWP through invasion of new clades and autochthonous production of
new taxa. Predictably, many more nonexclusive taxa of trees invade mangrove
habitat from terrestrial habitats in the IWP region than in the AEP region.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 49

2.4.2 Indian Subcontinent View of Diversification of IWP


Mangroves

In IWP, the collision of India and Australia with Asia would be the key tectonic
movement (Duke 2017). Further, this collision altered the geomorphology of South-
East Asia into massive fragmentation of landmasses with shallow shelf areas, which
are highly suitable for colonization of mangroves. Indian continent is the first one,
separated from Gondowana around 90–80 mya and move northward towards Asia in
tropical latitudes about 40 mya and collision with Asia began around 40 mya. This
complete isolation of India for 40 mya and vast continental shelf area, particularly
the Northern and North-eastern parts, around the Indian Island would be highly
suitable for terrestrial-mangrove transition/colonization and also for speciation. In
the past, India is seldom considered as part of tethyan shore, but actually Indian
subcontinent was an important part of the ancient Tethyan shore line (particularly
North and North-eastern parts) as wells as the part of Eocene west Tethyan and
Arabian hotspot (Renema et al. 2008). So it is more obvious that the Indian
subcontinent could be refugia of ancestral population of IWP rather than that of
South-East Asia. For instance, when Western tethyan hotspot matured in the late
Eocene, collision of India with Asia began, which would lead to the creation of
numerous shallow marine platforms in the Northern and North-eastern India. Further
the rifting of the Mediterranean also would create shallow areas in the Middle East.
Due to this contraction of west tethyan hotspot, the population of west tethyan
hotspot migrated toward east and colonized in adjacent suitable habitats, viz.,
Arabian hotspot and Indian coast. This is evident by the occurrence of mangrove
fossils since late cretaceous period (Ellison et al. 1999), the rich mangrove fossils in
India during Eocene (Srivastava and Prasad 2019) and the presence of more than
60% of extant species in Indian subcontinent. However, the role of the Indian
subcontinent in the current distribution pattern of extant mangrove species has
been overlooked in most of earlier bio-geographical distribution. Furthermore,
recent studies on phylogeography of widely distributed mangroves of IWP, mostly
excluded the sampling from Indian subcontinent, therefore, it is very important to
highlight the significance of the Indian Subcontinent in the current distribution
pattern of mangroves in IWP.
The mangroves of IWP consist of 59 mangrove species, which excludes 13 natural
hybrids. India subcontinents host 38 mangrove species, whereas South-East Asia
and Australasia host 55 and 43 species, respectively. Out of the 59 mangrove species
of IWP, 21 species are not known from the Indian subcontinent, of which, many
species are recently diverged from their sister/closely related species (Table 2.4).
Recent phylogeographical studies revealed that the divergence time of these species
is within the last 3 Myrs, and cycles of isolation interspersed by episodes of gene
flow during Pleistocene had resulted in this speciation in South-East Asia—a
boundary between the Indian Ocean and Pacific Ocean (He et al. 2019). Further,
recent molecular studies have revealed that the gene flow between the Indian and
Pacific Ocean is asymmetrical and eastward gene flow from North Indian Ocean to
South China Sea and Australia is stronger than that in the opposite direction (Guo
50 P. Ragavan et al.

Table 2.4 Recently diverged mangrove species and their close relatives
Species Remark
1. Acanthus xiamenensis Restricted distribution. Moreover identity is uncertain
2. Avicennia integra Recently diverged from A. officinalis
3. Avicennia rumphiana Recently diverged from A. alba
4. Diospyros littorea Independent terrestrial-mangrove transition (endemic to Australia)
5. Excoecaria ovalis Subspecies of E. agallocha
6. Sonneratia griffithii Though not found in SA. Possibilities for its occurrence are high.
Known from Bruma
7. Sonneratia lanceolata Recently diverged from S. caseolaris
8. Sonneratia ovata Recently diverged from S. caseolaris
9. Brownlowia argentea Recently diverged from B. tersa
10. Camptostemon Independent terrestrial-mangrove transition (endemic to South-East
philippinensis Asia)
11. Camptostemon Independent terrestrial-mangrove transition (known in South-East
schultzii Asia and Australia)
12. Aegiceras floridum Recently diverged from A. corniculatum
13. Osbornia octodonta Independent terrestrial-mangrove transition (known in South-East
Asia and Australia)
14. Aegialitis annulata Recently diverged from A. rotundifolia
15. Bruguiera exaristata Recently diverged from B. gymnorhiza
16. Ceriops australis Recently diverged from C. tagal
17. Ceriops Recently diverged from C. decandra
pseudodecandra
18. Ceriops zippeliana Recently diverged from C. decandra
19. Kandelia obovata Recently diverged from K. candel
20. Rhizophora Known in west pacific islands. Identity is uncertain
samoensisc
21. R. stylosa Recently diverged from R. mucronata

et al. 2016; Li et al. 2016). Particularly in summer, due to the effects of the strong
south-west monsoon and the absence of the north-east trades, strong current flows
from west to east, which facilitates the west to east migration rather than that in
opposite direction. Further, summer monsoon current completely obliterates the
north equatorial current and hence, there is also no counter-equatorial current, and
makes eastward migration more prominent.
All mangrove species of the Indian subcontinent, except Lumnitzera littorea, are
known from Sundarbans (considering both Indian and Bangladesh). In the past, the
formation of Bengal Basin was viewed as a result of the collision of India and
Eurasia, and the formation Sundarban delta is considered most recent. If the
Sundarban is of most recent origin, how does it host about 60% of extant mangrove
species of IWP. Are they dispersed from South-East Asia? This chance for dispersal
from South-East Asia to Sundarban is very rare considering the very rare long
distance dispersal of mangroves species and barrier of Sundaland during Pleistocene
climate oscillation. Though the Bengal basin is widely considered as Foreland basin
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 51

formed via continent–continent collision, there is speculation that the formation of


Bengal basin is not a recent one, and it began during the cretaceous period (Alam
et al. 2003), viz., formation of the Bengal basin started in the Jurassic with initiation
of rifting of the Pangaea and completed by the Miocene with docking of eastern
India with Burma Plate. Collision of India with Eurasia resulted in expansion of
Bengal basin due to huge sediment deposition from Himalayan Rivers. The presence
of rich mangrove fossils of Bengal basin during Eocene supports the occurrence of
ancestral populations along the Indian coast (Venkatachala and Rawat 1972; Mathur
and Chowdhary 1977; Kar 1985; Thanikaimoni 1987; Mathur 1984). India was the
part of West Tethyan and Arabian hotspot (Renema et al. 2008), and hence, it is more
plausible that Indian subcontinent could be the refugia of ancestral population of
IWP mangroves. Climate /geo-morphological changes that resulted from the colli-
sion of India with Eurasia would be attributed to the extinction of few mangroves
(e.g. Nypa) along west coast. However, formation of deltas along the east coast
would have enhanced the colonization of ancestral populations. Water buoyant
propagules of mangroves are suitable for long-shore dispersal, especially of vivipa-
rous species rather than transoceanic dispersal. Thus the formation of contiguous
coastline from India to SEA and seasonally directed eastward monsoon current as
well as East Indian Coastal current would have facilitated dispersal from the Indian
subcontinent to SEA prior to the Pleistocene. The collision of India and Australia
with Asia resulted in massive fragmentation of landmasses and shallow continental
shelf areas which would have facilitated the rapid colonization in SEA. Further,
cycles of isolation interspersed by episodes of gene flow during Pleistocene sea-level
fluctuation resulted in speciation event in SEA (He et al. 2019).
Collision of Australia with Asia began 10 mya and it separated from Antarctica
around 40 mya. So Australia spent more time in the temperate zone before its
collision with Asia. Thus, it is widely accepted that Australia had gained floristic
diversity when it came close to the SEA. The Sahul shelf formed due to collision of
Australia with SEA prevents gene flow between SEA and Australia during glacial
periods; but when the collision began, Australia was close to SEA, which might have
facilitated migration of mangrove species from SEA to Australia. It is particularly
true with Indonesia, where current flow was strong enough to facilitate gene flow.
Recent phylogeographical studies also reported the weak genetic differentiation of
many mangrove species due to the presence of gene flow between South-East Asia
and Australia facilitated by Indonesian throughflow current. Phylogeographical
studies also revealed the close genetic similarity of east African population of
Bruguiera gymnorhiza, Sonneratia alba, and Xylocarpus granatum with south-
east Asian populations and with Rhizophora mucronata with Australian population.
Further, a strong gene flow is found to be possible from Southeast Asia/Australia to
East Africa. Hence, the ancestors of African populations might have emigrated from
Southeast Asian/Australian populations, via major ocean currents (South equatorial
currents) (Lo et al. 2014). There are seven hypothetical dispersal routes have been
proposed for IWP mangroves (Duke 2017). Of which, IWP south-east and IWP
North-East routes seem to be more prominent considering the existing ocean current
pattern as supported by many phylogeographical studies. Based on above
52 P. Ragavan et al.

Fig. 2.2 Map showing dispersal route based on Indian Subcontinent view (1) East Indian Coastal
current, (2) Summer monsoon current (3) South China Sea Warm Current (4) Indonesia
throughflow current (5) South equatorial current (6) West Indian Coastal Current

speculation, more plausible dispersal routes in IWP are proposed here (Fig. 2.2).
However, it needs to be verified in future in the present context of very little fossil
evidences and lack of phylogeographical information on mangrove species of Indian
subcontinent and Middle East.

2.5 Conservation and Management

Despite the slowdown in average rate of mangrove loss in recent decade, still
mangroves experience an annual loss of 0.2–0.7% between 2000 and 2012, and
remain the most threatened ecosystems of the world (Hamilton and Casey 2016).
Anthropogenic activities and climate change consequences are the potential threat
for mangroves. Overexploitation, land use change, hydrological alteration, and
environmental pollution are the contemporary anthropogenic factors affecting man-
grove distributions, whereas sea-level rise, increasing frequency of natural
calamities such as drought, storm, cyclones, tsunami are the climate related threat
factors. Recently Goldberg et al. (2020) estimated that about 3363 km2 (2.1%) of
global mangrove areas was lost between 2000 and 2016, at an average annual rate of
0.13%. Anthropogenic causes are accounted for 62% of total mangrove loss,
whereas natural causes are attributed to 38% of total losses. Habitat conversion for
rice, shrimp, and oil palm cultivation remains the primary global driver of mangrove
loss, which represents 47% of global losses from 2000 to 2016; land reclamation for
non-productive conversions and human settlements represents 12 and 3% of global
losses, respectively. Shoreline erosion resulted from the increasing sea-level rise
represent the second highest as well as primary natural causes of global mangrove
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 53

losses, which represent 27% of global mangrove loss from 2000 to 2016;
while extreme weather events (EWE) contribute to 11% of losses (Goldberg et al.
2020). In order to reverse the mangrove loss, various legislative measures and
massive rehabilitation/restoration efforts have been undertaken in the last three
decades. Declaration of protected areas is a major legislative measure in many
mangrove nations. Worldwide, there are about 2500 protected areas that include
mangrove forests within their boundaries, which represent over 39% (around
54,000 km2; Table 2.1) of world’s remaining mangroves; and, 34 countries have
placed more than half of their mangroves under such protection (Worthington and
Spalding 2019; IUCN and UNEP-WCMC 2019). An extensive coverage of
mangroves under protected areas represents a strong positive trend in coastal con-
servation However, Southeast Asia- a hotspot of mangroves in terms of area cover,
species diversity, and deforestation rate—has only 20% area under protection
(Table 2.1). Protected areas prevent some drivers of degradation, such as unsustain-
able timber extraction. However, other drivers of degradation, such as upstream
water abstraction or changes to sediment supplies, cannot be influenced when they
occur beyond boundaries of the protected area (Worthington and Spalding 2019).
So, considering the hydrological and ecosystem connectivity is imperative to mini-
mize the ecological degradation within the protected areas.
In the last three decades, massive efforts were globally made to restore the
degraded mangrove areas. Incorporation of mangroves into engineered hard coastal
defence structures, monoculture plantations, and “ecological mangrove restoration”
(EMR) approaches are common methods of mangrove restoration (Ellison et al.
2020). However, despite existing guidance for successful restoration efforts (Lewis
et al. 2019), most of the efforts are not successful (Lee et al. 2019) and successful
cases are rare (e.g. Stubbs and Saenger 2002; Lewis et al. 2005; Bosire et al. 2008;
Rey et al. 2012; Dale et al. 2014; Brown et al. 2014a, b; Begam et al. 2017).
Unsuccessful restoration is a waste of time, money, and human resources. Further,
the habitat characteristics altered for the purpose of mangrove restoration render the
natural corridors unsuitable for natural migration of mangroves to cope-up with the
impacts of climate change. Further, functionality of restored areas is rarely moni-
tored to fully ascertain the restoration success based on faunal diversity, vegetation
structure (e.g., basal area, species diversity), and function (e.g., net primary produc-
tivity, carbon storage, resilience) (López-Portillo et al. 2017). The costs of mangrove
restoration are around US$3000 per hectare (Bayraktarov et al. 2016), however, the
cost is high over US$100,000 per hectare when large-scale engineering to restore
hydrology, combined with high staffing (Worthington and Spalding 2019). So,
mangrove restoration is expensive for many developing nations’ especially South-
East Asian countries. Thus, investments in large-scale planting must be the last
option. Rehabilitation of abandoned aquaculture areas should be directed to the
hydrological correction (Lewis et al. 2016) to facilitate natural recruitment of
mangroves and also improving soil fertility. Recently, Ellison et al. (2020) proposed
the adaptive management of restored mangroves areas—a structured, iterative
process of “learning-by-doing” and decision-making in the face either of continu-
ous change (environmental, social, cultural, or political) or uncertainty, through
54 P. Ragavan et al.

regular monitoring of key indicators of the objectives and goals of a restoration


project and identifying the clear triggers or decision-points for appropriate interven-
tion and action if the objectives or goals are not being met.
As a result of last three decadal conservation measures, human-driven mangrove
loss decreased significantly since 2000. For instance, the area of mangroves
converted by direct human intervention has declined by 73% over the 16-year period
(viz., from 1186 km2 between 2000 and 2005 to 314 km2 between 2010 and 2016)
and during the same period, the total loss of mangroves area is attributed to natural
causes declined by 60% from 624 km2 between 2000 and 2005 to 249 km2 between
2010 and 2016 (Goldberg et al. 2020). The decreasing rate of natural loss (3.75% per
year) was significantly lower than that of anthropogenic loss (4.56% per year);
however, the relative contribution of natural drivers to global mangrove losses
increased by 10% over the 16-year period (2000–2016) (Goldberg et al. 2020).
Shoreline erosion and extreme weather events are the key drivers of natural losses of
mangroves. So it is certain that the emergence of natural drivers as the primary
causes of modern mangrove loss in upcoming decades. For example, the recent
massive mangrove dieback in the Gulf of Carpentaria, Australia, for instance, is
considered to be an impact of climate change-induced extreme weather events (Duke
et al. 2017; Lovelock et al. 2017).
Despite being threatened by human developmental activities and climate change,
the mangrove forests are highly resilient ecosystems that have the potential to adapt
and adjust to changing conditions (Woodroffe et al. 2016). This is evident by the
recent expansion of mangroves towards the polar regions in response to increasing
minimum winter temperature (Saintilan et al. 2014) as well expansion and contrac-
tion in response to temperature fluctuations and Pleistocene sea-level drop and rise in
the long past (Ludt and Roacha 2014). Vertical adjustment and horizontal movement
across the landscape are the processes that govern the responses of mangrove forest
to sea-level rise (Woodroffe et al. 2016; Krauss et al. 2014; Lovelock et al. 2015).
Furthermore, the observed higher expression of diversity than the genetic diversity
and smaller genome size warrants the significant evolutionary potential of
mangroves (Lira-Medeiros et al. 2010; Wee et al. 2018; Lyu et al. 2018). So
enhancing the adaptive potential of mangroves is the need of the hour to cope-up
with the climate change. Since, widely distributed mangroves species exhibit low
genetic diversity and high genetic differentiation among the populations across their
distribution ranges, it is imperative to delineate the Conservation Units, including
Evolutionarily Significant Units (ESUs) and Management Units (MUs), which
warrant management as a separate unit (Frankham et al. 2010) Thus, use of genetic
information in conservation is crucial for its long-term effectiveness to preserve the
adaptive and evolutionary potentials of ecosystem/species (Hoffmann and Sgro
2011). However, little has been translated into on-the-ground conservation (Wee
et al. 2018). Further, structure and functions of mangrove are highly site-specific,
and they even differ significantly along the same coast. The mangrove ecosystem
responses to climate change processes are expected to be greatly influenced by plant-
mediated processes. So the site-specific and species-specific knowledge of
mangroves is highly imperative to enhance the sustainability of mangroves.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 55

Acknowledgement The author (P.R.) is thankful to the Department of Space, Government of


India for funding under Post-Doctoral Fellowship at Physical Research Laboratory and the
authorities of the Physical Research Laboratory for providing facilities.

References
Alam M, Alam MM, Curray JR, Chowdhury ALR, Gani MR (2003) An overview of the sedimen-
tary geology of the Bengal Basin in relation to the regional tectonic framework and basin-fill
history. Sediment Geol 155: 179–208.
Banerjee AK, Guo W, Qiao S, Li W, Xing F, Lin Y, Hou Z, Li S, Liu Y, Huang Y (2020) Land
masses and oceanic currents drive population structure of Heritiera littoralis, a widespread
mangrove in the Indo-West Pacific. Ecol Evol 00:1–15. doi: https://doi.org/10.1002/ece3.6460.
Bayraktarov E, Saunders MI, Abdullah S, Mills M, Beher J, Possingham HP, Mumby PJ, Lovelock
CE. 2016. The cost and feasibility of marine coastal restoration. Ecol Appl 264: 1055–1074.
Begam MM, Sutradhar T, Chowdhury R, Mukherjee C, Basak SK, Ray K (2017) Native salt-
tolerant grass species for habitat restoration, their acclimation and contribution to improving
edaphic conditions: a study from a degraded mangrove in the Indian Sundarbans. Hydrobiologia
803: 373–387.
Bosire JO, Dahdouh-Guebas F, Walton M, Crona BI, Lewis RR, Field C, Koedam N (2008)
Functionality of restored mangroves: a review. Aquat Bot 89(2): 251–259.
Brown B, Fadilla R, Nurdin Y, Soulsby I, Ahmad R (2014b) Community based ecological
mangrove rehabilitation (CBEMR) in Indonesia. SAPIENS 7: 53–64.
Brown B, Yuniati W, Ahmad R, Soulsby I (2014a) Observations of natural recruitment and human
attempts at mangrove rehabilitation after seismic (tsunami and earthquake) events in Simulue
Island and Singkil Lagoon, Acheh, Indonesia. In Santiago-Fandino V, Kontar YA, Kaneda Y
(eds), Post-Tsunami Hazard Reconstruction and Restoration. Springer, New York: 311–327.
Bryan-Brown DN, Connolly RM, Richards DR, Adame F, Friess DA, Brown CJ (2020) Global
trends in mangrove forest fragmentation. Sci Rep 10: 7117.
Bunting P, Rosenqvist A, Lucas RM, Rebelo LM, Hilarides L, Thomas N, Hardy A, Itoh T,
Shimada M, Finlayson CM (2018) The Global Mangrove Watch-A New 2010 Global Baseline
of Mangrove Extent. Remote Sens 10: 1669.
Cerón-Souza I, Gonzalez EG, Schwarzbach AE, Salas-Leiva DE, Rivera-Ocasio E, Toro-Perea N,
Bermingham E, McMillan WO (2015) Contrasting demographic history and gene flow patterns
of two mangrove species on either side of the central American isthmus. Ecol Evol 5:
3486–3499.
Chapman VJ (1975) Mangrove Biogeography. In: Proceedings of the International Symposium on
Biology and Management of Mangroves, eds. Walsh GE, Snedaker SC, Teas HJ, University of
Florida Press, Gainesville. pp. 3–22.
Chapman VJ (1976) Mangrove vegetation. J. Cramer, Lehre, Germany, 447 pp.
Churchill DM (1973) The ecological significance of tropical mangroves in the early Teniary floras
of southern Australia. Geological Society of Australia-Special Publication 4: 79-86.
Cooper WE, Kudo H, Duke NC (2016) Bruguiera hainesii C.G.Rogers (Rhizophoraceae), an
endangered species recently discovered in Australia. Austrobaileya 9:481–488.
Dale PER Knight JM, Dwyer PG (2014) Mangrove rehabilitation: a review focusing on ecological
and institutional issues. Wetl Ecol Manag 22: 587–604.
Duke NC (1984) A mangrove hybrid, Sonneratia  gulngai (Sonneratiaceae) from north-eastern
Australia. Austrobaileya 2: 103–105.
Duke NC (1992) Mangrove floristics and biogeography. In: Robertson AI, Alongi DM (eds)
Tropical mangrove ecosystems. American Geophysical Union, Washington DC, USA,
pp. 63–100
Duke NC (1994) A Mangrove Hybrid, Sonneratia x urama (Sonneratiaceae) from Northern
Australia and Southern New Guinea. Aust Syst Bot 7:521–526.
56 P. Ragavan et al.

Duke NC (1995) Genetic diversity, distributional barriers and rafting continents—more thoughts on
the evolution of mangroves. Hydrobiologia 295: 167-181.
Duke NC (2010) Overlap of eastern and western mangroves in the South-western Pacific:
hybridization of all three Rhizophora (Rhizophoraceae) combinations in New Caledonia.
Blumea 55:171–188.
Duke NC (2017) Mangrove floristics and biogeography revisited: further deductions from biodi-
versity hot spots, ancestral discontinuities, and common evolutionary processes. In: Rivera-
Monroy VH, Lee SY, Kristensen E, Twilley RR (e) (eds) Mangrove ecosystems: a global
biogeographic perspective. Springer, Cham, pp. 17–53.
Duke NC (2020)Asystematic revision of the vulnerable mangrove genus Pelliciera
(Tetrameristaceae) in equatorialAmerica. Blumea 65(2):107–120. https://doi.org/10.3767/
blumea.2020.65.02.04
Duke NC, Ball MC, Ellison JC (1998a) Factors influencing biodiversity and distributional gradients
in mangroves. Glob Ecol Biogeogr 7(1):27–47.
Duke NC, Benzie JAH, Goodall JA, Ballment ER (1998b) Genetic structure and evolution of
species in the mangrove genus Avicennia (Avicenniaceae) in the Indo-West Pacific. Evolution
52(6):1612–1626.
Duke NC, Bunt JS (1979) The genus Rhizophora (Rhizophoraceae) in north-eastern Australia. Aust
J Bot 27: 657–678.
Duke NC, Kovacs JM, Griffith A, Preece L, Hill DJ, van Oosterzee P, Mackenzie J, Morning HS,
Burrows D (2017) Large-scale dieback of mangroves in Australia’s Gulf of Carpentaria: a
severe ecosystem response, coincidental with an unusually extreme weather event. Mar Freshw
Res doi:https://doi.org/10.1071/MF16322.
Duke NC, Kudo H (2018) Bruguiera  dungarra, a new hybrid between mangrove species
B. exaristata and B. gymnorhiza (Rhizophoraceae) recently discovered in north-east
Australia. Blumea 63:279–285
Ellison AM, Farnsworth EJ, Merkt RE (1999) Origins of mangrove ecosystems and the mangrove
biodiversity anomaly. Glob Ecol Biogeogr 8:95–115. https://doi.org/10.1046/j.1466-822X.
1999.00126.x
Ellison AM, Felson AJ, Friess DA (2020) Mangrove Rehabilitation and Restoration as Experimen-
tal Adaptive Management. Front Mar Sci 7: 327. doi: https://doi.org/10.3389/fmars.2020.
00327.
Feller IC, Friess DA, Krauss KW, Lewis III RR (2017) The state of the world’s mangroves in the
21st century under climate change. Hydrobiologia 803: 1–12.
Frankham DA, Ballou R, Briscoe JD (2010) Introduction to conservation genetics, 2nd
ed. Cambridge University Press, Cambridge.
Ge XJ (2001) Genetic diversity and conservation genetics of mangrove species in South China and
Hong Kong. Hong Kong: University of Hong Kong.
Ge XJ, Sun M (2001) Population genetic structure of Ceriops tagal (Rhizophoraceae) in Thailand
and China. Wetl Ecol Manag 9: 213–219.
Goldberg L, Lagomasino D, Thomas N, Fatoyinbo T (2020) Global declines in human-driven
mangrove loss. Global Change Biol 00:1–12. doi: https://doi.org/10.1111/gcb.15275.
Guo Z, Huang YJ, Chen Y, Duke NC, Zhong C, Shi S (2016) Genetic discontinuities in a dominant
mangrove Rhizophora apiculata (Rhizophoraceae) in the Indo-Malesian region. J Biogeogr 43
(9): 1856–1868.
Guo Z, Li X, He Z, Yang Y, Wang W, Zhong C, Greenberg AJ, Wu C-I, Duke NC, Shi S (2018a)
Extremely low genetic diversity across mangrove taxa reflects past sea level changes and hints at
poor future responses. Global Change Biol 24:1741–1748.
Guo W, Ng WL, Wu H, Li W, Zhang L, Qiao S, et al. (2018b). Chloroplast phylogeography of a
widely distributed mangrove species, Excoecaria agallocha, in the Indo-West Pacific region.
Hydrobiologia 807: 333–347. doi: https://doi.org/10.1007/s10750-017-3409-7.
Guo Z, Guo W, Wu H, Fang X, Ng WL, et al. (2018c) Differing phylogeographic patterns within
the Indo-West Pacific mangrove genus Xylocarpus (Meliaceae). J Biogeogr 45: 676–689.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 57

Guo W, Banerjee AK, Ng WL, Yuan Y, Li W and Huang Y (2020) Chloroplast DNA
phylogeography of the Holly mangrove Acanthus ilicifolius in the Indo-West Pacific.
Hydrobiologia 847:3591–3608.
Hamilton SE, Casey D (2016) Creation of a high spatio-temporal resolution global database of
continuous mangrove forest covers for the 21st century (CGMFC-21). Glob Ecol Biogeogr 25
(6): 729–738.
Hamrick JL, Godt MJW, Sherman-Broyles SL (1992) Factors influencing levels of genetic diversity
in woody plant species. New For 6: 95–124. https://doi.org/10.1007/BF00120641.
He Z, Li X, Yang M, et al. (2019) Speciation with gene flow via cycles of isolation and migration:
insights from multiple mangrove taxa. Natl. Sci. Rev. 6: 275–88.
Hoffmann AA, Sgro CM (2011) Climate change and evolutionary adaptation. Nature 470:
479–485.
Huang L, Li X, Huang V, Shi S, Zhou R (2014) Molecular evidence for natural hybridization in the
mangrove genus Avicennia. Pak J Bot 46: 1577–1584.
Huang Y, Tan F, Su G, Deng S, He H, Shi S (2008) Population genetic structure of three tree species
in the mangrove genus Ceriops (Rhizophoraceae) from the Indo West Pacific. Genetica 133:
47–56. doi: https://doi.org/10.1007/s10709-007-9182-1.
Inomata N, Wang X-R, Changtragoon S, Szmidt AE (2009) Levels and patterns of DNA variation
in two sympatric mangrove species, Rhizophora apiculata and R. mucronata from Thailand.
Genes Genet Syst 84: 277–286. doi: https://doi.org/10.1266/ggs.84.277.
IUCN and UNEP-WCMC (2019) www.protectedplanet.net.
Kar RK (1985) The fossil floras of Kachchh-IV. Tertiary Palynostratigraphy. Palaeobotanist 34:
1-280.
Kathiresan K (1995) Rhizophora annamalayana: a new species of mangroves. Environ Ecol 13:
240–241.
Kathiresan K, Bingham BL (2001) Biology of mangroves and mangrove ecosystems. Adv Mar Biol
40:81–251.
Ko WC (1985) Notes on the genus Sonneratia (Sonneratiaceae) in S. E. Asia. Acta Phytotax Sin 23:
311–314.
Krauss KW, McKee KL, Lovelock CE, Cahoon DR, Saintilan N, Reef R, Chen L (2014) How
mangrove forests adjust to rising sea level. New Phytol 202: 19–34.
Lee SY, Hamilton S, Barbier EB, Primavera J, Lewis III RR (2019) Better restoration policies are
needed to conserve mangrove ecosystems. Nat Ecol Evol 3: 870–872.
Lee SY, Primavera JH, Dahdouh-Guebas F, McKee K, Bosire JO, Cannicci S, Diele K, Fromard F,
Koedam N, Marchand C, Mendelssohn I, Mukherjee N, Record S (2014) Ecological role and
services of tropical mangrove ecosystems: a reassessment. Glob Ecol Biogeogr 23: 726–743.
Lewis RR, Brown BM, Flynn LL (2019) “Methods and criteria for successful mangrove forest
rehabilitation,” In Coastal Wetlands: An Integrated Ecosystem Approach, 2nd Edn, eds
G. Perillo ME, Wolanski E, Cahoon DR, Hopkinson CS (Amsterdam: Elsevier), pp. 863–887.
doi: https://doi.org/10.1016/B978-0-444-63893-9.00024-1.
Lewis RR, Hodgson AB, Mauseth GS (2005) Project facilitates the natural reseeding of mangrove
forests (Florida). Ecol Restor 23: 276–277.
Lewis RR, Milbrandt EC, Brown B, Krauss KW, Rovai AS, Beever JW, Flynn LL (2016) Stress in
mangrove forests: early detection and pre-emptive rehabilitation are essential for future success-
ful worldwide mangrove forest management. Mar Pollut Bull 109: 764–771.
Li J, Yang Y, Chen Q, Fang L, He Z, Guo W, Qiao S, Wang Z, Guo M, Zhong C, Zhou R, Shi S
(2016) Pronounced genetic differentiation and recent secondary contact in the mangrove tree
Lumnitzera racemosa revealed by population genomic analyses. Sci Rep 6: 29486. doi: https://
doi.org/10.1038/srep29486.
Liao PC, Havanond S, Huang S (2007) Phylogeography of Ceriops tagal (Rhizophoraceae) in
Southeast Asia: The land barrier of the Malay Peninsula has caused population differentiation
between the Indian Ocean and South China Sea. Conserv Genet 8: 89–98. doi: https://doi.org/
10.1007/s10592-006-9151-8.
58 P. Ragavan et al.

Lira-Medeiros CF, Parisod C, Fernandes RA, Mata CS, Cardoso MA, Ferreira PCG (2010)
Epigenetic variation in mangrove plants occurring in contrasting natural environment. PLoS
ONE 5: e10326.
Lo EY, Duke NC, Sun M (2014) Phylogeographic pattern of Rhizophora (Rhizophoraceae) reveals
the importance of both vicariance and long-distance oceanic dispersal to modern mangrove
distribution. BMC Evol Biol 14: 83. doi:https://doi.org/10.1186/1471-2148-14-83.
Lomolino MV, Riddle BR, Brown JH (2006) Biogeography, 3rd edn. Sinauer, Sunderland, MA.
López-Portillo J, Lewis RR, Saenger P, Rovai A, Koedam N, Dahdouh-Guebas F, Agraz-
Hernández C, Rivera-Monroy VH (2017) Mangrove forest restoration and rehabilitation. In:
Rivera-Monroy VH, Lee SY, Kristensen E, Twilley RR (Eds.), Mangrove Ecosystems: A Global
Biogeographic Perspective. Springer, New York, New York, USA, pp. 301–345
Lovelock CE, Cahoon DR, Friess DA, Guntenspergen GR, Krauss KW, Reef R, Rogers K,
Saunders ML, Sidik F, Swales A, Saintilan N, Thuyen AX, Triet T (2015) The vulnerability
of Indo-Pacific mangrove forests to sea-level rise. Nature 526: 559–563.
Lovelock CE, Feller IC, Reef R, Hickey S, Ball MC (2017) Mangrove dieback during fluctuating
sea levels. Sci Rep 7: 1680.
Ludt WB, Rocha LA (2014) Shifting seas: the impacts of Pleistocene sea-level fluctuations on the
evolution of tropical marine taxa. J Biogeogr 42(1): 25–38.
Lyu H, He Z, Wu C-I, Shi S (2018) Convergent adaptive evolution in marginal environments:
unloading transposable elements as a common strategy among mangrove genomes. New Phytol
217: 428–438.
Maguire TL, Saenger P, Baverstock P, Henry R (2000) Microsatellite analysis of genetic structure
in the mangrove species Avicennia marina (Forsk.) Vierh. (Avicenniaceae). Mol Ecol
9:1853–1862.
Mathur YK (1984) Cenozoic palynofossils, vegetation, ecology and climate of the North and North-
western Sub-Himalayan region. In R. O. Whyte (Ed.), The Evolution of the East Asian
Environment, 2 (pp. 504–551). Hong Kong, China: Centre of Asian Studies, Univ. Hong Kong.
Mathur YK, Chowdhary LR (1977) Palaeoecology of the Kalol Formation, Cambay Basin, India. In
B. S. Venkatachala, & V. V. Sastri (Eds.), (pp. 164–178). Proceedings of the 4th Colloquium on
Indian Micropalaeontology and Stratigraphy, Dehradun, 1974–75. Dehradun, Uttarakhand:
Institute of Petroleum Exploration, Oil and Natural Gas Commission.
McCoy ED, Heck KL Jr (1976). Biogeography of corals, seagrasses, and mangroves: An alternative
to the center of origin concept. Syst Zool 25:201–210. doi: https://doi.org/10.2307/2412488.
Mepham RH (1983) Mangrove floras of the southern continents. Part 1. The geographical origin of
Indo-Pacific mangrove genera and the development and present status of the Australian
mangroves. South Afr J Bot 2: 1–8.
Minobe S, Fukui S, Saiki R, Kajita T, Changtragoon S, Shukor NAA, Latiff A, Ramesh BR,
Koizumi O, Yamazaki T (2010) Highly differentiated population structure of a Mangrove
species, Bruguiera gymnorhiza (Rhizophoraceae) revealed by one nuclear Gap Cp and one
chloroplast intergenic spacer trnF–trnL. Conserv Genet 11: 301–310.
Mori GM, Zucchi MI, Sampaio I, Souza AP (2015) Species distribution and introgressive
hybridization of two Avicennia species from the Western Hemisphere unveiled by
phylogeographic patterns. BMC Evol Biol 15: 1–15.
Ono J (2016) Conservation genetic study of mangrove plant genus Bruguiera. Ph D thesis, Chiba
University.
Ono J, Yong JWH, Takayama K, Saleh MNB, Wee AKS, Asakawa T, Yllano OB, Salmo SG,
Meenakshisundaram SH, Watano Y, Webb EL, Kajita T (2016) Bruguiera hainesii, a critically
endangered mangrove species, is a hybrid between B. cylindrical and B. gymnorhiza
(Rhizophoraceae). Conserv Genet 17:1137–1144. doi: https://doi.org/10.1007/s10592-016-
0849-y.
Qiu S, Zhou RC, Li YQ, Havanond S, Jaengjai C, Shi SH (2008) Molecular evidence for natural
hybridization between Sonneratia alba and S. griffithii. J Syst Evol 46: 391–395.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 59

Ragavan P, Saxena A, Jayaraj RSC, Ravichandran K, Saravanan S (2015). Rhizophora  mohanii:


a putative hybrid between Rhizophora mucronata and Rhizophora stylosa from mangroves of
the Andaman and Nicobar Islands, India. ISME/GLOMIS Electronic Journal 13: 3–7.
Ragavan R, Saxena A, Jayaraj RSC, Mohan PM, Pavichandran K, Saravanan S, Vijayaraghavan A
(2016) A review of the mangrove floristics of India. Taiwania 61:224–242.
Renema W, Bellwood DR, Braga JC, Bromfield K, Hall R et al (2008) Hopping hotspots: global
shifts in marine biodiversity. Science 321:654–657.
Rey JR, Carlson DB, Brockmeyer RE (2012) Coastal wetland management in Florida: environ-
mental concerns and human health. Wetl Ecol Manag 20: 197–211.
Riddle BR, Dawson MN, Hadly EA, Hafner DJ, Hickerson MJ, Mantooth SJ, Yoder AD (2008)
The role of molecular genetics in sculpting the future of integrative biogeography. Prog Phys
Geogr 32: 173–202.
Saenger P, Hegerl EJ, Davie JDS (1983) Global Status of Mangrove Ecosystems. The Environ-
mentalist 3 (Supplement): 1–88.
Saenger P, Ragavan P, Sheue CR, López-Portillo J, Yong JWH, Mageswaran T (2019) Mangrove
Biogeography of the Indo-Pacific. In Sabkha Ecosystems Volume VI: Asia/Pacific, Gul B,
Böer B, Khan, MA, Cluesener-Godt M, Hameed A (Eds.) Chapter 23, pp. 379–400.
Saintilan N, Wilson NC, Rogers K, Rajkaran A, Krauss KW (2014). Mangrove expansion and salt
marsh decline at mangrove poleward limits. Glob Change Biol 20: 147–157.
Sheue CR, Liu HY, Tsai CC, Rashid SMA, Yong JWH, Yang YP (2009b) On the morphology and
molecular basis of segregation of two species Ceriops zippeliana Blume and C. decandra
(Griff.) Ding Hou (Rhizophoraceae) from southeastern Asia. Blumea 54:220–227
Sheue CR, Liu HY, Tsai CC, Yang YP (2010) Comparison of Ceriops pseudodecandra sp. nov.
(Rhizophoraceae), a new mangrove species in Australasia, with related species. Bot Stud
51:237–248.
Sheue CR, Liu HY, Yong JWH (2003) Kandelia obovata (Rhizophoraceae), a new mangrove
species from Eastern Asia. Taxon 52:287–294
Sheue CR, Yang YP, Liu HY, Chou FS, Chang SC, Saenger P, Mangion CP, Wightman G, Yong
JWH, Tsai CC (2009a) Reevaluating the taxonomic status of Ceriops australis
(Rhizophoraceae) based on morphological and molecular evidence. Bot Stud 50:89–100
Sheue CR, Yong JWH, Yang YP (2005) The Bruguiera (Rhizophoraceae) species in the mangroves
of Singapore, especially on the new record and the rediscovery. Taiwania 50:251–260
Specht RL (1981) Biogeography of halophytic angiosperms (saltmarsh, mangrove and sea-grass).
In A. Keast (ed.), Ecological Biogeography of Australia. W. Junk, The Hague.
Srivastava J, Prasad V (2019) Evolution and paleobiogeography of mangroves. Mar. Ecol. 00:
e12571. doi:https://doi.org/10.1111/maec.12571.
Srivastava SK, Binda PL (1991) Depositional history of the early eocene shumaysi formation, Saudi
Arabia. Palynology 15: 47–61.
Stubbs BJ, Saenger P (2002). The application of forestry principles to the design, execution and
evaluation of mangrove restoration projects. Bois et Foréts des Tropiques 56: 5–21.
Su G, Huang Y, Tan F, Ni X, Tang T, Shi S (2007) Conservation genetics of Lumnitzera littorea
(Combretaceae), an endangered mangrove, from the Indo-West Pacific. Mar Biol 150: 321–328.
doi: https://doi.org/10.1007/s00227-006-0357-6.
Su GH, Huang YL, Tan FX, Ni XW, Tang T, Shi SH (2006) Genetic variation in Lumnitzera
racemosa, a mangrove species from the Indo-West Pacific. Aquat Bot 84: 341–346. https://doi.
org/10.1016/j.aquabot.2006.01.001.
Sun M, Lo EY (2011) Genomic markers reveal introgressive hybridization in the Indo-West Pacific
mangroves: a case study. PLoS One 6: e19671.
Takayama K, Tamura M, Tateishi Y, Webb EL, Kajita T (2013) Strong genetic structure over the
American continents and transoceanic dispersal in the mangrove genus Rhizophora
(Rhizophoraceae) revealed by broad-scale nuclear and chloroplast DNA analysis. Am J Bot
100: 1191–1201.
60 P. Ragavan et al.

Tan F, Huang Y, Ge X, Su G, Ni X, Shi S (2005) Population genetic structure and conservation


implications of Ceriops decandra in Malay Peninsula and North Australia. Aquat Bot
81:175–188. https://doi.org/10.1016/j.aquabot.2004.11.004.
Thanikaimoni G (1987) Mangrove palynology. Tr Sec. Sci Tech. Institute of French Pondichery 24:
100.
Thomas N, Lucas R, Bunting P, Hardy A, Rosenqvist A, Simard M (2017) Distribution and drivers
of global mangrove forest change, 1996–2010. PLoS ONE 12(6): e0179302. doi: https://doi.
org/10.1371/journal.pone.0179302.
Tomizawa Y, Tsuda Y, Saleh M, Wee A, Takayama K, Yamamoto T, et al. (2017) Genetic structure
and population demographic history of a widespread mangrove plant Xylocarpus granatum
J. Koenig across the Indo-West Pacific Region. Forests 8: 480. doi: https://doi.org/10.3390/
f8120480.
Tomlinson BP, Womersley JG (1976) A Rhizophora new to Queensland and New Guinea, with
notes relevant to the genus. Contrib Herb Aust 19: 1–10.
Tomlinson PB (1978) Rhizophora in Australasia—some clarification of taxonomy and distribution.
J Arnold Arbor 59: 156–169.
Tomlinson PB (1986) The Botany of Mangroves. Cambridge University Press, Cambridge, 413 pp.
Tomlinson PB (2016) The botany of mangroves, 2nd edn. Cambridge University Press, Cambridge.
Tomlinson PB, Bunt JS, Primack RB, Duke NC (1978) Lumnitzera  rosea (Combretaceae)—its
status and floral morphology. J Arnold Arbor 59: 342–351.
Tsai CC, Li SJ, Su YY, Yong JWH, Saenger P, Chesson P, Das S, Wightman G, Yang YP, Liu HY,
Sheue CR (2012) Molecular phylogeny and evidence for natural hybridization and historical
introgression between Ceriops species (Rhizophoraceae). Biochem Syst Ecol 43: 178–191.
Urashi C, Teshima KM, Minobe S, Koizumi O, Inomata N (2013) Inference of evolutionary history
of a widely distributed mangrove species, Bruguiera gymnorrhiza, in the Indo-West Pacific
region. Ecol Evol 3: 2251–2261.
Van Steenis CGGJ (1962) The distribution of mangrove plant genera and its significance for
Palaeogeography. Nederl. Akad. Von Wetenschappen 65:164–169.
Venkatachala BS, Rawat MS (1972) Palynology of the Tertiary sediments in the Cauvery basin
1. Palaeocene-Eocene palynoflora from the subsurface. In Proceedings of the Seminar on
Palaeopalynology and Indian Stratigraphy, Calcutta (pp. 292–335).
Wang PX, Wang LJ, Bian YH, Jian ZM (1995) Late Quaternary paleoceanography of the South
China Sea: surface circulation and carbonate cycles. Mar Geol 127:145–165.
Wang RJ, Chen ZY, Chen EY, Zheng XR (1999) Two hybrids of the genus Sonneratia
(Sonneratiaceae) from China. Guihaia 19: 199–204.
Wee A, Teo J, Chua J, Takayama K, Asakawa T, Meenakshisundaram S, et al. (2017) Vicariance
and oceanic barriers drive contemporary genetic structure of widespread mangrove species
Sonneratia alba J. Sm in the Indo-West Pacific. Forests 8: 483. doi: https://doi.org/10.3390/
f8120483.
Wee AKS, Mori GM, Lira CF, Núñez-Farfán J, Takayama K, Faulks L, Shi S, Tsuda Y, Suyama Y,
Yamamoto T, Iwasaki T, Nagano Y, Wang Z, Watanabe S, Kajita T (2018) The integration and
application of genomic information in mangrove conservation. Conserv Biol 33: 206–209.
Wee AKS, Noreen AME, Ono J, et al. (2020) Genetic structures across a biogeographical barrier
reflect dispersal potential of four Southeast Asian mangrove plant species. J Biogeogr 00: 1–14.
doi: https://doi.org/10.1111/jbi.13813.
Wee AKS, Takayama K, Asakawa T, Thompson B, Sungkaew S, Tung NX et al. (2014) Oceanic
currents, not land masses, maintain the genetic structure of the mangrove Rhizophora
mucronata Lam. (Rhizophoraceae) in Southeast Asia J Biogeogr 41: 954–964.
Wee AKS, Takayama K, Chua JL, Asakawa T et al. (2015) Genetic differentiation and
phylogeography of partially sympatric species complex Rhizophora mucronata Lam. and
R. stylosa Griff. using SSR markers. BMC Evol Biol 15: 57. https://doi.org/10.1186/s12862-
015-0331-3.
2 Biogeography of the Mangrove Ecosystem: Floristics, Population Structure,. . . 61

Woodroffe CD, Rogers K, McKee KL, Lovelock CE, Mendelssohn IA, Saintilan N (2016)
Mangrove sedimentation and response to relative sea-level rise. Ann Rev Mar Sci 8: 243–266.
Worthington T, Spalding M (2019) Mangrove restoration potential: A global map highlighting a
critical opportunity. 34pp.
Wyrtki K (1961) Scientific results of marine investigations of the South China Sea and the Gulf of
Thailand 1959–1961. NAGA report 2. pp. 164–169.
Xie W, Zhong G, Li X, Guo Z, Shi S (2020) Hybridization with natives augments the threats of
introduced species in Sonneratia mangroves. Aquat Bot 160: 103166.
Yan Y-B, Duke NC, Sun M (2016) Comparative Analysis of the Pattern of Population Genetic
Diversity in Three Indo-West Pacific Rhizophora Mangrove Species. Front Plant Sci 7: 1434.
doi: https://doi.org/10.3389/fpls.2016.01434.
Yang Y, Duke NC, Peng F, Li J, Yang S, Zhong C, Zhou R, Shi S (2016) Ancient Geographical
Barriers Drive Differentiation among Sonneratia caseolaris Populations and Recent Divergence
from S. lanceolata. Front Plant Sci 7:1618. doi: https://doi.org/10.3389/fpls.2016.01618.
Yang Y, Li J, Yang S, Li X, Fang LU, Zhong C, et al. (2017) Effects of Pleistocene sea-level
fluctuations on mangrove population dynamics: A lesson from Sonneratia alba. BMC Evol Biol
17: 22. doi: https://doi.org/10.1186/s12862-016-0849-z.
Zhang R, Liu T, Wu W, Li Y, Chao L, Huang L, Huang Y, Shi S, Zhou R (2013) Molecular
evidence for natural hybridization in the mangrove fern genus Acrostichum. BMC Plant Biol 13:
74.
Zhang ZH, Zhou RC, Tang T, Huang YL, Zhong Y, Shi SH (2008) Genetic variation in central and
peripheral populations of Excoecaria agallocha from Indo-West Pacific. Aquat Bot 89: 57–62.
Zhong C, Li D, Zhang Y (2020) Description of a new natural Sonneratia hybrid from Hainan Island,
China. PhytoKeys 154: 1–9.
Mangroves as Feeding and Breeding
Grounds 3
D. Arceo-Carranza, X. Chiappa-Carrara, R. Chávez López, and
C. Yáñez Arenas

Abstract

Mangroves are considered as ecosystems that provide shelter, food and breeding
grounds for many groups of inhabiting fauna. Much of the fauna present are
organisms in different stages of their life cycle, mostly juveniles. The three-
dimensional structure of the mangrove roots and the combination of the aquatic
and terrestrial environments are factors that bring together a great diversity. Such
diversity within mangrove sites includes aquatic and terrestrial vertebrates and
invertebrates such as fish, amphibians, reptiles, mammals and birds. Present
within the fauna are representatives of different trophic guilds that perform key
functions in the ecosystem, such as pollination, seed dispersal and nutrient
recirculation. The food produced by the ecosystem is based on the production
of detritus caused by leaf litter and its decomposition, where transformation of
energy and accumulation of biomass for higher trophic levels begins with the

D. Arceo-Carranza (*)
Facultad de Ciencias, Unidad Multidisciplinaria de Docencia e Investigación Sisal, Universidad
Nacional Autónoma de México, Sisal, Yucatán, Mexico
e-mail: [email protected]
X. Chiappa-Carrara
Facultad de Ciencias, Unidad Multidisciplinaria de Docencia e Investigación Sisal, Universidad
Nacional Autónoma de México, Sisal, Yucatán, Mexico
Escuela Nacional de Estudios Superiores Mérida, Universidad Nacional Autónoma de México,
Mérida, Yucatán, Mexico
R. Chávez López
Facultad de Estudios Superiores Iztacala, Universidad Nacional Autónoma de México, Mexico
City, Mexico
C. Yáñez Arenas
Facultad de Ciencias, Unidad Multidisciplinaria de Docencia e Investigación Sisal, Sede Parque
Científico, Universidad Nacional Autónoma de México, Mérida, Mexico

# The Author(s), under exclusive license to Springer Nature Singapore Pte 63


Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_3
64 D. Arceo-Carranza et al.

invertebrates. With regard to breeding activity, many fish families spawn in the
mangrove roots (e.g. Fundulidae), or nest in the canopy (herons and cormorants)
that provides protection against predators and food for juvenile organisms.
Undoubtedly, mangroves function as feeding and breeding grounds that are
essential in maintaining populations of marine organisms, especially fish. Many
fish species that grow in the mangroves are important for fisheries. Unequivo-
cally, maintaining these breeding sites for marine and terrestrial fauna is crucial to
the general functioning of adjacent ecosystems.

Keywords
Guilds · Nursery · Life stage · Juvenile · Refuge · Trophic ecology

3.1 Introduction

Coasts throughout the world are comprised of ecosystems formed by vegetation


types where environmental adaptations are necessary for life in an environment with
constant changes in hydrological characteristics, prominently, salinity and flooding.
These halophytes from the mangrove ecosystems nurture a great diversity world-
wide. The niche diversity of this ecosystem occurs at the aquatic, sediment, roots,
trunk, and foliage level, serving as a habitat and food source to many animal species.
Therefore, mangrove fauna is diverse due to the geographic, hydrological and
climatic characteristics as well as ecological processes that occur in these
ecosystems. This chapter will discuss important mangrove characteristics that
make this ecosystem an important habitat for fauna, in relation to food availability,
and as a nursery for several species during different stages of their life cycles,
making mangroves an essential habitat.

3.2 Mangroves, a Habitat for Fauna

Mangrove ecosystems are considered as a type of coastal wetland dominated by


 
woody plants, that are distributed on particular latitudinal gradients (30 N to 37 S),
tidal height (<1 m to >4 m), with particular geomorphology (from continental river
systems to oceanic islands), various sediment composition (from organic peat to
alluvial zones), to types of climate (from temperate warm, arid zones to humid
tropics), various nutrient availability (from oligotrophic to eutrophic levels), and
with soils that are principally anaerobic. Their success is the result of highly
developed morphological and physiological adaptations.
All these characteristics categorize mangroves as biocomplex systems (Feller
et al. 2010) due to the processes that occur at the individual, population and
ecosystem level, as well as their natural biotic and abiotic interactions. Mangroves
have been characterized over time as being very productive sites that hold great
diversity, this is because they are located on the borders between marine and
3 Mangroves as Feeding and Breeding Grounds 65

terrestrial ecosystems. In this coastal ecotone, we can find fauna with distinctive
features, those of saltwater and freshwater origins, or fauna that adapt and tolerate
fluctuating environmental conditions (e.g. salinity, flooding, pH and dissolved
oxygen concentration in sediments). Inhabiting the mangrove is local fauna known
as permanent residents, found at birth, growing, reproducing and dying within this
ecosystem. Also, there is fauna that can be considered as temporary visitors,
inhabiting mangroves for short periods, such as migratory bird species (e.g. herons
and pelicans), fish (e.g. grunts and snappers), marine invertebrates (e.g. shrimp) and
amphibians, to mention a few. These species use mangroves to reproduce, spawn or
as resting sites (Kathiresan and Bingham 2001; Nagelkerken et al. 2008; De Dios
et al. 2019). Occasional or facultative visitors can be found in mangroves due to its
high food source, but these species do not require mangroves to fulfill their life cycle.
These systems are considered relicts for endemic species or for species with a need
for protection (Luther and Greenberg 2009). Ecologically, mangroves are crucial for
estuarine ecosystems, and are one of the most utilized and threatened habitats due to
anthropogenic activities, as increased degradation in the last 30 years has caused a
35% loss of mangrove surface (Barbier et al. 2011). The most deteriorated ecosystem
functions are the nursery areas (consequently, causing the collapse of some coastal
fisheries), filtration and purification (services provided by detritus feeders,
e.g. bivalves, barnacles and sponges), decreased water quality, loss of biodiversity
and their role in protecting the coastline (Worm et al. 2006).
The connectivity of this ecosystem is a feature resulting from the interactions
between the mainland and geomorphology, hydrology, weather and tidal systems,
structural characteristics, accessibility to fauna, and the borders of the mangrove
forests; altogether this stimulates the exchange between the ecosystem limits
(Nagelkerken et al. 2008). This connectivity not only contributes to the economic
value and ecosystem services, but also increases its vulnerability to natural and
human disturbances (Alongi 2008).

3.2.1 Canopy

Found in the aerial section of mangroves, comprising a considerable portion of the


fauna, are the insects, either as permanent residents or temporary visitors, their
presence is important given that they carry out different functions in this dynamic
ecosystem. Due to their mobility, many insect species are temporary visitors of
mangroves, and thus create links to other habitats they inhabit (Balakrishnan et al.
2014). The mangrove canopy sustains terrestrial fauna, which as in other ecosystems
is dominated by insects, but also includes spiders, crabs and members of all
terrestrial invertebrates. However, only a few endemic species are reported (Luther
and Greenberg 2009). Brooks and Bell (2001) indicate that many species are
subgroups of terrestrial fauna that scatter in the mangrove by flying, swimming or
even floating on debris. Accordingly, species that feed on flowers and fruits such as
frugivores, folivores and palynivores can be found (e.g. butterflies, bees, crickets,
bats, rodents, birds, among others). Although the presence of dragonflies, saddles
66 D. Arceo-Carranza et al.

and termites is common (Balakrishnan et al. 2018), a considerable number of studies


have focused on bees, ants and mosquitos, and as per the former, honey produced in
mangroves is valued in places such as India, the Caribbean, Florida and Brazil
(Fernandes et al. 2011).
In Panama, Adams (1994) demonstrated that four species of ants divided the
resources of the mangrove canopy by marking territories with a combination of
tactile and chemical pheromone signals, whereas other animals (e.g. birds, reptiles)
use the upper part of mangroves for resting or as a refuge site (Gopal and Chauhan
2006). In Mexico, De Dios et al. (2019) reported five species of marine migratory
birds that nest in the mangrove canopy of the Yucatan Peninsula.

3.2.2 Trunks and Branches

Terrestrial insects that inhabit mangroves face severe environmental conditions such
as insolation, high temperatures and high probabilities of desiccation. Nonetheless,
insects as well as other arthropods deal with these adversities through nocturnal
behaviour or by living inside plants. Moths and beetles in Belize have been found to
dig tunnels through mangroves, a habitat modification used by more than 70 species
of ants, spiders, mites, moths, cockroaches, termites and scorpions (Feller 2002).
Insects and spiders also temporarily use trunks and branches in flooding events,
while other species (e.g. isopods, amphipods and myriapods) show affinity for trees
in the intertidal zone to avoid desiccation (Feller and Mathis 1997). The feeding
habits of some organisms such as herbivorous insects can harm the mangrove
vegetation through drilling of the bark and wood, they also cause defoliation, and
seedlings can be vulnerable when growing next to adult trees (Jenoh et al. 2016).
One of the reasons mangroves are disregarded and unpopular is due to them
harbouring large populations of mosquitos; holes inside tree barks made by insects
and crab burrowers are ideal sites, particularly in species of Avicennia spp. (Ismail
et al. 2018). The insects that inhabit mangroves can be grouped as cryptic and
endophytic species, e.g. miners, borers and gall creators. In the intertidal zone
substrates such as trunks, aerial roots and muddy saline plains are used by some
benthonic organisms.

3.2.3 Roots

Considering that mangroves are surrounded by sand or lime sediments, with roots,
branches and trunks, converts them into ecological islands that are an attractive
habitat for a variety of epifaunal communities (Guerra-Castro et al. 2011). This type
of biota can include an assemblage of different invertebrates, such as sponges,
hydroids, anemones, polychaetes, bivalves, barnacles, bryozoans and ascidians;
some of these organisms show morphological adaptations to mangrove life, includ-
ing a display of different morphology when inhabiting roots (Díaz et al. 2004).
3 Mangroves as Feeding and Breeding Grounds 67

A basic function of the mangrove roots at the water–sediment interface is to


recycle nutrients, these chemical compounds are carried by freshwater currents and
tidal circulation patterns, also included, are inorganic particles, organic matter,
sediments and other contaminants.
Found associated with the roots is a community of various bacteria, that are
involved in the decomposition of organic matter. These bacterial communities
occupy different niches and are important elements for the mangrove to function,
since they control all the chemical aspects. For instance, sulphate reducing bacteria
Desulfovibrio, Desulfotomaculum, Desulfosarcina and Desulfococcus are the first
participants in the decomposition processes. These bacteria control the dynamics of
the ions, iron, phosphorous and sulphur, which contribute to soil formation and
vegetation patterns (Varon-Lopez et al. 2014). Methanogenic cyanobacteria are
temporally abundant in Avicennia species and are critical for recycling nitrogen,
and as components of the mangrove microbiota, as they provide an important
nitrogen source (Arai et al. 2016). The function of aerial roots is to influence the
tidal flow rate, which determines the sedimentation and particle retention rates. The
variation in water flow and height of the water column affects the availability of food
particles for filtering organisms such as zooplankton (Srikanth et al. 2016).
In the entangled roots below water are spaces or tunnels shaped by the tidal level
and root system, where a zooplankton community can be found with abundances
that can be very high in densities of up to 10,000 organisms/m3 and 623 mg/m3 of
biomass, surpassing records obtained from marine waters (Saravanakumar et al.
2007). Another important group that finds refuge and food between the roots is the
fish community, mostly juvenile marine fish that use this habitat for a short period of
their life cycle.

3.2.4 Sediments

The basic structure and functional attributes of the mangrove ecosystem are usually
represented by models that show the main external energy sources and the pressures
that affect them (Lugo and Snedaker 1974; Mukherjee et al. 2014). In these
representations, the ecosystem is divided into two compartments. The first, groups
the structures found above the ground, and the second includes, what remains
flooded for most of the time (i.e. roots and sediments) with aerobic and anaerobic
processes occurring as well. In these models a high primary productivity is
recognized for mangrove ecosystems, with a continuous nutrient supply essential
for plant growth. Besides availability, nutrient requirements are possible due to
efficient capture, absorption and nutrient recycling systems (Woodroffe 1992;
Alongi et al. 1993; Alongi 1994; Kristensen et al. 1994).
Mangrove systems offer a wide range of benthonic substrates for the infauna
(Hsieh 1995), from poorly consolidated sludge, sand mixed with detritus from the
mangrove itself or neighbouring ecosystems, to hard substrates such as roots and
other supporting structures (Alongi 1989; Day et al. 2013). The complexity of this
68 D. Arceo-Carranza et al.

habitat enables the assumption that mangroves host a great diversity of communities
(Eller and Grassle 1992).
Three main carbon sources available for invertebrates are recognized in the
surface layers of the sediment: mangrove leaf litter, plant detritus from neighbouring
systems and microphytobenthos. However, reports by some authors (Bouillon et al.
2002; Oakes et al. 2010) show that mangroves provide for carbon assimilation which
is important to a limited number of species. The production of local and imported
microphytobenthos is the most important carbon source of the invertebrate ben-
thonic communities within the intertidal mangroves. Nonetheless, most studies
conclude that mangroves are the carbon source that sustains the dominant
communities, both in numbers and biomass (e.g. crabs and gastropods) (Wells
1984; Camilleri 1992).
These species play an important role in leaf litter degradation, unlike other
components of the meiofauna which are normally found low in abundance in the
mangrove sediments (Alongi 1987). Small-sized infauna from mangrove sediments
is not well studied, even though these organisms might play an important role in the
food web, being an important food source for fish and crustaceans that use
mangroves as a nursery area (Daniel and Robertson 1990; Dittmann 2000).
Several factors associated with hydrodynamics are responsible for the presence
and abundance of the infauna in mangrove sediments. Other biological factors such
as competition with epifauna, predation, food quality and the chemical defence from
mangroves (Alongi 1989) should be studied, in order to explain spatial distribution
patterns and species composition. Recently, taxonomic diversity in water and man-
grove sediment has been explored using next generation sequencing technology, as
well as, isotope techniques to demonstrate the rich diversity of bacteria consortia,
diverse in metabolic functions (Pascal et al. 2016).

3.2.5 The Aquatic Environment

The hydrological characteristics in mangroves vary according to season; these


changes can follow a daily cycle relative to the tides, or stationary, according to
precipitation input during the rainy season. The main change is observed in the
salinity gradient, and as a result, fauna distributes itself according to osmoregulatory
capacity.
A great variety of marine invertebrates in juvenile phases (zooplankton) can be
found in the water that floods the mangrove, some of these species return to the sea
as adults to continue forming part of zooplankton (e.g. Mysidacea) or turn towards a
marine benthic life (e.g. shrimps, gastropods), whereas others such as encrusting
species (e.g. oysters and barnacles) settle in estuarine waters to grow and develop
(Guerra-Castro et al. 2011). The zones permanently flooded are used by some fish
species as a refuge during daylight, and move towards other sites during the night to
feed (Kimirei et al. 2011; Ramírez-Martínez et al. 2016).
Turbid waters high in quantities of suspended solids are another feature that offers
protection and food for juvenile species, with notable migrations to these waters
3 Mangroves as Feeding and Breeding Grounds 69

observed in crustaceans and distinct fish species, according to ecological hypotheses


mentioned by various authors (Dorenbosch et al. 2004; Wang et al. 2009; Arceo-
Carranza et al. 2016). This migratory behaviour by some species of fish, and use of
the mangrove during early, or juvenile life stages is known as estuarine dependence
(Nagelkerken et al. 2001; Dorenbosch et al. 2004; Able 2005; Kimirei et al. 2011);
some examples are Ocyurus chrysurus, Haemulon flavolineatum, Gerres cinereus,
Lagodon rhomboides and Elops saurus (Nagelkerken et al. 2001; Able 2005). This
behaviour can be facultative or obligatory, depending on the biology of the species,
or due to temporality and hydrological characteristics of the system.

3.3 Mangrove as Food Resources

Mangroves are one of the most productive ecosystems worldwide (Robertson and
Alongi 1992). Their high productivity is reflected on its diverse food resources, such
as leaves, flowers, seeds and propagules, and a high detritus production generated by
leaf litter decomposition. Mangrove ecosystems are constantly defoliating their
trees, and thus, detritus production occurs throughout the year. Due to this amount
of resources, it is possible to find diverse trophic guilds within mangroves. Among
the primary consumers, we can find palynivores, herbivores, frugivores and
folivores, as well as detritivores and omnivores that obtain their energy from the
system, which is transferred to carnivores and ichthyophages at higher levels.
Overall, these food resources form the base of the food chain in mangrove
ecosystems. According to Azam et al. (1983), an essential element in mangrove
productivity is the microbial loop defined as a group of microorganisms that
transform particle and dissolved organic matter into food for higher trophic levels.
Therefore, bacteria, viruses, fungi and protozoans play an important role in nutrient
assimilation and detritus production, which is the base of the food web responsible in
supporting secondary productivity for mangroves (Pascal et al. 2016).

3.3.1 Primary Production

The organic matter produced by mangroves has been used to estimate its primary
production, which is directly correlated to biomass, i.e. density of trees and canopy
regulated by environmental variables such as precipitation, temperature and nutrient
availability (Vitousek and Sanford 1986).
Primary production in mangroves comprised of biomass (organic matter)
generated by mangrove trees (including roots and pneumatophores), seedlings and
periphyton (a biofilm of organisms contained in a polysaccharide matrix). Mangrove
productivity is magnified by the microbial associations occurring in the sediments,
resulting in recycling and biogeochemical processes for better nutrient use within the
ecosystem. Bacterial communities help to sustain mangroves, as well as in produc-
tivity, based on three mechanisms: (1) mineralization of organic matter under
anaerobic and microaerophilic conditions mostly by sulphur-reducing bacteria;
70 D. Arceo-Carranza et al.

(2) increase in the rate of nitrogen fixation and (3) bacterial symbiosis at the
rhizosphere that provide nutrients and regulate growth substances (Holguin and
Bashan 2007).
Benthonic macrofauna such as crabs (e.g. Grapsid and Sesarma) also have an
important role in the leaf litter dynamics, in the cycling of nutrients, energy flow, and
consequently, in the productivity of the ecosystem (Smith III et al. 1991). Predatory
interactions between herbivores and non-herbivores (e.g. interaction between
isopods and barnacles) have a positive influence on the ecosystem’s productivity,
since predation between species controls populations of organisms that can feed on
propagules, causing a decrease in production and growth of mangrove roots (Perry
1988).

3.3.2 Detritus

Detritus production in mangrove ecosystems is the main energy source and most
detritus in mangroves come from leaf litter. Detritus is defined as all types of
biogenic material at various decomposition stages by microbes (McLusky and Elliott
2004). Some authors have demonstrated transfer of carbon to the meiofauna and
macroinvertebrates (Oakes et al. 2010); both fauna are important links between
producer and consumer levels in this system, as they contribute by adding nitrogen
waste to the sediments, intervene with organic matter decomposition, and serve as
prey for predatory invertebrates, fish and birds. Different authors agree that detritus
is a key element in the diet of many mangrove species (Nagelkerken et al. 2008;
Oakes et al. 2010; Pascal et al. 2016). Holguin and Bashan (2007) mentioned that
from 120 mangrove species, one-third feeds on detritus, serving as a link to energetic
interactions in the ecosystem; some species are penaeid shrimp, insect larvae,
nematodes, oysters, polychaetes, isopods, bivalves and some fish families
(e.g. Mugilidae). Additionally, some authors (Odum and Heald 1972) have noted
that detritus is a food source that is exported to other neighbouring ecosystems such
as sea grasses and coral reefs through suspended particles.

3.3.3 Secondary Production

Mangroves are sites with high secondary productivity due to the growth of
inhabiting fauna. Although mangrove productivity is higher on coastal and riverine
forests, many species are associated to subtidal habitats. Island mangroves with
distinguished clear water and fewer salinity shifts have groups of species rich in
sessile epibionts (e.g. algae, sponges, tunicates and anemones), where numerous
parasitic and mutualistic associations have been detected (Guerra-Castro et al. 2011).
Species diversity in mangroves can also be explained through habitat varieties
created by the heterogenous structures of these plants. Even though these ecosystems
are recognized with low species richness, studies have shown the importance of
3 Mangroves as Feeding and Breeding Grounds 71

interactions between fauna and vegetation for forest development, productivity and
structural complexity (Nagelkerken et al. 2008).
Increase in growth and biomass of aquatic organisms is one of the hypotheses
why mangroves are considered nursery habitats (Beck et al. 2001). This has been
recorded by various authors in marine species; for instance, Lutjanidae and
Haemulidae fish families feed, increase size and weight in mangroves during
juvenile stages, reaching to adult phases that eventually migrate to, and inhabit
coral reefs (Dorenbosch et al. 2004; Kimirei et al. 2011). This was analysed by
Lara-Domínguez and Yáñez-Arancibia (1999) in mangroves of the Gulf of Mexico,
where different uses were evaluated for distinct lengths and abundance in juveniles
(Eucinostomus gula and Orthopristis chrysoptera) and adult species (Arius
melanopus and Lutjanus synagris) for different habitats; and these contribute to
the biomass generated by the ecosystem.

3.4 Fauna Diversity and Habitat Utilization

3.4.1 Terrestrial Invertebrates

Arthropods are the most diverse groups on the planet; so for mangroves, insects and
spiders are found in all structural levels of this ecosystem, from aerial roots to,
pneumatophores, trunks and canopies, hence, different guilds of these organisms can
be found at various trophic levels. Despite their great diversity and many functions,
little attention has been given to this fauna, and their role in mangrove ecosystems.
Although some studies have revealed that mangroves are important sites for
feeding and refuge of terrestrial invertebrates (Clay and Andersen 1996; Musyafa
et al. 2020). Nagelkerken et al. (2008) mention three main trophic groups within the
insect community: (1) herbivore insects that feed on leaves, flowers, seeds and any
other plant part (e.g. crickets, butterflies and bees); (2) saprophagous and saproxylic
invertebrates that feed on dead or decaying organic matter (e.g. flies and beetles) and
(3) parasitic or predatory insects (e.g. dragonflies, bedbugs and spiders).
Many terrestrial invertebrates are involved in key processes such as pollination
and nutrient recycling, in particular, detritivore organisms inhabiting sediments
(e.g. springtails, woodlouse). These organisms are an essential part of the food
web as they are the main food source for birds, fish and amphibians of the mangrove
ecosystems (Rajpar and Zakaria 2014).
Palynivore insects are classified within the herbivore group, and besides feeding
in the mangrove ecosystems, they are vital in maintaining the genetic variability of
mangroves. Worldwide, insects have been recorded in these ecosystems;
e.g. Diptera, Hymenoptera and Lepidoptera (Sánchez-Nuñez and Mancera-Pineda
2012), with up to 25 species in Rhizophora mangle, Avicennia germinans and
Laguncularia racemosa in mangroves from the Caribbean.
Veenakumari et al. (1997) found 276 species of insects in mangroves off islands
of the Indian Ocean, of which 197 were herbivores, 43 parasitic, and 36 predatory
72 D. Arceo-Carranza et al.

species. A similar composition in diversity and abundance was reported for


mangroves from Thailand by Murphy (1990).
In Zanzibar, Mchenga and Ali (2013) reported 103 species of orders in which
hymenopterans were the most abundant. In Indonesian mangroves, Musyafa et al.
(2020) recorded 10 insect orders, where coleopterans were the most abundant. In
New Zealand, Dencer-Brown et al. (2020) recorded a total of four classes of
arthropods, where spiders were the most abundant.
In the south of the Gulf of Mexico, García-Martínez et al. (2019) recorded
11 orders of arthropods within a mangrove system formed by A. germinans and
R. mangle, where spiders and ants from the Crematogaster genus were the most
relevant in number and abundance. In Brazil, 22 ant species have been reported, the
most common genus being Camponotus and Solenopsis (Gissel 2010); in Australia,
16 species have been registered (Clay and Andersen 1996), including reports of
Polyrhachis sokolova, an endemic species of the area that are capable of building
their nests in sludge. Terrestrial arthropods are an important group in terms of energy
transfer, with links to different groups of terrestrial and aquatic vertebrates, whether
these arthropods obtain energy directly from detritus, or by adding it through
herbivory. As observed, fish predation on the aquatic larval phase of mosquitoes
serves as a natural control, with some studies showing low mosquito densities when
fish densities are high (Griffin and Knight 2012).

3.4.2 Aquatic Invertebrates

Mangroves are important habitats for aquatic invertebrate populations such as


crustaceans, molluscs and worms (Zakaria and Rajpar 2015). Crustaceans such as
shrimps, crabs and prawns contribute to the recycling of organic matter through
bioturbation of mangrove sediments, but other groups such as barnacles can cause
considerable damage to plants since they grow on aerial roots and pneumatophores
(e.g. Euraphia, Elminius and Hexaminius). Apparently, only desiccation and an
increase in temperature can limit the colonization of barnacle larvae on the root
systems (Ross and Underwood 1997).
Isopods known as a burrowing crustacean group (e.g. Sphaeroma terebrans and
S. peruvianum) have also damaged mangroves of the Atlantic, Caribbean and east
Pacific regions, as juveniles and adults embed themselves in roots and stems,
affecting root growth (Brooks and Bell 2005; Davidson et al. 2016). There are
other crustaceans that temporarily inhabit mangrove waters during a certain phase
of their life cycle, for example, the Caribbean spiny lobster (Panulirus argus) that
uses the mangrove habitat as nursery areas. Lobsters, shrimps and prawns migrate
from the mangrove once they have grown, with adults remaining in the mangrove
vegetation, only if it is adjacent to a coral reef. However, if several species are
present, then, they will tend to distribute themselves differently (Hing et al. 2014).
Crabs occupy a great diversity of habitats within the complex mangrove structure.
Some species have obligatory relationships with mangroves while others are facul-
tative, allowing them to occupy seagrass beds and algae located within the mangrove
3 Mangroves as Feeding and Breeding Grounds 73

system (Ravichandran et al. 2007). Six of the 30 Brachyura families of crabs are
frequent colonizers of mangroves; the most common families are Mictyridae,
Grapsidae, Geocarcinidae, Portunidae, Ocypodidae and Xanthidae, with around
127 species. The Ocypodidae family (8 of the 19 genera are the most numerous of
at least 80 species) and the Grapsidae family (the Sesarma genus has 60 species
inhabiting mangroves) where 30 species are found in the Indo-Malaysian region,
16 on the coasts of Africa, 14 in Australia and 5 in tropical America (Diele et al.
2010).
Among the Anomura crustaceans, hermit crabs from the Diogenidae family
(e.g. Clibanarius spp. and Coenobita spp.) form numerous groups as they climb
and rest on the mangrove roots during the immersion period (Teoh et al. 2014). Crab
behaviour in mangroves varies, with nocturnal activity for some, perhaps avoiding
high temperatures or predators (e.g. Coenobita rugosus and C. cavipes), whereas
others are active all day (Barnes 2001). Mangrove crabs are divided into groups
according to their feeding mode; Uca and Macrophthalmus spp. are detritivores;
Scylla serrata are opportunistic diggers, and Thalamita crenata are predators that
occupy the marine edges of the mangrove, as they feed on slowly moving bivalves
and crustaceans (Cannicci et al. 2008). Herbivore crabs are another group of
consumers that feed directly on mangrove leaf litter, for instance, in the diet of
Sesarmidae crabs from Thailand up to 82% of leaf litter was found to be consumed
(Diele et al. 2010).
These crustaceans may be greatly important for leaf degradation and transport of
organic matter to other marine and estuarine habitats (Ravichandran et al. 2006;
Andreetta et al. 2014). Some studies have indicated that crabs play a key role in
consuming mangrove propagules. Crabs from the Grapsidae family in Malaysia and
Australia consume 95% of the post-dispersion propagules (Sousa et al. 2007), such
herbivore influences explaining the pattern of distribution and composition of
mangroves in the intertidal zone; at the same time determining the assemblages of
associated fauna, and successive stages of vegetation (Lindquist et al. 2009).
Molluscs are considered a dominant aquatic group in the mangrove fauna. They
are an important link for the transfer of energy from organic matter produced there to
secondary consumers such as invertebrates, mammals, fish and birds (Alfaro 2006).
The nature of the mollusc community is strongly influenced by physical conditions;
they live inside, above the sediment, and are firmly attached to the roots.
In mangroves from China, the density and biomass of 52 mollusc species were
higher at high tide, decreased with depth, and abundance was more related to an
increase in salinity (Printrakoon et al. 2008). Mollusc fauna associated with
mangroves is mostly composed of bivalves and gastropods, but other groups, more
marine based, are also found in mangroves, such as nudibranchs, chitons and
scaphopods (Kathiresan and Bingham 2001).
Mangrove molluscs can be divided into three groups: (1) native to the habitat
(e.g. Cerithidea, Terebralia and Nerita); (2) facultative molluscs (e.g. Littoraria and
Crassostrea) and (3) migratory molluscs (e.g. Nerita and Clypeomorus) (Irma and
Sofyatuddin 2012). Molluscs occupy all level of the food web, but detritus and filter
feeders predominate (Cannicci et al. 2008). Intertidal zones have been documented
74 D. Arceo-Carranza et al.

to function as critical habitats of some organisms (Krauss et al. 2008), since they
facilitate aggregation of other species that form an association, such as Rhizophora
mangle and the Crassostrea virginica oyster (Aquino-Thomas and Proffitt 2014).

3.4.3 Amphibians

Amphibians play an important role in the food web of terrestrial and aquatic
ecosystems. Due to their predatory behaviour they act as primary and secondary
consumers of insects, some of which are crop pests or disease vectors (Behangana
2004). Additionally, many amphibian species are considered as biomarkers, since
changes in their population are observed when disturbances occur in their habitat
(Welsh Jr and Ollivier 1998). However, amphibian diversity in mangroves is
comparatively low relative to other vertebrate groups (Alfred and Ramakrishna
2004).
Most amphibians have a limited osmoregulation capacity and are particularly
sensitive to saltwater, which generally restricts their presence in brackish and salt
environments (Gomez-Mestre et al. 2004). Approximately, 144 amphibian species
capable of tolerating saltwater (between 0.5 and 32 ppt) and brackish conditions
have been reported (Wu and Kam 2009; Hopkins and Brodie Jr 2015). However,
only 26 amphibian species have been globally recorded in mangrove ecosystems
(Rog et al. 2017). Among the species recorded in these ecosystems, only adult stages
have been observed, thus, it could be assumed that they inhabit mangroves only in
certain seasons (Hopkins and Brodie Jr 2015). The facultative dependence on
mangroves shown by these species is probably due to them searching for, food
resources, reproduction sites, dispersion pathways between primary habitats, and/or
as a temporary refuge against biotic (e.g. predators, competition), abiotic
(e.g. extreme temperatures, desiccation) or anthropogenic stress (Rog et al. 2017).
Amphibians can tolerate salinity conditions typical of mangrove ecosystems.
They are predators in these environments feeding on small organisms such as insects
(e.g. beetles, bees, ants, termites and crickets), snails, small frogs, shrimp and fish
(Rajpar and Zakaria 2014). Frogs from the Fejervarya genus (located in southeast
Asia) (Yodthong et al. 2019) have successfully colonized mangroves and are closely
linked to them due to their high tolerance to brackish waters (up to 35 ppt) and show
specialized feeding on crabs (Wright et al. 2004; Hopkins and Brodie Jr 2015). Other
authors have observed larvae of these frog species, including reports of egg laying in
these ecosystems (Uchiyama et al. 1990). The commonly known cane toad (Rhinella
marina) is another species linked to these habitats (Rajpar and Zakaria 2014); the
larval presence of this species in mangroves suggests a similar pattern regarding
salinity tolerance. In other species of frogs, larvae and egg laying have been
observed such as for Smilisca baudinii, Trachycephalus typhonius (both from
Costa Rica; Sasa et al. 2009) and Leptodactylus macrosternum (in mangroves
from Ceara, northeast Brazil; Ferreira et al. 2019). However, the latter of these
species has been reported for the rainy season, and authors speculate that low salinity
3 Mangroves as Feeding and Breeding Grounds 75

concentrations, characterized by this season, allow for their presence (Ferreira et al.
2019).
The presence of larva and/or egg laying suggests that some individuals, or
populations of the species mentioned above, complete their life cycle within
mangroves; species that have records in other ecosystems, such that they are
regarded as facultative inhabitants of mangroves (Rog et al. 2017). On the contrary,
the distribution of the Caribbean robber frog (Eleutherodactylus caribe) is strictly
restricted to mangroves. This species was first documented for a marsh dominated by
red mangroves (Rhizophora mangle) on the western side of the Tiburon peninsula in
Haiti (10 males, 4 females and 2 juveniles). Frogs were observed on trees of
R. mangle that were completely flooded by brackish water, with no epiphytes, or
other elements where this species could potentially lay eggs. Authors of this result
concluded that mangroves could be the preferred habitat for E. caribe, since other
wetland communities without mangroves, also analysed, had no sightings of
E. caribe. To date, this species, worldwide, is considered the only endemic amphib-
ian species, or with a restricted distribution to mangrove ecosystems (Luther and
Greenberg 2009).

3.4.4 Reptiles

Reptiles are an essential component in mangrove ecosystems, as their unique life


history and diverse role in food webs are as, prey, predators, herbivores, seed
dispersers and commensal species (Raxworthy et al. 2008). This group constitutes
a large percentage of the fauna biomass, and their presence is vital for the proper
functioning of many ecological processes (Campbell and Campbell 2000). They also
serve as biomarkers of environmental health (Read 1998; Raxworthy et al. 2008) due
to their sensitivity to habitat loss and degradation, contamination, disease and
climate change (Todd et al. 2010). The wide diversity of shapes, and the nature of
reptiles, has enabled them to conquer all continents (expect Antarctica) and to
inhabit a variety of diverse environments (Roll et al. 2017), including saltwater
and brackish environments (Schmidt-Nielsen and Fange 1958), such as mangroves
(Voris and Murphy 2012).
Currently 118 reptile species have been reported in these ecosystems, majority as
predators, since more than 90% are carnivores (Rog et al. 2017). Reptiles have
successfully colonized mangroves, temporally inhabiting, or relying on their
resources; these species come from independent evolutionary lines: freshwater
turtles, sea turtles, crocodiles, marine and terrestrial lizards and snakes (Nagelkerken
et al. 2008).
Some turtle species are known to rely on estuaries and other brackish
environments although little is known on their specific requirements. For example,
the northern river terrapin, Batagur baska (distributed in central and southwest
Asia), and the painted terrapin Batagur borneoensis (distributed in Thailand and
west Malaysia and Indonesia), both species classified as critically endangered
(IUCN 2020), feed mainly on riverside vegetation including mangrove fruits
76 D. Arceo-Carranza et al.

(Blanco et al. 1991). In other areas, various studies have documented the fidelity of
sea turtles to feeding areas within mangroves (Limpus and Limpus 2000; Godley
et al. 2002). Green sea turtles (Chelonia mydas) have a pelagic distribution during
the first 3 to 5 years, thereafter, recruiting occurs up to sexual maturity in coastal
waters, where they occupy a series of habitats for development, including the
mangroves (Makowski et al. 2006). These distribution changes coincide with onto-
genetic diet shifts from omnivore to herbivore, as they consume fruits, cotyledons
and propagules from Avicennia marina (Pendoley and Fitzpatrick 1999; Limpus and
Limpus 2000). Similarly, mangroves can indirectly provide food for turtles, since
algae, also a food source, grow on the mangrove roots, trunks and pneumatophores.
Rhizophora mangle has been reported important for foraging, and the development
of Kemp turtles (Lepidochelys kempii) (Schmid 2000). Finally, the diamondback
terrapin (Malaclemys terrapin rhizophorarum), which is native to the eastern and
southern saltmarshes of the USA, is considered an endemic species to mangroves
(Luther and Greenberg 2009). It mainly feeds on decapod crustaceans, bivalves,
gastropods, fish and the typical vegetation of this habitat (Tucker et al. 1995; Butler
et al. 2012).
Regarding crocodiles, most of the fish and marine animals they feed on use
mangroves for their reproduction and nesting sites. Notwithstanding, its importance
as a habitat varies depending on the species. The saltwater crocodile Crocodylus
porosus is closely associated to mangroves, and lays eggs in the vegetation of
adjacent ecosystems (Webb et al. 1977; Magnusson 1980). A study from Sri
Lanka reported that the population decrease in this species was correlated to an
increase in coastal runoff due to mangrove tree cuttings, and thus destruction of nests
and eggs (Santiapillai and de Silva 2001). According to Santiapillai and de Silva
(2001), prop roots of Rhizophora spp. provide an important structural shelter for
C. porosus juveniles. Other crocodile species are closely linked to mangroves
(Table 3.1).
Many lizards from geckos to iguanas inhabit intertidal mangroves. Some are
terrestrial species that enter the ecosystem in an opportunistic manner to access the
resources, while others such as monitor lizards are semi-aquatic with a distribution
that is more restricted to these environments (Nagelkerken et al. 2008). The littoral
whiptail-skink (Emoia atrocostata) from Southeast Asia and the Pacific is frequently
observed in mangrove habitats feeding on crabs, fish and insects (Voris and Murphy
2012). The oriental garden lizard (Calotes versicolor), widely distributed in Asia, is
also a common species of these ecosystems (Ghosh 2011).
Finally, within the reptile group, are the snakes. Snakes are by far the most
successful and diverse reptiles in marine and brackish habitats, with several species
found living in the mangroves (Voris and Murphy 2012). For instance, 19 species of
mangrove snakes have been reported in Nigeria (Luiselli and Akani 2002). To date,
at least 45 species associated with these ecosystems have been globally documented
(Rog et al. 2017); some only access mangroves occasionally, while others are regular
inhabitants. According to Luther and Greenberg (2009), worldwide, approximately
11 snake species are restricted to mangroves. Some species distinctive of North
America’s mangroves are Nerodia clarkii and Agkistrodon piscivorus conanti (Voris
3 Mangroves as Feeding and Breeding Grounds 77

Table 3.1 Main vertebrate species and their protection category (IUCN 2020) in mangrove
ecosystems
Common Conservation
Class name Specie status IUCN Reference
Mammals Vordermann’s Hypsugo DD Luther and
Pipistrelle vordermanni LC Greenberg (2009),
Northern Pipistrellus EN Barlow et al. (2011),
Pipistrelle westralis CR Rajpar and Zakaria
Proboscis Nasalis larvatus CR (2014), Adhya et al.
Monkey Bradypus EN (2011), Giesen et al.
Pygmy three- pygmaeus VU (2007), Nowak
toed Sloth Mysateles garridoi CR (2012), Galat-Luong
Garrido’s Mesocapromys VU and Galat (2005) and
Hutia angelcabrerai EN Nowak and Lee
Cabrera’s Panthera tigris EN (2011)
Hutia tigris VU
Bengal Tiger Prionailurus EN
Fishing Cat viverrinus VU
Sumatran Panthera tigris VU
Tiger sumatrae EN
Leopard Panthera pardus
Temminck’s Procolobus badius
Red Colobus temminck
Zanzibar Red Procolobus kirkii
Colobus Cebus capucinus
Colombian Tragulus nigricans
White-faced Lutrogale
capuchin perspicillata
Balabac Aonyx cinereus
Mouse Deer Trichechus
Smooth coated manatus
Otter
Asian small
clawed otter
Caribbean
manatee
Birds Mangrove Coccyzus minor LC Polidoro et al.
cuckoo Camarhynchus CR (2010), Buelow and
Mangrove heliobates LC Sheaves (2015),
Finch Amazilia buocardi NT Canales-Delgadillo
Mangrove Setophaga petechia VU et al. (2019), Zakaria
Hummingbird bryanti VU and Rajpar (2015)
Mangrove Vireo pallens EN and Rajpar and
Warbler Egretta rufescens LC Zakaria (2014)
Mangrove Egretta eulophotes LC
Vireo Leptoptilos LC
Reddish Egret javanicus
Chinese Egret Tringa guttifer
Lesser Phoenicopterus
Adjutant ruber
Spotted Halcyon
Greenshank senegaloides
(continued)
78 D. Arceo-Carranza et al.

Table 3.1 (continued)


Common Conservation
Class name Specie status IUCN Reference
American Haliaeetus
flamingo leucogaster
Mangrove
kingfisher
White-Bellied
Sea-eagle
Amphibians Mangrove frog Eleutherodactylus CR Luther and
Crab eating caribe LC Greenberg (2009),
frog Fejervarya LC Nagelkerken et al.
Cuban treefrog cancrivora LC (2008) and Rajpar
Cane toad Osteopilus and Zakaria (2014)
septentrionalis
Rhinella marina
Reptiles Mangrove Malaclemys VU Luther and
diamondback terrapin CR Greenberg (2009),
Terrapin rhizophorarum VU Rajpar and Zakaria
Northern River Batagur baska LC (2014), Macintosh
Terrapin Crocodylus VU and Ashton (2002)
Mugger palustris LC and Nagelkerken
Nile crocodile Crocodylus LC et al. (2008)
American niloticus LC
crocodile Crocodylus acutus VU
Common Caiman crocodilus LC
caiman Ephalophis greyae LC
North-western Myron
Mangrove Sea rischardsonii
Snake Ophiophagus
Richardson’s Hannah
Mangrove Varanus
Snake semiremex
King Cobra Trimeresurus
Rusty monitor purpureomaculatus
Mangrove Pit
Viper
Fish Mayan Cichlid Mayaheros LC Arceo-Carranza and
Silver Jenny urophthalmus LC Vega-Cendejas
Bay Anchovy Eucinostomus gula LC (2009), Arceo-
Flathead Anchoa mitchilli LC Carranza et al.
Mullet Mugil cephalus LC (2016), Faunce and
Common Centropomus EN Serafy (2006),
Snook undecimalis VU Huxam et al. (2004),
Yucatan Fundulus persimils LC Igulu et al. (2013),
Killifish Fundulus LC Verweij and
Giant Killifish grandissimus NT Nagelkerken (2007),
Grey Snapper Lutjanus griseus LC Zakaria and Rajpar
Great Sphyraena LC (2015), McDonald
Barracuda barracuda LC et al. (2009) and Able
Rainbow Scarus guacamaia LC (2005)
Parrotfish Lutjanus apodus LC
(continued)
3 Mangroves as Feeding and Breeding Grounds 79

Table 3.1 (continued)


Common Conservation
Class name Specie status IUCN Reference
Schoolmaster Gerres cinereus LC
Snapper Gerres LC
Yellowfin filamentosus LC
Mojarra Haemulon VU
Whipfin flavolineatum LC
Mojarra Ariopsis felis VU
French Grunt Sphoeroides LC
Hardhead Sea testudineus
Catfish Boleophthalmus
Checkered dussumieri
Puffer Bairdiella
Mud Skipper chrysoura
Silver croaker Megalops
Tarpon atlanticus
Northern Elops saurus
ladyfish Poecilia velifera
Sail fin Molly Cyprinodon
Yucatan artifrons
pupfish
Source: DD ¼ data deficient; LC ¼ least concern; NT ¼ near threatened; VU ¼ vulnerable;
EN ¼ endangered; CR ¼ critically endangered

and Murphy 2012). The latter of these species has been observed below the herons
and cormorants searching for food, at the same time, they nest in the intertidal zone
(Wharton 1966). In the Neotropics (a region with few studies on mangrove
associated fauna) some snake genera have been identified (e.g. Helicops, Hydrops,
Liophis and Tretanorhinus, inhabiting mangroves, although occasionally (Voris and
Murphy 2012). In Africa, the Grayia smythii and Crotaphopeltis hotamboeia species
have been documented for brackish waters, mangroves and freshwater (Luiselli and
Akani 2002). The relative frequency in which the African rock python (Python
sebae) has been reported in mangroves, in comparison to other habitats, suggests this
habitat could act as an important refuge or dispersal corridor for this species
(Nagelkerken et al. 2008). In Asia, the Indian python (Python molurus) and king
cobra (Ophiophagus hannah) are two examples of terrestrial snakes that commonly
enter the mangroves for foraging (Macintosh and Ashton 2002). In Australasia,
some sea snakes (Hydrophiidae family) access the mangroves during high tide,
while other species such as the Grey’s mud snake (Ephalophis greyae) have
conserved its mode of terrestrial movement, migrating to dry mangrove substrate
for fish foraging, during low tide (Storr et al. 2002). Other snakes from this region
depend on mangrove trees as a physical structure in the habitat, species such as
Myron richardsonii (Guinea et al. 2004) and Boiga dendrophila (Norhayati et al.
2009). The dog-faced water snake (Cerberus rynchops) (Lim et al. 2001), little file
snake (Acrochordus granulatus) (Gorman et al. 1981) and mangrove Pit viper
80 D. Arceo-Carranza et al.

(Trimeresurus purpureomaculatus) (Lim et al. 2001) are equally common in the


mangroves of Australasia (Rajpar and Zakaria 2014).

3.4.5 Mammals

Worldwide, mammals are a key component of biodiversity since they perform roles
that are vital for the proper functioning of most ecosystems, such as controlling the
population of species at lower trophic levels (McLaren and Peterson 1994); they are
also seed dispersers (Asquith et al. 1997), some large-size species are landscape
modifiers (Sinclair 2003), and many are top predators on land (Van Valkenburgh
1999) and aquatic ecosystems (Bowen 1997). Moreover, various mammals are
considered as flagship species, because they are a great attraction to the general
public, and thus helps the creation and maintenance of natural protected areas
(Câmara and Oliveira 2012).
In mangrove forests, mammals are the terrestrial vertebrates with the highest
species recorded (~320 species) (Rog et al. 2017), although only a few mammals are
exclusive to mangroves, since these ecosystems have extreme conditions that only a
few species can tolerate (Vanucci 2001).
Most mammal species reported in mangrove ecosystems temporarily live in them
by means of opportunistic circumstances, or may change their distribution so as to
shelter in mangroves, due to the destruction of their terrestrial habitats by humans
(Hogarth 2015). Due to widespread deforestation in the tropics, mangroves and other
wetland ecosystems that are not easily accessible to human populations have become
an important refuge for the persistence of various populations of threatened mammal
species (Nowak 2012). In West Africa, human disturbance has had a strong impact
on predators like the spotted hyena (Crocuta crocuta) and African civet (Civettictis
civetta), species that have been observed hunting for the Patas monkey
(Erythrocebus patas), harnessed bushbuck (Tragelaphus scriptus) and warthogs
(Phacochoerus africanus) in mangroves with more frequency than in their preferred
habitats (Galat-Luong and Galat 2007).
Similarly, poaching has caused the disappearance of majority of the Javan
rhinoceros populations (Rhinoceros sondaicus) from their preferred habitats; this
species has currently found refuge in mangrove ecosystems (Macintosh and Ashton
2002).
Mangroves are also fundamental for the conservation of wild cat species
(Table 3.1). Primates are frequent inhabitants of mangroves; in West Africa,
(south of Senegal) guenons (Cercopithecus) feed on fiddler crabs (Uca tangeri)
and fruits, flowers and young leaves of Rhizophora. The Macaques monkeys
(Macaca) from southwest Asia feed in the mud looking for crabs and bivalves;
although this has been documented to have a negative impact on mangrove restora-
tion projects, since Macaques tend to uproot the Rhizophora propagules. The Surilis
monkeys (Presbytis) also in Southeast Asia feed mainly on mangrove leaves and
fruits. Lastly, the Proboscis monkeys (Nasalis larvatus) feed on large quantities of
leaves and fruits per day (Macintosh and Ashton 2002). The importance of
3 Mangroves as Feeding and Breeding Grounds 81

mangrove ecosystems for wild cats and primates of Africa and Asia has been
documented by Nowak (2012). The data shows mangroves being used by
39 primates and 2 felines (species and subspecies) from Africa, and 29 primates
and 18 felines from Asia.
Even though mangroves provide multiple benefits to several mammalian species,
mammals also return important services in these ecosystems. Besides being impor-
tant links in the food web, some species are key in transporting nutrients from
terrestrial to marine ecosystems through the use of mangroves (Reef et al. 2014).
In Australia, the role of herbivorous mammals and their nutrient input to the
mangroves were evaluated. On the one side, the relationship and presence of the
red kangaroo (Macropus rufus) and that of the Wallaroo (Macropus robustus)
showed high nutrient concentrations coming from terrestrial environments. On the
other hand, a positive relationship was observed between the presence of the black
fruit bat (Pteropus alecto) and nutrition levels, as well as for growth in mangrove
trees. A colony of thousands of P. Alecto was observed resting on the canopies of
Rhizophora stylosa, Ceriops tagal and Lumnitzera rosea mangroves, making inland
trips for fruit and nectar. Bats defecate where they rest, thus providing and
transporting nutrients coming from inland mangroves (Reef et al. 2014).
Mangroves around the world are of great importance to several mammal species
with aquatic lifestyles, such as dugongs, porpoises, dolphins and some whales. In
Sundarbans, Bangladesh, the South Asian River dolphin (Platanista gangetica) and
Irrawaddy dolphin (Orcaella brevirostris) are noticeable inhabitants of mangrove
ecosystems. Also found, are the Sea otters, another mammalian group that use the
mangrove to feed. In Southeast Asia, aquiculturist considered them as pests because
they steal their produce (Macintosh and Ashton 2002).

3.4.6 Birds

Most mangroves have experienced a loss in connectivity and decrease in habitat


heterogeneity, which has reduced the diversity in fauna, including land and aquatic
migratory bird populations (Mohd-Azlan et al. 2015; Amir 2018). Additionally,
there are opportunistic bird species that use the canopy and the aerial root system to
shelter from predators, for reproduction, or as feeding sites (Naranjo 1997).
Mangroves together with neighbouring environments offer the resources for nest
building, including a great diversity of potential prey. Although mangroves are an
important habitat for birds, few endemic species are found in these ecosystems
(Table 3.1); majority of the birds found in the mangroves are also found in other
habitats.
Species composition of birds has been used to indicate mangrove health
(Behrouzi-Rad 2014) and to evaluate the impacts of climate change and coastal
development (Ogden et al. 2014). In addition, this allows each site to be
characterized due to the similarity between sites being relatively low, which suggests
that physical characteristics such as habitat heterogeneity and vegetation, including
the flooding surface of each zone, allows for the presence of particular bird groups.
82 D. Arceo-Carranza et al.

Other species can use mangroves when their preferred habitats are unavailable or
because mangrove trees provide an additional habitat to bird populations that would
mainly be found in neighbouring habitats. Nonetheless, fewer mangrove bird species
are observed in highly disturbed sites as compared to those less disturbed (Mohd-
Taib et al. 2020). Results from Mohd-Azlan and Lawes (2011) show that the number
of specialized species depend on the mangrove patch area, which suggests that these
species are limited by resource availability.
The presence of aquatic birds in mangrove ecosystems, mainly piscivore guild,
has been used as a status indicator of habitat conservation. These birds are found at
the top of the food chain, and are susceptible to environment and habitat changes.
These groups reflect the conditions of the terrestrial environment, as well as that of
the aquatic environment since they rely on both environments for feeding, refuge
and reproduction (Catterall et al. 2012; Zakaria and Rajpar 2015; De Dios et al.
2019). Observations from Canales-Delgadillo et al. (2019) indicate that bird abun-
dance in tropical mangroves is more influenced by habitat condition
(i.e. hydroperiod and forest structure), rather than water quality variables. Although
without disregard, environmental characteristics possibly have an influence over
resource availability due to its effects on primary producers, as well as, the presence
of benthic communities, fish and crustaceans (that bird species feed on).
Birds mainly congregate on the soft bottom substrates found close to the borders
of the mangrove forests; this has been linked to their feeding behaviour. Sites far
from the canopy provide better feeding areas for wading birds (Curado et al. 2013).
Furthermore, open and semi-open areas are probably chosen to reduce the risk of
predation and increases feeding efficiency, since this type of substrate favours the
presence of aquatic invertebrates (Pomeroy 2006; Chacin et al. 2015).
The dynamics of bird populations depend on multiple environmental factors and
food availability (Goodsell 1990; Halse et al. 1993; Murkin et al. 1997); it is one of
the reasons why piscivorous birds are distributed according to prey availability
(Kerekes et al. 1997). Another factor of great importance is the vegetation structure,
where there is the likelihood that it can be used as a habitat for roosting, as a refuge,
and nesting sites (Zakaria and Rajpar 2015).
Some species of aquatic birds, mostly the Suliformes and Ciconiiformes orders,
choose wetlands for their nesting sites because it provides the necessary resources
for reproduction, nesting and breeding (Cairns and Kerekes 2000). The location of
colonies depends on the accessibility to nearby sites with the available food to satisfy
nutritional requirements of both parents and chicks (Buckley and Buckley 1980;
Cairns and Kerekes 2000). The coastal wetlands boarded by mangrove forests on the
northern Yucatan coast host colonies of aquatic birds with different feeding
strategies and whose success depends on the water depth. The shallowest area of
the wetland is particularly used by Platalea ajaja adults, the intertidal area is used by
Ardea alba, Egretta rufescens, E. thula and E. tricolor adults, the deepest area is
used by Phalacrocorax brasilianus adults, and the peripheral area is used by
Cochlearius cochlearius which displays a foraging bout where it attacks prey
starting from trunks or branches, unlike wading birds that feed directly in the water.
3 Mangroves as Feeding and Breeding Grounds 83

Overall, leg length acts as a limiting factor in resource partitioning. Birds with
longer legs have access to deeper areas, and thus have more fishing sites than smaller
birds that may be restricted to fishing areas. Due to their wide trophic plasticity,
herons and cormorants feed on a wide variety of prey, such as fish, crustaceans,
insects, amphibians, reptiles and other birds (Ramo and Busto 1993; Kushlan and
Hancock 2005; Nelson 2005); however, they are mostly considered piscivorous
because their diet heavily depends on fish (Kneib 1982; Britton and Moser 1982;
Barquete et al. 2008; Ẑydelist and Kontautas 2008). Therefore, the permanency of a
colony is dependent on sites that provide food high in nutritional quality, food such
as fish (Kushlan et al. 2002). These birds form a colony in an islet on the northern
coast of Yucatan to reproduce. Robles-Toral (2019) mentioned that mangrove
forests are exploited by aquatic bird colonies for nest building, because the com-
plexity of the vegetation structure protects nests from inclement weather and
predators, and showing a significant correlation between the abundance of nests of
the colony and vegetation cover. Thus, trees with greater coverage provide more
shelter from predators and inclement weather.

3.4.7 Fish

Mangroves and other shallow aquatic habitats are sites that have recorded densities
of large numbers of juvenile species, including invertebrates and fish groups,
observed higher in numbers than in nearby areas without vegetation. These
observations gave rise to the hypothesis that mangroves act as nurseries for many
species that migrate to other habitats once they become adults (Lee et al. 2014).
Beck et al. (2001) proposed three reasons to explain the great abundance of
crustaceans and fish juveniles within the mangroves: (1) food availability,
(2) reduced predation due to the presence of microhabitats found in shallow waters,
and turbidity which is high for nearby habitats that lack vegetation, and (3) the
complex physical structure created by the aerial and submerged mangrove roots.
These factors could act in synergy to favour the presence of juvenile organisms,
given that mangroves offer optimal conditions for their growth and survival. The
structure of mangrove plants creates a canopy that reduces light penetration and the
presence of fine sediments that increase turbidity, decreasing the interaction between
predator and prey (Lee 2008).
These hypotheses, without a doubt are important; the relative importance of each
depends on the specific conditions of each system, the type of mangrove and the
species present t (Pittman et al. 2004; Lugendo et al. 2006). Recent studies using
stable isotopes have shown that carbon comes from other primary producers such as
phytoplankton, microphytobenthos, macroalgae and epiphytes, which are necessary
so as to maintain trophic levels (Bouillon et al. 2008; Nagelkerken et al. 2008).
Carbon from leaf litter has been shown to enter the marine food web through serranid
and snapper fish species that visit mangroves to feed on crabs through a process
called “short-circuit”. Sheaves and Molony (2000) noted that a considerate amount
of mangrove production incorporated in Sesarmidae crabs is exported to
84 D. Arceo-Carranza et al.

neighbouring ecosystems once fish complete their incursion of the mangroves. The
low frequency in fish consumption by serranid and snapper species in Australian
estuaries (Sheaves and Molony 2000) supports theories that a reduction in predatory
pressure over ichthyofauna adds value to tropical mangroves as nursing sites for
many fish species.
For decades, studies have been done to describe the diet of fish inhabiting
mangrove ecosystems, all in an effort to understand trophic relationships and thus
expand on the biology of species, their ecological role in aquatic systems and the
factors that regulate their periodic presence and absence (Braga et al. 2012). Notably,
tidal cycles have been shown to promote fish movement although there are existing
differences imposed by the amplitude of the tide. In areas with small amplitudes,
mangrove ecosystems seem to function mainly as habitat shelters (Kimirei et al.
2013), however, in sites with greater tidal amplitudes, mangroves are used as feeding
sites during high tides by various fish species (Unsworth et al. 2007; Pülmanns et al.
2018; Loera-Pérez et al. 2020). However, these observed differences in fish feeding
patterns suggest that it is important to consider other factors, such as local flooding
regimes, time of flooding and the geomorphological context of the basin (Unsworth
et al. 2007; Pülmanns et al. 2018; Loera-Pérez et al. 2020).
Unequivocally, it is difficult to conclude on the importance of mangroves as
feeding sites, considering the logistical dilemmas in obtaining information from
areas difficult to access, or with spatial differences in the structure of the landscape at
different scales, and with uneven geomorphologies. Additionally, Igulu et al. (2013)
show that there are other sources of uncertainties related with: (1) terminology, given
that some studies inappropriately use the concepts “mangrove microhabitat” and
“mangroves”, (2) studies in tidal regimen, since these have been done in areas with
distinct tidal amplitudes and sampled at specific times of the tide, (3) the nature of the
species studied, leading to intrinsic variations of results depending on the
bioecological features of each species, (4) the spatial-temporal variability of a
dynamic system, that are mangroves, occurs at distinct scales that promotes changes
in the presence and abundance of fish species that use these habitats, and (5) the
study methods, which, even though, after decades of efforts in analysing stomach
contents and nutritional value of prey, it has not yet been standardized, making it
difficult to compare between species in distinctive sites, and within a specific time
frame, further reducing a potential meta-analysis (Amundsen and Sánchez-
Hernández 2019; Fonteles de Vasconcelos Filho et al. 2019).
Moreover, several fish species show ontogenic migrations between habitats, with
shifts in their diet, possibly due to patterns in prey distribution (Ramírez-Martínez
et al. 2016). The feeding activities shown by sexually immature individuals found in
the mangroves prove vital for this stage of their life cycle. However, it has been
shown that available prey in these systems do not always have a high nutritional
value, but not all fish migrations between mangroves and nearby ecosystems are
related to food (Igulu et al. 2013), hence, these habitats are important refuges thanks
to the architecture of mangrove roots. Thereby, understanding the processes and
linkages of ecological interactions between neighbouring coastal ecosystems is
fundamental. The importance of documenting the spatial and temporal variability
3 Mangroves as Feeding and Breeding Grounds 85

of fish communities in coastal ecosystems, particularly in mangroves, has been


highlighted by Nagelkerken et al. (2008). Nonetheless, environmental dynamics
imposes at particular stages are distinctive to time and space attempting to establish
the relationships between the variables that determine fish distribution (Bonilla-
Gómez et al. 2013). Techniques in direct observations, based on underwater video
monitoring, reveal that only a small number of species frequent to these different
tropical estuary habitats, actually make extensive use of mangroves (Sheaves et al.
2016), suggesting a re-evaluation of patterns in which fish use habitat.
An inventory of fish biodiversity and studies on their life history allows
the comprehension of the complexity and functioning of coastal food webs, and
the importance of connectivity between different habitats. It has been shown that the
reduction and fragmentation of mangrove and seagrass habitats, as well as loss of
connectivity between ecosystems, negatively affect productivity. There are studies
that have addressed the consequences of fisheries and ecosystem services would
suffer if these socio-environmentally fragile mangrove systems are affected. Con-
servation efforts should be directed towards protecting habitat patches instead of
single habitats, or small portions of an ecosystem since it is known that species move
throughout the borders of these ecosystems.
Studies comprising large extensions of mangroves done for coastal wetlands in
the southern Gulf of Mexico show the presence of piscivorous fish, benefitting from
the richness and abundance of fish prey (Arceo-Carranza and Chiappa-Carrara
2017). Icthyophagous species show significant values in diet overlap, which,
under conditions of low prey abundance, may trigger competition, and hence, add
a predatory pressure to juveniles of various marine fish species using the mangroves
as a nursery, and for feeding.
Migrations of organisms with sufficient swimming capacity to move between
adjacent coastal systems are common, and occur at different time scales. Many fish
use the mangrove systems differentially, resulting in species composition, abun-
dance and feeding variability throughout a 24 hour cycle, and annual cycle (Arceo-
Carranza et al. 2013). Consequently, many authors have indicated that these
migrations occur to maximize feeding opportunities, or reduce predation risk, but
without specifying the relative value each process has for each species (Zárate-
Hernández et al. 2012; Kruse et al. 2016).
Globally, the loss of mangrove coverage is an important subject, due to the
functions and services it provides. Coastal development and tourism infrastructure
are the main impacts disrupting water flow, and causing deforestation. Throughout
this chapter, descriptions have been provided on the diversity and functions
displayed by the mangrove fauna, including the role of indicator species in deter-
mining the status of habitat health (e.g. amphibians, birds and fish), necessary to
determine changes resulting from the effects of impacts on the mangroves. In recent
years, different groups of fauna have been evaluated, some to a certain degree of
impact, or fauna in the process of restoration, with the objective of proving ecologi-
cal hypotheses, as it pertains to the functions mangroves offer, i.e. for feeding and as
nurseries (Arceo-Carranza et al. 2016; García-Martínez et al. 2019; Hernández
Mendoza 2020).
86 D. Arceo-Carranza et al.

The important role of mangrove ecosystems merits a global and monumental


response, with efforts geared towards emphasizing the significance of these
ecosystems for both terrestrial and aquatic fauna.

References
Able KW (2005) A re-examination of fish estuarine dependence: Evidence for connectivity between
estuarine and ocean habitats. Estuar Coast Shelf Sci 64:5–17.
Adams ES (1994) Territory defense by the ant Azteca trigona: Maintenance of an arboreal ant
mosaic. Oecol 97: 202–208.
Adhya T, Dey P, Das U, Hazra P (2011) Status survey of Fishing Cat (Prionailurus viverrinus) in
Howrah and Hooghly, West Bengal. Intermediate report submitted to the small grants
programme, WWF, India. WWF. India.
Alfaro AC (2006) Benthic macro-invertebrate community composition within a mangrove seagrass
estuary in northern New Zealand. Estuar Coast Shelf Sci 66:97–110.
Alfred JRB, Ramakrishna (2004) Faunal Resources in Mangrove Ecosystem. Envis Forestry Bull
4:24–31.
Alongi DM (1987) Intertidal zonation and seasonality of meiobenthos in tropical mangrove
estuaries. Mar Biol 95: 447–458.
Alongi DM (1989) Benthic processes across mixed terrigenous-carbonate sedimentary facies on the
central great barrier reef continental shelf. Cont Shelf Res 9(7): 629–663.
Alongi DM, Christoffersen P, Tirendi F (1993) The influence of forest type on microbial-nutrient
relationships in tropical mangrove sediments. J Exp Mar Biol Ecol 171: 201–223.
Alongi DM (1994) The role of bacteria in nutrient recycling in tropical mangrove and other coastal
benthic ecosystems. Hydrobiology 285: 19–32.
Alongi D (2008) Mangrove forests: resilience, protection from tsunamis, and responses to global
climate change. Estuar Coast Shelf Sci 76:1–13. https://doi.org/10.1016/j.ecss.2007.08.024.
Amir AA (2018) Mitigate risk for Malaysia’s mangroves. Science 359(6382): 1342–1343
Amundsen PA, Sánchez-Hernández J (2019) Feeding studies take guts–critical review and
recommendations of methods for stomach contents analysis in fish. J Fish Biol 95: 1364–1373.
Andreetta A, Fusi M, Cameldi I, Cimo F, Carnicelli S, Cannicci S (2014) Mangrove carbon sink.
Do burrowing crabs contribute to sediment carbon storage? Evidence from a Kenyan mangrove
system. J Sea Res 85:524–533. https://doi.org/10.1016/j.seares.2013.08.010.
Arai H, Yoshioka R, Hanazawa S, Minh VQ, Tuan VQ, Tinh TK, Phu TQ, Jha CS, Reddy SR,
Dadhwal VK, Mano M, Inubushi K (2016) Function of the methanogenic community in
mangrove soils as influenced by the chemical properties of the hydrosphere. Soil Sci Plant
Nut 62(2):150–163. https://doi.org/10.1080/00380768.2016.1165598.
Arceo-Carranza D, Vega-Cendejas ME (2009) Spatial and temporal characterization of fish
assemblages in a tropical coastal system influenced by freshwater inputs: northwestern Yucatan
peninsula. Rev Biol Trop57(1-2):89–103.
Arceo-Carranza D, Vega-Cendejas ME, Hernández de Santillana M (2013) Day and night trophic
variations of dominant fish species in a lagoon influenced by freshwater seeps. J Fish Biol
82:54–68. https://doi.org/10.1111/j.1095-8649.2012.03463.x.
Arceo-Carranza D, Gamboa E, Teutli-Hernández C, Badillo-Alemán M, Herrera Silveira JA (2016)
Los peces como indicador de restauración de áreas de manglar en la costa norte de Yucatán. Rev
Mex Biodiv 87:489–496. https://doi.org/10.1016/j.rmb.2016.03.001.
Arceo-Carranza D, Chiappa-Carrara X (2017) Feeding ecology of juvenile marine fish in a shallow
coastal lagoon of southeastern Mexico. Lat Am J Aquat Res 43(4): 621–631.
Aquino-Thomas JS, Proffitt E (2014) Oysters Crassostrea virginica on red mangrove Rhizophora
mangle prop roots: Facilitation of one foundation species by another. Mar Ecol Prog Ser. https://
doi.org/10.3354/meps10742.
3 Mangroves as Feeding and Breeding Grounds 87

Asquith NM, Wright SJ, Clauss MJ (1997) Does mammal community composition control recruit-
ment in neotropical forests? Evidence from Panama. Ecology 78:941–946.
Azam F, Fenchel T, Field JG, Gray JS, Meyer-Reil, Thingstad F (1983) The ecological role of
water-column microbes in the sea. Mar Ecol Prog Ser 10: 257–263.
Balakrishnan S, Srinivasan M, Mohanraj J (2014) Diversity of some insect fauna in different coastal
habitats of Tamil Nadu, southeast coast of India. J Asia-Pacific Biodiv. https://doi.org/10.1016/j.
japb.2014.10.010
Balakrishnan S, Srinivasan M, Santhanam P (2018) Insect Fauna of Pitchavaram and Parangipettai
Mangroves of Southeast Coast of India. Proc Zool Soc. https://doi.org/10.1007/s12595-016-
0182-y.
Barbier EB, Hacker SD, Kennedy C, Koch EW, Stier AC, Silliman BR (2011) The value of
estuarine and coastal ecosystem services. Ecol Monog 81:169–193.
Barlow AC, Smith JL, Ahmad IU, Hossain AN, Rahman M, Howlader A (2011) Female tiger
Panthera tigris home range size in the Bangladesh Sundarbans: the value of this mangrove
ecosystem for the species’ conservation. Oryx 45:125–128.
Barnes DKA (2001) Hermit crabs, humans and Mozambique mangroves. Afr J Ecol. https://doi.org/
10.1046/j.1365-2028.2001.00304.x
Barquete V, Bugoni L, Vooeren CM (2008) Diet of neotropic cormorant (Phalacrocorax
brasilianus) in an estuarine environment. Mar Biol 153(3):431–443. https://doi.org/10.1007/
s00227-007-0824-8.
Beck MW, Heck KL Jr, Able KW, Childers DL, Eggleston DB, Gillanders BM, Halpern B, Hays
CG, Hoshino K, Minello TJ, Orth RJ, Sheridan PF, Weinstein MP (2001) The identification,
conservation, and management of estuarine and marine nurseries for fish and invertebrates.
BioScience 51: 633–641.
Behangana M. (2004) The diversity and status of amphibians and reptiles in the Kyoga Lake Basin.
Afr J Ecol 42:51–56. https://doi.org/10.1111/j.1365-2028.2004.00461.x.
Behrouzi-Rad T (2014) Population dynamic and species diversity of wintering waterbirds in
mangroves wetland (Persian Gulf) in 1983 and 2013. Int J Mar Sci 4(63):1–7. https://doi.org/
10.5376/ijms.2014.04.63.
Blanco S, Behler J, Kostel F (1991) Propagation of the Batagurine Turtles Batagur baska and
Callagur borneoensis at the Bronx Zoo. In: Beaman KR, Caporaso F, McKeown S, Graff MD
(ed) Proceedings of the 1st international symposium on turtles and tortoises: conservation and
captive husbandry. pp. 63–65.
Bonilla-Gómez JL, Badillo M, López K, Gallardo A, Galindo C, Arceo D, Chiappa-Carrara X
(2013) Environmental influences on the abundance of dominant fishes in a very shallow tropical
coastal lagoon in Northwestern Yucatan Peninsula, Mexico. J Marine Sci Res Dev 3:118.
https://doi.org/10.4172/2155-9910.1000118.
Bouillon S, Koedam N, Raman AV, Dehairs F (2002) Primary producers sustaining macro-
invertebrate communities in intertidal mangrove forests. Oecol 130: 441–448.
Bouillon S, Connolly R, Lee SY (2008) Organic matter exchange and cycling in mangrove
ecosystems: recent insights from stable isotope studies. J Sea Res 59: 44–58.
Bowen W (1997) Role of marine mammals in aquatic ecosystems. Mar Ecol Prog Ser 158:
267–274.
Britton RH, Moser ME (1982) Size specific predation by herons and its effect on the sex-ratio of
natural populations fish Gambusia affinis Baird and Girard. Oecol 53: 14–151.
Braga RR, Bornatowski H, Vitule JRS (2012) Feeding ecology of fishes: an overview of worldwide
publications. Rev Fish Biol Fish 22: 915–929.
Brooks RA, Bell SS (2001) Colonization of a dynamic substrate: factors influencing recruitment of
the wood-boring isopod, Sphaeroma terebrans, onto red mangrove, Rhizophora mangle prop
roots. Oecol 127:522–532.
Brooks RA, Bell SS (2005) The distribution and abundance of Sphaeroma terebrans, a wood-
boring isopod of red mangrove (Rhizophora mangle) habitat within Tampa Bay. Bull Mar Sci.
76:27–46.
88 D. Arceo-Carranza et al.

Buckley FG, Buckley PA (1980) Habitat selection and marine birds. In: Burger J, Olla BL, Winn
HE (eds) Behavior of marine animals, Plenum Press, New York. pp. 69–112.
Buelow C, Sheaves M (2015) A birds-eye view of biological connectivity in mangrove systems.
Estuar Coast Shelf Sci 152: 33–43. https://doi.org/10.1016/j.ecss.2014.10.014.
Butler JA, Heinrich GL, Mitchell ML (2012) Diet of the Carolina diamondback terrapin
(Malaclemys terrapin centrata) in northeastern Florida. Chelonian Conserv Bi 11:124–128.
Cairns DK, Kerekes JJ (2000) Fish harvest by common loons and common mergansers in
Kejimkujik National Park, Nova Scotia, Canada, as estimated by bioenergetic modelling. In:
Comín FA, Herrera-Silveira JA, Ramírez-Ramírez J (eds): Limnology and aquatic birds.
Monitoring, modelling and management, Proceedings of the 2nd International Symposium on
Limnology and Aquatic Birds, Universidad Autónoma de Yucatán, Mérida, México.
pp. 125–135.
Câmara E, Oliveira L (2012). Mammals of Serra do Cipó National Park, southeastern Brazil. Check
List 8:355–359. https://doi.org/10.15560/8.3.355.
Camilleri JC (1992) Leaf-litter processing by invertebrates in a mangrove forest in Queensland. Mar
Biol 114: 139–145.
Campbell KR, Campbell TS (2000) Lizard contaminant data for ecological risk assessment. In: de
Voogt P (ed) Reviews of Environmental Contamination and Toxicology. Springer, Netherlands.
pp. 39–116.
Canales-Delgadillo JC, Perez-Ceballos R, Zaldivar-Jimenez MA, Merino-Ibarra M, Cardoza G,
Cardoso-Mohedano J (2019) The effect of mangrove restoration on avian assemblages of a
coastal lagoon in southern Mexico. PeerJ 7:e7493. https://doi.org/10.7717/peerj.7493.
Cannicci S, Burrows D, Fratini S, Smith TJ III, Offenberg J, Dahdouh-Guebas F (2008) Faunal
impact on vegetation structure and ecosystem function in mangrove forests: a review. Aquat Bot
89:186–200.
Catterall CP, Freeman AND, Kanowski J, Freebody K (2012) Can active restoration of tropical
rainforest rescue biodiversity? A case with bird community indicators. Biol Conserv 146(1):
53–61. https://doi.org/10.1016/j.biocon.2011.10.033.
Chacin DH, Giery ST, Yeager LA, Layman CA, Brian-Langerhans R (2015) Does hydrological
fragmentation affect coastal bird communities? A study from Abaco Island, The Bahamas. Wetl
Ecol Manag 23: 551–557. https://doi.org/10.1007/s11273-014-9389-8.
Clay RE, Andersen AN (1996) Ant fauna of a mangrove community in the Australian seasonal
tropics, with particular reference to zonation. Aust J Zool 44:521–533.
Curado G, Figueroa E, Sanchez MI, Castillo JM (2013) Avian communities in Spartina maritima
restored and non-restored salt marshes. Bird Study 60: 185–194.
Daniel PA, Robertson AI (1990) Epibenthos of mangrove waterways and open embayments:
community structure and the relationship between exported mangrove detritus and epifaunal
standing stocks. Estuar Coast Shelf Sci 31:599–619.
Davidson TM, Ruiz GM, Torchin ME (2016) Boring crustaceans shape the land–sea interface in
brackish Caribbean mangroves. Ecosphere. https://doi.org/10.1002/ecs2.1430.
Day JW, Crump BC, Kemp WM, Yáñez-Arancibia A (2013) Estuarine ecology. 2nd ed. John Wiley
and Sons, New Jersey.
De Dios Arcos C, Badillo-Alemán M, Arceo-Carranza D, Chiappa-Carrara X (2019) Feeding
ecology of the waterbirds in a tropical mangrove in the southeast Gulf of Mexico. Stud Neotrop
Fauna Environ 55(1): 1–9. https://doi.org/10.1080/01650521.2019.1682232.
Dencer-Brown AM, Alfaro AC, Bourgeois C, Sharma S (2020) The secret lives of mangroves:
Exploring New Zealand’s urban mangroves with integrated biodiversity assessments. Ocean
Coast Manag 191:1–16. https://doi.org/10.1016/j.ocecoaman.2020.105185.
Díaz MC, Smith KP, Rützler K (2004) Sponge species richness and abundance as indicators of
mangrove epibenthic community health. Atoll Res Bull. https://doi.org/10.5479/si.00775630.
518.1.
Diele K., Koch V., Abrunhosa F, Darlan JB, Simith (2010) The Brachyuran Crab Community of the
Caeté Estuary, North Brazil: Species Richness, Zonation and Abundance. In: Saint-Paul U,
3 Mangroves as Feeding and Breeding Grounds 89

Schneider H (eds) Mangrove Dynamics and Management in North Brazil. Ecological studies
vol 211. Springer, Berlin.
Dittmann S (2000) Zonation of benthic communities in a tropical tidal flat of north-east Australia. J
Sea Res 43: 33–51.
Dorenbosch M, Van Riel MC, Nagelkerken I, Van der Velde G (2004) The relationship of reef fish
densities to the proximity of mangrove and seagrass nurseries. Estuar Coast Shelf Sci 60: 37–48.
Eller RJ, Grassle JF (1992) Patterns of species diversity in the deep-sea as a function of sediment
particle size. Nature 360: 576–578.
Faunce CH, Serafy JE (2006) Mangroves as fish habitat: 50 years of field studies. Mar Ecol Prog Ser
318:1–18.
Feller IC, Mathis WN (1997) Primary herbivory by wood-boring insects along an architectural
gradient of Rhizophora mangle. Biotropica 29: 440–451.
Feller IC (2002) The role of herbivory by wood-boring insects in mangrove ecosystems in Belize.
Oikos 97:167–176.
Feller IC, Lovelock CE, Berger U, McKee KL, Joye SB, Ball MC (2010) Biocomplexity in
Mangrove ecosystems. Annu Rev Mar Sci 2:395–417.
Fernandes C, Fernandes CP, Barth MO (2011) Pollen analysis of honey and beebread derived from
Brazilian mangroves. Braz J Bot. https://doi.org/10.1590/S0100-84042012000100009.
Ferreira AC, Cascon P, Matthews-Cascon H (2019) Occurrence and egg-laying of Leptodactylus
macrosternum Miranda-Ribeiro, 1926 in mangrove habitat in Ceará, Northeast Brazil. Herpetol
Notes 12:865–868.
Fonteles de Vasconcelos Filho JI, Camargo Maia R, Salles R (2019) Desempenho dos diferentes
métodos de amostragem para caracterização da ictiofauna associada ao manguezal da praia de
Arpoeiras em Acaraú, Ceará. Arq Ciên Mar 52(1):81–98.
Galat-Luong A, Galat G (2005) Conservation and survival adaptations of Temminck’s red colobus
(Procolobus badius temmincki), in Senegal. Internat J Primat 26:585–603.
Galat-Luong A, Galat G (2007). Influence of anthropization on the distribution of large wildlife: the
mangroves, a refuge environment. In: Fournier A, Sinsin B, Mensah GA (eds) Quelles aires
protégées pour l’Afrique de l’Ouest? Conservation de la biodiversité et développement, IRD
Editions, Paris 568–569.
García-Martínez G, Arceo-Carranza D, Teutli Hernández C, Flores Rivero MA (2019) Diversidad
de artrópodos terrestres asociados a un ecosistema de manglar en estado conservado y en
proceso de restauración dentro de la ciénega de Progreso, Yucatán In: Molina Moreira
(ed) Manglares de América. Ed Compas Pp 83–93.
Ghosh D (2011) Mangroves. Resonance. 16:47–60.
Giesen W, Wulffraat S, Zieren M, Scholten L (2007). Mangrove guidebook for Southeast Asia.
Wageningen, FAO and Wetlands International.
Gissel NM (2010) Ants (Hymenoptera: Formicidae) of mangrove and other regularly inundated
habitats: Life in physiological extreme. Myrmecol News 14:113–121.
Godley B, Richardson S, Broderick A, Coyne M, Glen F, Hays G (2002) Long-term satellite
telemetry of the movements and habitat utilisation by green turtles in the Mediterranean.
Ecography 25:352–362.
Gomez-Mestre I, Tejedo M, Ramayo E, Estepa J (2004) Developmental alterations and osmoregu-
latory physiology of a larval anuran under osmotic stress. Physiol Biochem Zool 77:267–274.
Goodsell JT (1990) Distribution of waterbird broods relative to wetland salinity and pH in south-
western Australia. Aust Wildl Res 17: 219–229.
Gopal B, Chauhan M (2006) Biodiversity and Its Conservation in the Sundarban Mangrove
Ecosystem. Aquat Sci 68:338–354. https://doi.org/10.1007/s00027-006-0868-8.
Gorman GC, Licht P, McCollum F (1981) Annual reproductive patterns in three species of marine
snakes from the central Philippines. J Herpetol 335–354.
Griffin LF, Knight JM (2012) A review of the role of fish as biological control agents of disease
vector mosquitoes in mangrove forests: reducing human health risks while reducing environ-
mental risk. Wetl Ecol Manag. https://doi.org/10.1007/s11273-012-9248-4.
90 D. Arceo-Carranza et al.

Guerra-Castro E, Cruz Motta JJ, Conde JE (2011) Cuantificación de la diversidad de especies


incrustantes asociadas a las raíces de Rhizophora mangle L. en el Parque Nacional Laguna de la
Restinga. Interciencia. 36(12): 923–930.
Guinea M, Limpus C, Whiting S (2004) Marine snakes. In: Description of key species groups in the
northern planning area. National Oceans Office Hobart, Australia.
Halse SA, Williams MR, Jaensch RP, Lane JAK (1993) Wetland characteristics and waterbird use
of wetlands in south-western Australia. Wildl Res 20: 103–125.
Hernández Mendoza LC (2020) Ictiofauna como indicador de recuperación de las funciones
ecológicas de una zona de manglar en la Reserva de la Biosfera Sian Ka’an, Quintana Roo,
México. MSc thesis, Universidad Nacional Autónoma de México, México.
Hing MH, Chong VC, Sasekumar A (2014) Distribution and burrow morphology of three sympatric
species of Thalassina mud lobsters in relation to environmental parameters on a Malayan
mangrove shore. J Sea R. 95: 75–83. https://doi.org/10.1016/j.seares.2014.10.006.
Hogarth PJ (2015) The biology of mangroves and seagrasses. Oxford University Press.
Holguin G, Bashan Y (2007) La importancia de los manglares y su microbiología para el
sostenimiento de las pesquerías costeras. In R. Ferrera Cerrato, A. Alarcón (eds) Microbiología
agrícola: hongos, bacterias, micro y macrofauna, control biológico, planta organismo. Trillas,
México, México. p 239–252.
Hopkins GR, Brodie Jr ED (2015) Occurrence of amphibians in saline habitats: a review and
evolutionary perspective. Herpetol Monogr 29:1–27.
Hsieh HL (1995) Laonome albicingillum, a new fan worm species (Polychaeta: Sabellidae:
Sabellinae) from Taiwan. Proc Biol Soc Wash 108(1):130–131.
Huxam M, Kimani E, Augley J (2004) Mangrove fish: a comparison of community structure
between forested and cleared habitats. Estuar Coast Shelf Sci. 60:637–647.
Igulu MM, Nagelkerken I, van der Velde G, Mgaya YD (2013) Mangrove fish production is largely
fuelled by external food sources: a stable isotope analysis of fishes at the individual, species, and
community levels from across the globe. Ecosystems 16:1336–1352.
Irma D, Sofyatuddin K (2012) Diversity of Gastropods and Bivalves in mangrove ecosystem
rehabilitation areas in Aceh Besar and Banda Aceh districts, Indonesia. AACL Bioflux 5
(2):55–59.
Ismail T, Kassim A, Rahman A, Yahya AK, Webb CE (2018) Day biting habits of mosquitoes
associated with mangrove forests in Kedah, Malaysia. Trop Med Inf Dis. https://doi.org/10.
3390/tropicalmed3030077.
IUCN 2020. The IUCN Red List of Threatened Species. Version 2020-2 https://www.iucnredlist.
org. Downloaded on 09 July 2020.
Jenoh EM, Robert EM, Lehmann I, Kioko E, Bosire JO, Ngisiange N, Dahdouh-Guebas F, Koedam
N (2016) Wide ranging insect infestation of the pioneer mangrove Sonneratia alba by two insect
species along the Kenyan Coast. PloS One. https://doi.org/10.1371/journal.pone.0154849.
Kathiresan K, Bingham BL (2001) Biology of Mangroves and Mangrove Ecosystems. Adv Mar
Biol 40:81–251.
Kerekes JJ, Duggan M, Tordon R, Boros G, Bronkhorst M (1997) Abundance and distribution of
fish-eating birds in Kejimkujik National Park, Canada (1988–1994). In: Faragó S, Kerekes JJ
(eds) Limnology and waterfowl. Monitoring, modelling and management, Wetlands Interna-
tional Publication. p 211–227.
Kimirei IA, Nagelkerken I, Griffioen B, Wagner C, Mgaya YD (2011) Ontogenetic habitat use by
mangrove/seagrass-associated coral reef fishes shows flexibility in time and space. Estuar Coast
Shelf Sci 92:47–58.
Kimirei IA, Nagelkerken I, Mgaya YD, Huijbers CM (2013) The mangrove nursery paradigm
revisited: otolith stable isotopes support nursery-to-reef movements by Indo-Pacific fishes.
PLoS ONE 8(6): e66320 https://doi.org/10.1371/journal.pone.0066320.
Kneib RT (1982) The effects of predation by wading birds (Ardeidae) and blue crabs (Callinectes
sapidus) on the population size structure of the common mummichog Fundulus heteroclitus.
Estuar Coast Shelf Sci 14:159–165.
3 Mangroves as Feeding and Breeding Grounds 91

Krauss KW, Lovelock CE, McKee KL, Lopez-Hoffman L, Ewe SML, Sousa WP (2008) Environ-
mental driver in mangrove establishment and early development: a review. Aquat Bot
89:105127.
Kristensen E, King GM, Holmer M, Banta GT, Jensen MH, Hansen K, Bussarawit N (1994) Sulfate
reduction, acetate turnover and carbon metabolism in sediments of Ao Nam Bor mangrove,
Phuket, Thailand. Mar Ecol Prog Ser 109:245–255.
Kruse M, Taylor M, Muhando CA, Reuter H (2016) Lunar, diel, and tidal changes in fish
assemblages in an East African marine reserve. Reg Stud Mar Sci 3:49–57.
Kushlan JA, Hancock JA (2005) The Herons. Oxford University Press, Oxford.
Kushlan JA, Steinkamp MJ, Parsons KC, Capp J, Acosta-Cruz M, Coulter M, Davidson I,
Dickson L, Edelson N, Elliot R, Erwin RM, Hatch S, Kress S, Milko R, Miller S, Mills K,
Paul R, Phillips R, Saliva JE, Sydeman B, Trapp J, Wheeler J, Wohl K (2002) Waterbird
conservation for the Americas: the North American waterbird conservation plan, Version
1, Waterbird Conservation for the Americas Washington, DC.
Lara-Domínguez AL, Yáñez-Arancibia A (1999) Productividad secundaria, utilización del hábitat y
estructura trófica. In: Yáñez-Arancibia A, Lara-Domínguez AL (eds.). Ecosistemas de Manglar
en América Tropical. Instituto de Ecología A.C. México, UICN/ORMA, Costa Rica, NOAA/
NMFS Silver Spring MD, USA p. 153–166.
Lee SY (2008) Mangrove macrobenthos: assemblages, services, and linkages. J Sea Res 59(1–2):
16–29.
Lee SY, Primavera JH, Dahdouh-Guebas F, McKee K, Bosire JO, Cannicci S, Diele K, Fromard F,
Koedam N, Marchand C, Mendelssohn I, Mukherjee N, Record S (2014) Reassessment of
mangrove ecosystem services. Global Ecol Biogeogr 23:726–743.
Lim KK, Murphy D, Morgany T, Sviasothi N, Ng P, Soong B, Tan H, Tan H, Tan T (2001) A guide
to the mangroves of Singapore. Raffles Museum of Biodiversity Research, National University
of Singapore, and Singapore Science Centre, Singapore, 1.
Limpus CJ, Limpus DJ (2000) Mangroves in the diet of Chelonia mydas in Queensland, Australia.
Mar Turt Newsl 89:13–15.
Lindquist ES, Krauss KW, Green PT, O’Dowd DJ, Sherman PM, Smith TJ (2009) Land crabs as
key drivers in tropical coastal forest recruitment. Biol Rev. https://doi.org/10.1111/j.1469-
185X.2008.00070.x.
Loera-Pérez J, Hernández-Stefanoni J, Chiappa-Carrara X (2020) How do spatial and environmen-
tal factors affect the fish community structure in seasonally flooded karst systems? Lat Am J
Aquat Res 48(2):268–279.
Lugendo BR, Nagelkerken I, Van Der Velde G, Mgaya YD (2006) The importance of mangroves,
mud and sand flats, and seagrass beds as feeding areas for juvenile fishes in Chwaka Bay,
Zanzibar: gut content and stable isotope analyses. J Fish Biol 69(6): 1639–1661.
Lugo AE, Snedaker SC (1974) The Ecology of Mangroves. Annu Rev Ecol Syst 5: 39–64.
Luiselli L, Akani GC (2002) An investigation into the composition, complexity and functioning of
snake communities in the mangroves of south-eastern Nigeria. Afr J Ecol 40, 220–227.
Luther D, Greenberg R (2009) Mangroves: a global perspective on the evolution and conservation
of their terrestrial vertebrates. BioScience 59:602–612.
Macintosh D, Ashton E (2002) A Review of Mangrove Biodiversity Conservation and Manage-
ment. Centre for Tropical Ecosystems Research. University of Aarhus, Denmark.
Magnusson WE (1980) Habitat Required for Nesting by Crocodylus porosus (Reptilia:
Crocodilidae). Wildl Res 7:149–156.
Makowski C, Seminoff JA, Salmon M (2006) Home range and habitat use of juvenile Atlantic
green turtles (Chelonia mydas L.) on shallow reef habitats in Palm Beach, Florida, USA. Mar
Biol 148:1167–1179.
MacDonald J, Shahrestani S, Weis J (2009) Behavior and space utilization of two common fishes
within Caribbean mangroves: implications for the protective function of mangrove habitats.
Estuar Coast Shelf Sci 84(2):195–201.
92 D. Arceo-Carranza et al.

McLusky DS, Elliott M. (2004) The Estuarine Ecosystem. Ecology, threats and management. Third
edition. Oxford University Press. New York, USA.
Mchenga ISS, Ali AI (2013). Macro-fauna communities in tropical mangrove forest of Zanzibar
island, Tanzania. GJBB 2(1):260–266.
McLaren BE, Peterson RO (1994) Wolves, moose, and tree rings on Isle Royale. Science
266:1555–1558.
Mohd-Azlan J, Lawes MJ (2011) The effect of the surrounding landscape matrix on mangrove bird
community assembly in north Australia. Biol Conserv 144(9):2134–2141.
Mohd-Azlan J, Noske RA, Lawes MJ (2015) The role of habitat heterogeneity in structuring
mangrove bird assemblages. Diversity 7:118–136.
Mohd-Taib FS, Mohd-Saleh W, Asyikha R, Mansor MS, Ahmad-Mustapha M, Mustafa-Bakray
NA, Mod-Husin S, Md-Shukor A, Amat-Darbis ND, Sulaiman N (2020) Effects of anthropo-
genic disturbance on the species assemblages of birds in the back mangrove forests. Wetlands
Ecol Manage 28:479–494.
Mukherjee N, Sutherland WJ, Khan MNI, Berger U, Schmitz N, Dahdouh-Guebas F, Koedam N
(2014) Using expert knowledge and modeling to define mangrove composition, functioning,
and threats and estimate time frame for recovery. Ecol Evol 4(11):2247–2262.
Murkin HR, Murkinand EJ, Ball JP (1997) Avian habitat selection and prairie wetland dynamics: A
10-year experiment. Ecol Appl 7:1144–1159.
Murphy DH (1990) The recognition of some insects associated with mangroves in Thailand. In:
Mangrove Ecosystem Occasional Paper 7. UNDP /UNESCO, New Delhi.
Musyafa B, Hardiwinoto S, Ahbudin A (2020) Changes in insect biodiversity on rehabilitation sites
in the southern coastal areas of Java Island, Indonesia. Biodiversitas 21(1):1–7.
Nagelkerken I, Kleijnen S, Klop T, Van den Brand CJ, Cocheret de la Moriniere E, Van der Velde G
(2001) Dependence of Caribbean reef fishes on mangroves and seagrass beds as nursery
habitats: a comparison of fish faunas between bays with and without mangroves/seagrass
beds. Mar Ecol Prog Series 214:225–235.
Nagelkerken I, Blaber S, Bouillon S, Green P, Haywood M, Kirton L, Meynecke J-O, Pawlik J,
Penrose H, Sasekumar A (2008) The habitat function of mangroves for terrestrial and marine
fauna: a review. Aquat Bot 89:155–185.
Naranjo L G (1997) A note on the birds of the Pacific mangroves of Colombia. In: Kjerfve B, de La
Cerda LD, Diop HS (eds) Mangrove ecosystem studies in Latin America and Africa, UNESCO–
International Society for Mangrove Ecosystems, Forest Service Department of Agriculture,
Paris p 64–70.
Nelson B (2005) Pelicans, cormorants and their relatives: Pelecanidae, Sulidae, Phalacrocoracidae,
Anhingidae, Phaethontidae. Oxford University Press, Oxford.
Norhayati A, Shukor M, Juliana S, Wa WJ (2009) Mangrove flora and fauna of Klang islands
mangrove forest reserves, Selangor, Malaysia. MJS 28:275–288.
Nowak K (2012) Mangrove and peat swamp forests: refuge habitats for primates and felids. Folia
Primatol 83:361–376.
Nowak K, Lee PC (2011) Demographic structure of Zanzibar red colobus populations in unpro-
tected coral rag and mangrove forests. Internat J Primatol 32:24–45.
Oakes JM, Connolly RM, Revill AT (2010) Isotope enrichment in mangrove forests separates
microphytobenthos and detritus as carbon sources for animals. Limnol Oceanogr. https://doi.
org/10.4319/lo.2010.55.1.0393.
Odum WE, Heald EJ (1972) Trophic Analyses of an Estuarine Mangrove Community. Bull Mar Sci
22(3):671–738.
Ogden JC, Baldwin JD, Bass OL, Browder JA, Cook MI, Frederick PC, Frezza PE, Galvez RA,
Hodgson AB, Meyer KD, Oberhofer LD, Paul AF, Fletcher PJ, Davis SM, Lorenz JJ (2014)
Waterbirds as indicators of ecosystem health in the coastal marine habitats of Southern Florida:
2. Conceptual ecological models. Ecol indic 44:128–147.
Pascal PY, Gros O, Boschker HTS (2016) Temporal fluctuations in the trophic role of large benthic
sulfur bacteria. Food webs 7:20–28.
3 Mangroves as Feeding and Breeding Grounds 93

Pendoley K, Fitzpatrick J (1999) Browsing of mangroves by green turtles in Western Australia. Mar
Turtl Newsl 84:10.
Perry D (1988) Effects of associated fauna on growth and productivity in the red mangrove.
Ecology 69(4):1064–1075.
Pittman SJ, McAlpine CA, Pittman KM (2004) Linking fish and prawns to their environment: a
hierarchical landscape approach. Mar Ecol Prog Ser 283:233–254.
Polidoro BA, Carpenter KE, Collins L, Duke NC, Ellison AM, Ellison JC, Farnsworth EJ, Fernando
ES, Kathiresan K, Koedam NE, Livingstone SR, Miyagi T, Moore GE, Nam VN, Ong JE,
Primavera JH, Salmo SG III, Sanciangco JC, Sukardjo S, Wang Y, Yong JWH (2010) The Loss
of Species: Mangrove Extinction Risk and Geographic Areas of Global Concern. PLoS ONE 5
(4): e10095. https://doi.org/10.1371/journal.pone.0010095.
Pomeroy AC (2006) Tradeoffs between food abundance and predation danger in spatial usage of a
stopover site by western sandpipers, Calidris mauri. Oikos 112: 629–637.
Printrakoon C, Wells FE, Chitramvong Y (2008) Distribution of mollusks in mangrove at six sites
in the upper gulf of Thailand. Raffles Bull Zool 18:247–257.
Pülmanns N, Castellanos-Galindo GA, Krumme U (2018) Tidal-diel patterns in feeding and
abundance of armed snook Centropomus armatus from macrotidal mangrove creeks of the
tropical eastern Pacific Ocean. J Fish Biol 93: 850–859.
Rajpar MN, Zakaria M (2014) Mangrove fauna of Asia. In: Faridah-Hanum I, Latiff A, Hakeem K,
Ozturk M (eds). Mangrove Ecosystems of Asia. Springer, New York, p. 153–197.
Ramírez-Martínez GA, Castellanos-Galindo GA, Krumme U (2016) Tidal and diel patterns in
abundance and feeding of a marine-estuarine-dependent fish from macrotidal mangrove creeks
in the Tropical Eastern Pacific (Colombia). Estuaries Coasts 39:1249–1261.
Ramo C, Busto B (1993) Resource use by herons in Yucatan wetland during the breeding season.
Wilson Bull 105(4):573–586.
Ravichandran S, Kannupandi T, Kathiresan K (2006) Mangrove leaf litter processing by sesarmid
crabs. Cey J Sci (Bio Sci) 35:107–114.
Ravichandran S, Anthonisamy T, Kannupandi S, Balasubramanian T (2007) Habitat Preference of
crabs in Pichavaram mangrove environment, Southeast coast of India. J Fish Aq Sci 2:47–55.
Raxworthy C, Pearson R, Zimkus B, Reddy S, Deo A, Nussbaum R, Ingram C (2008) Continental
speciation in the tropics: contrasting biogeographic patterns of divergence in the Uroplatus leaf-
tailed gecko radiation of Madagascar. J Zool 275:423–440.
Read J (1998) Are geckos useful bioindicators of air pollution? Oecol 114:180–187.
Reef R, Feller IC, Lovelock CE (2014) Mammalian herbivores in Australia transport nutrients from
terrestrial to marine ecosystems via mangroves. J Trop Ecol 30:179–188.
Robertson AI, Alongi DM (1992) Tropical mangrove ecosystems, American Geophysical Union,
Washington DC.
Robles-Toral PJ (2019) Uso de hábitat de una colonia de aves acuáticas (Aves: Ardeidae,
Threskiornithidae y Phalacrocoracidae) en la Reserva Estatal de Ciénegas y Manglares de la
Costa Norte de Yucatán, México. MSc thesis, Universidad Nacional Autónoma de México,
México.
Rog SM, Clarke RH, Cook CN (2017) More than marine: revealing the critical importance of
mangrove ecosystems for terrestrial vertebrates. Divers Distrib 23:221–230.
Roll U, Feldman A, Novosolov M, Allison A, Bauer AM, Bernard R, Böhm M, Castro-Herrera F,
Chirio L, Collen B (2017) The global distribution of tetrapods reveals a need for targeted reptile
conservation. Nat Ecol Evol 1:1677–1682.
Ross PM, Underwood AJ (1997) The distribution and abundance of barnacles in a mangrove forest.
Aust J Ecol 22:37–47.
Sasa M, Chaves GA, Patrick LD (2009) Marine reptiles and amphibians. In: Wehrtmann IS, Cortés
J (eds) Marine Biodiversity of Costa Rica, Central America. Springer, p. 459–468.
Santiapillai C, de Silva M (2001) Status, distribution and conservation of crocodiles in Sri Lanka.
Biol Conserv 97:305–318.
94 D. Arceo-Carranza et al.

Sánchez-Nuñez DA, Mancera-Pineda JE (2012) Pollination and fruit set in the main neotropical
mangrove species from the southwestern Caribbean. Aquat bot 103:60–65.
Saravanakumar A, Rajkumar M, Serebiah JS, Thivakaran GA (2007) Abundance and seasonal
variations of zooplankton in the arid zone mangroves of gulf of Kachchh-Gujarat, westcoast of
India. Pak J Biol Sci. https://doi.org/10.3923/pjbs.2007.3525.3532.
Schmid JR (2000) Activity patterns and habitat associations of Kemp’s ridley turtles, Lepidochelys
kempi, in the coastal waters of the Cedar Keys, Florida. PhD Thesis University of Florida.
Schmidt-Nielsen K, Fange R (1958) Salt glands in marine reptiles. Nature 182:783–785.
Sinclair A (2003) The role of mammals as ecosystem landscapers. Alces 39:161–176.
Sheaves M, Johnston R, Baker R (2016) Use of mangroves by fish: new insights from in-forest
videos. Mar Ecol Prog Ser 549: 167–182.
Sheaves M, Molony B (2000) Short-circuit in the mangrove food chain. Mar Ecol Prog Ser
199:97–109.
Sousa WP, Kennedy PG, Mitchell BJ, Ordonez BM (2007) Supply-side ecology in mangroves: Do
propagule dispersal and seedling establishment explain forest structure? Ecol Monogr.
77:53–76.
Smith III TJ, Boto KG, Frusher SD, Giddins RL (1991) Keystone species and mangrove forest
dynamics: the influence of burrowing by crabs on soil nutrient status and forest productivity.
Estuar Coast Shelf Sci 33:419–432.
Srikanth S, Yamauchi LSK, Chen Z (2016) Mangrove root: adaptations and ecological importance.
Trees. https://doi.org/10.1007/s00468-015-1233-0.
Storr GM, Smith LA, Johnstone RE (2002) Snakes of Western Australia. Western Australian
Museum.
Teoh HW, Hussein MAS, Chong VC (2014) Influence of habitat heterogeneity on the assemblages
and shell use of hermit crabs (Anomura: Diogenidae). Zool Stud https://doi.org/10.1186/
s40555-014-0067-6.
Todd BD, Willson JD, Gibbons JW (2010) The global status of reptiles and causes of their decline.
In: Sparling D, Linder G, Bishop C, Krest S (eds.) Ecotoxicology of amphibians and reptiles,
CRC Press Boca Raton, Fl, p. 47–67.
Tucker AD, Fitzsimmons NN, Gibbons JW (1995) Resource partitioning by the estuarine turtle
Malaclemys terrapin: trophic, spatial, and temporal foraging constraints. Herpetologica
167–181.
Uchiyama M, Murakami T, Yoshizawa H (1990) Notes on the development of the crab-eating frog,
Rana cancrivora. Zool Sci 7:73–78.
Unsworth R, Bell J, Smith D (2007) Tidal fish connectivity of reef and sea grass habitats in the Indo-
Pacific. J Mar Biol Ass UK 87(5):1287–1296.
Van Valkenburgh B (1999) Major patterns in the history of carnivorous mammals. Annu Rev Earth
Planet Sci 27:463–493.
Vanucci M (2001) What is so special about mangroves? Braz J Biol 61:599–603.
Varon-Lopez M, Dias AC, Fasanella CC (2014) Sulphur-oxidizing and sulphate-reducing
communities in Brazilian mangrove sediments. Environ Microbiol. https://doi.org/10.1111/
1462-2920.12237.
Veenakumari K, Mohanraj P, Bandyopadhyay AK (1997) Insect herbivores and their natural
enemies in the mangals of the Andaman and Nicobar Islands. J Nat Hist 31:1105–1126.
Verweij M, Nagelkerken I (2007) Short and long-term movement and site fidelity of juvenile
Haemulidae in back reef habitats of a Caribbean embayment. Hydrobiologia 592(1): 257–270.
Vitousek PM, Sanford Jr RL (1986) Nutrient cycling in moist tropical forest. Ann. Rev. Ecol. Syst
17: 137–167.
Voris HK, Murphy JC (2012) Sampling marine and estuarial reptiles. In: McDiarmid RW, Foster
MS, Guyer C, Gibbons JW, Chernoff N (eds). Reptile biodiversity: standard methods for
inventory and monitoring. University of California Press, Berkeley, 192–196.
Wang M, Huang Z, Shi F, Wang W (2009) Are vegetated areas of mangroves attractive to juvenile
and small fish? Estuar Coast Shelf Sci 85:208–216.
3 Mangroves as Feeding and Breeding Grounds 95

Webb GJ, Messel H, Magnusson W (1977) The nesting of Crocodylus porosus in Arnhem Land,
northern Australia. Copeia 238–249.
Wells FE (1984) Comparative distribution of macromolluscs and macrocrustaceans in a north-
western Australian mangrove system. Aust J Mar Freshw Res 35:591–596.
Welsh Jr HH, Ollivier LM (1998) Stream amphibians as indicators of ecosystem stress: a case study
from California’s redwoods. Ecol Appl 8:1118–1132.
Wharton CH (1966) Reproduction and growth in the cottonmouths, Agkistrodon piscivorus
Lacépede, of Cedar Keys, Florida. Copeia 149–161.
Woodroffe CD (1992) Mangrove sediments and geomorphology. In: Robertson AI, Alongi DM
(eds) Tropical mangrove ecosystems, American Geophysical Union, Washington, DC. p 7–41.
Worm B, Barbier EB, Beaumont N, Duffy JE, Folke C, Halpern BS, Jackson J B, Lotze HK,
Micheli F, Palumbi SR, Sala E, Selkoe KA, Stachowicz JJ, Watson R (2006) Impacts of
biodiversity loss on ocean ecosystem services. Science. https://doi.org/10.1126/SCIENCE.
1132294.
Wright P, Anderson P, Weng L, Frick N, Wong WP, Ip YK (2004) The crab-eating frog, Rana
cancrivora, up-regulates hepatic carbamoyl phosphate synthetase I activity and tissue osmolyte
levels in response to increased salinity. J Exp Zool A Comp Exp Biol 301:559–568.
Wu CS, Kam YC (2009) Effects of salinity on the survival, growth, development, and metamor-
phosis of Fejervarya limnocharis tadpoles living in brackish water. Zool Sci 26:476–482.
Yodthong S, Stuart BL, Aowphol A (2019) Species delimitation of crab-eating frogs (Fejervarya
cancrivora complex) clarifies taxonomy and geographic distributions in mainland Southeast
Asia. ZooKeys 883:119–153.
Zakaria M, Rajpar MN (2015) Assessing the fauna diversity of Marudu Bay mangrove forest,
Sabah, Malaysia, for future conservation. Diversity 7:137–148. https://doi.org/10.3390/
d7020137.
Zárate-Hernández R, Castillo-Rivera M, Sanvicente-Añorve L, Ortiz-Burgos S (2012) Spatial, diel,
and seasonal changes in the fish community structure of a Mexican tropical estuary. Cienc Mar
38(4): 665–676.
Ẑydelist R, Kontautas A (2008) Piscivorous birds as top predators and fishery competitors in the
lagoon ecosystem. Hydrobiologia 611: 45–54.
Factors Influencing Mangrove Ecosystems
4
Joanna C. Ellison

Abstract

Mangroves occur in coastal settings of estuaries, deltas, lagoons, open coasts and
oceanic low islands. In these settings, mangrove attributes are influenced by
physical factors of temperature, coastal typology, ocean currents and land
barriers, wave action and sediment supply, river catchment discharge and sedi-
ment yield, and tidal range and inundation frequencies. Factors of gradients and
tidal ranges control the lateral extent of mangroves through inundation frequency,
and factors influencing accretion rates in the context of relative sea level change
can shift or eliminate mangrove extents over time. Mangroves are however
resilient systems within steady state equilibrium, that allows recovery from
minor perturbations. Factors influencing mangroves can however exceed tipping
points of tolerance, bringing a sudden change in ecosystem function and break-
down of equilibrium. Stressors that may cause critical reduction of mangrove
resilience are the impacts from humans, climate becoming significantly drier,
increased inundation, reduced sedimentation supply, and relative sea level rise.
Rehabilitation can be successful if ecological guidance on mangrove restoration
is followed, particularly topographic positioning with respect to tidal inundation
frequency factors. Understanding of the physical factors that influence mangrove
ecosystems that contribute to variation in processes, that result in spatial and
temporal differences in mangrove attributes, is essential to effective management.

J. C. Ellison (*)
School of Geography, Planning and Spatial Sciences, University of Tasmania, Launceston, TAS,
Australia
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 97


Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_4
98 J. C. Ellison

Keywords

Tidal inundation · Coastal typology · Oceanic currents · Temperature · Rainfall ·


Tipping points

4.1 Introduction

Mangrove forests occur most extensively on low latitude, low energy, sedimentary
shorelines, between intertidal elevations. Global extends are greatest in Asia
(38.7%), of a global 137,600 km2, with 20.3% in Central America and the Carib-
bean, 20% in Africa, 11.9% in Oceania, 8.4% in North America and 0.7% in
European Territories (Bunting et al. 2018). The IUCN Redlist recognises 70 species
of mangroves in 17 families (Polidoro et al. 2010), the majority of which grow in
mid to high intertidal positions indicating the stress of deeper water. Mangrove trees
are also more diverse in Asia/SW Pacific (Spalding et al. 2010). Adaptations of
mangroves to their intertidal, anaerobic, salty habitat are four main types: aerial
roots, waxy leaves, salt regulation and viviparity, which is the germination of seeds
before they drop into the low-oxygen salty water (Hutchings and Saenger 1987;
Tomlinson 2016).
Adapting general concepts from Odum (1972), mangrove ecosystems include all
of the organisms in the mangrove area interacting with the physical environment, so
that a flow of energy leads to exchange of materials between the living and
non-living parts of the system. The interactions with the physical environment are
the factors influencing mangroves. Ecosystems are a dynamic network consisting of
organisms, and interconnection with their environment and the definition includes
abiotic factors which include those factors that limit distributions (Putman and
Warren 1984).
Mangrove ecosystems therefore comprise mangrove trees, and floral associates,
particularly at the landward edge, as well as fauna linked through food-chains, in a
physical setting of sunlight, rainfall, tidal waters and sediments. The mangrove trees
are the structuring species group creating habitats variety for other biotas, and the
mangroves can biogeomorphically influence processes to become self-maintaining
systems (Lugo and Snedaker 1974; Ellison 2019). Factors influence the spatial
ranges of species, and particularly include the climatic factor of temperature, and
physical factors that in intertidal systems include tidal inundation and hydroperiod
(Sasmito et al. 2020), salinity, and position in the estuary (Valiela et al. 2018).
Dispersal capability of propagules, restricted availability of mangrove habitats and
adverse oceanic currents are major limiting factors of mangroves (Saenger et al.
2019).
Factors, attributes, characteristics and processes are all nouns that are applied to the
analysis of mangrove systems, but each has different meanings (Lincoln et al. 1982;
Macmillan 2020) (Fig. 4.1). Factors are an element or circumstance contributing to a
result, and other words for the same definition are component, constituent or
contributing element. A process is rather the series of actions or procedures or
mechanisms that cause change, such as through explanatory verbs that may apply to
4 Factors Influencing Mangrove Ecosystems 99

Fig. 4.1 Relationships between mangrove attributes and influencing factors and processes

Fig. 4.2 Schematic illustration of the factors influencing mangroves (Source: adapted from Ellison
2019)

the mangroves of respire, photosynthesise, grow, breed, propagate, accumulate,


metabolise, break down, transfer, adsorb, compress, or store. An attribute is a charac-
teristic, quality, property so is a descriptive term for results of factors influencing
processes. Hence the nouns all have associated meanings that help scientists under-
stand how ecosystems such as mangroves work (Fig. 4.1). This chapter reviews
physical factors that influence mangrove ecosystems that contribute to variation in
processes to result in spatial and temporal differences in mangrove attributes.
In the well cited article “Factors influencing biodiversity and distributional
gradients in mangroves”, Duke et al. (1998) review the influences of environmental
factors prevalent at global, regional, estuarine and intertidal scales. Environmental
factors of influence on mangrove distributions are climate, hydrology, geomorphol-
ogy and water conditions. Climatic factors include temperature, rainfall and storms,
and hydrological factors include tides, currents and sea level. Geomorphological
factors include sediments, catchment size and slope, and water condition factors
include salinity, nutrients, oxygen and pH (Duke et al. 1998). The extent of
influencing abiotic factors on mangroves are the climate (temperature and rainfall),
tidal inundation, salinity fluctuations, exposure to wind and waves, sea level,
currents variations and sedimentation (Duke 2016, 2017; Horstman et al. 2018).
These have notable and profound influences on the distributions, functioning and
well-being of mangrove habitats, along with human disturbance (Duke 2016;
Saenger et al. 2019). Limiting abiotic factors controlling mangrove attributes are
schematically illustrated in Fig. 4.2.
100 J. C. Ellison

4.2 Climate: Temperature and Rainfall

Mangrove global extents are usually attributed to temperature limits of sea surface
temperatures, based on mangrove distributions extending into higher latitudes with
prevalent warmer currents (Tomlinson 2016). Temperature influences both photo-
synthetic and respiratory processes, controlling a large number of internal processes
such as salt regulation, excretion and root respiration (Hutchings and Saenger 1987).
Mangroves are tolerant of high temperature stresses, but mangrove photosynthesis is
limited by low temperature at higher latitudes (Steinke and Naidoo 1991). Coastal
distributions in the sub-tropics are influenced by the origin of influencing oceanic
currents indicated by the 20  C isotherm, with frost incidence also a limiting factor
(Duke et al. 1998). Avicennia marina is the most tolerant of low temperatures in the
southern hemisphere, where the highest latitudes for mangrove distributions are
reached (Hutchings and Saenger 1987; Duke et al. 1998). Climatic factors primarily
result in mangrove latitudinal expansion and contraction (Van der Stocken et al.
2019), particularly temperature (Saenger et al. 2019).
Winter air temperature extremes are an especially important factor controlling
mangrove limits along the coasts of eastern North America, western Gulf of Mexico
and eastern Asia (Osland et al. 2017). Frequency of freeze events controls the
distribution, canopy height and coverage of Avicennia germinans at the species
northern limits in Louisiana, USA, showing the limiting factor of winter temperature
extremes (Osland et al. 2020). Freeze events cause mangrove mortality, leaf damage,
while warmer conditions favour more abundant and tall mangroves. Concomitant
abiotic factors are inundation, salinity and surface elevation gradients. However, for
mangroves to establish in productive marshes at high latitudes an entry point is
permitted by an external disturbance such as drought, fire, or storm damage causing
marsh dieback (Osland et al. 2020). Coastal wetlands in northeast Florida have
shown shifts between mangroves and saltmarshes at least 6 times in the last two
centuries due to decadal scale fluctuations in frequency and intensity of extreme cold
events (Cavanaugh et al. 2019). Northern extents of mangroves in the Gulf of

Mexico and Florida approach 30 N. Indicators of vegetation photosynthetic activity
and leaf area index at these mangrove limits are strongly controlled by climate
factors, with significant relationships with annual minimum temperature
(Cavanaugh et al. 2018). The influences on mangrove distributions of warm ocean

currents are evident from the occurrence of mangroves in Bermuda at 32 N (Ellison
1997).
At latitudinal extremes such as New Zealand, mangrove diversity is restricted by
factors such as temperature and day length, causing mean tree height to reduce with
increasing latitude (Horstman et al. 2018). Temperature stress causes reduced net
primary productivity resulting in smaller tree sizes, leading to stunted or dwarfed tree
architecture at higher latitudes. For example, at the northern latitudinal limit of

mangroves in the Atlantic at Bermuda 32 N, Rhizophora mangle reaches a maxi-
mum height of 6 m and Avicennia germinans 10 m (Ellison 1997), with strong
seasonal patterns in productivity in a mean annual temperature range of 17–29  C.
4 Factors Influencing Mangrove Ecosystems 101

This contrasts with close to the equator in central Africa, where mangroves including
these species can exceed 40 m in height (Ajonina et al. 2014).
As well as temperature limits on mangrove extents, along arid or semi-arid high
latitude coastlines, increase or decrease in rainfall can lead to contraction and
expansion. This is demonstrated from western North America, western Gulf of
Mexico and Western Australia, though west central Africa lacks data to confirm
the trend (Osland et al. 2017). The northern most arid mangrove margin of this

region at 20 N showed dwarf growth of Avicennia germinans at average height of
1.2 m, and poor regeneration rates (Otero et al. 2016), distributed among halophytic
vegetation.
Reduced rainfall and humidity cause reduction in mangrove diversity, photosyn-
thesis, productivity and growth rates (Waycott et al. 2011). In low rainfall areas, high
evaporation and soil salinity in higher intertidal mangrove zones cause stunting of
mangroves or replacement by hypersaline sand flats. During drought periods, exten-
sive mangrove mortality of such areas can occur (Duke et al. 2017; Lovelock et al.
2017). In the Gulf of Carpentaria, Australia, more than 74 km2 of mangroves died
during El Niño related drought, high temperatures and lower sea level (Duke et al.
2017; Duke 2020), mostly along the higher elevation zones. Similar losses have been
analysed from Kakadu, northern Australia (Ashbridge et al. 2019) this affecting the
landward Avicennia marina zone, and Western Australia (Lovelock et al. 2017)
where 20–30% increase in soil salinisation occurred.

4.3 Ocean Currents and Land Barriers

The limiting factor of oceanic currents and land barriers on dispersal of mangrove
propagules has influenced the species presence in different parts of the world, and
the biodiversity complements of mangrove areas (Duke et al. 1998; Spalding et al.
2010). Tertiary records of mangrove occurrence do not show clear trends of man-
grove migration with climate change, rather distributions developed over time are
due to biogeographic factors and habitat availability (Ellison 2008). Biogeographic
factors include oceanic connectivity, such as the western end of the Tethys Sea
became isolated with the enclosure of the Mediterranean by the collision of Africa
and southern Asia around 18 million years ago (Saenger 1998). At that time the
pantropical mangrove flora became disjunct and subsequently developed the differ-
ent species groups of the S.E. Asian and American centres of diversity (Ellison et al.
1999).
The Himalayan uplift and establishment of the Asian summer monsoon towards
the Late Neogene further affected the coastal dynamics affecting mangrove
distributions of the Indian subcontinent (Srivastava and Prasad 2019). Around
three million years ago the Panama Gap closed with the collision of North and
South America, to disjunct the mangroves on either side of the American land
masses (Saenger 1998).
Dispersal capability and the barriers to oceanic connection isolated the American
biodiversity centre from the Indo-West Pacific, demonstrating the combination of
102 J. C. Ellison

Fig. 4.3 World distributions of mangroves, and warm and cold currents that influence these
distributions. Source: Mangrove distributions are adapted from Spalding et al. (2010)

these other factors influencing mangrove distributions. The most widely distributed
genus in the Pacific Islands, Rhizophora shows high genetic variation across islands
indicating that long-distance dispersal of propagules is rare (Yan et al. 2016). Other
factors such as wind and water currents, estuarine geomorphology and coastal
topographies likely played important roles in facilitating or blocking the mangrove
gene flow (Yan et al. 2016). The distributions of mangroves across the world, and
influences of warm and cold currents and land barriers are illustrated in Fig. 4.3.

4.4 Coastal Typology Factors

Factors of catchment area and geomorphic setting are regional scale influences on
mangroves with geomorphological factors, including sediments and catchment size
(Duke et al. 1998). Geomorphology and coastal topographies are important in
facilitating or blocking the mangrove gene flow (Yan et al. 2016). At climate
extremes, lack of suitable habitats within the dispersal range of the propagules can
limit mangrove ranges (de Lange and De Lange 1994). These factors influence
mangroves through variation of sediment type and supply, and mangrove habitats.
Table 4.1 shows the prevalent categories of mangrove typological settings, each
illustrated by a location in Fig. 4.4.
Drainage basin size, gradients and land uses influence fluvial runoff, solutes and
sediment yield. Suspended sediment in river discharge allows coastal progradation
and mangrove substrate accretion, and if abundant can cause mangroves to become
opportunistic (Ellison 2019). Catchment geology and land uses influence sediment
yield, with climate determining weathering rates and types, rainfall characteristics
influencing runoff variability, vegetation and soil cover influencing slope hydrology,
and river channel erodibility (Tedford and Ellison 2018). Sediment yield can be
increased by catchment vegetation clearance to increase overland flow, riparian
vegetation disturbance that increases river bank erodibility, and increase in flow
4 Factors Influencing Mangrove Ecosystems 103

Table 4.1 Coastal typologies of mangroves


Deltas and Wave-dominated Oceanic karst or
estuaries, river Estuarine tidal berm or barrier, small islands,
dominated creeks, strong sheltering low lacking fluvial
Typology distributaries tidal currents energy lagoons discharge
Mangrove Seaward edge and Tidal creek Inside low energy Fringing or
locations river distributaries margins and lagoons basin
elongated islands
Sediment Catchment Catchment Sand barriers, Autochthonous
supply discharged allochthonous autochthonous mangrove
type allochthonous sediment mangrove fringed basins and
sediment, flood reworked by tidal lagoons fringes
deposits currents
Examples North Rufiji delta, Fly delta, Papua Pichavaram, NE Pipon Island,
shown in Tanzania New Guinea Cauvery River Queensland,
Fig. 4.4 Delta, India Australia

Fig. 4.4 Examples of coastal typologies of mangroves, imagery is adapted from Google Earth Pro
2020. (a) North Rufiji delta, Tanzania. (b) Fly delta, Papua New Guinea. (c) Pichavaram
mangroves, Eastern India. (d) Pipon Island, Queensland, Australia

velocities caused by drainage of wetlands and channel straightening. These factors


influence mangrove settings in estuarine, deltaic and riverine typologies (Table 4.1),
to cause coastal progradation (Fig. 4.4A and B) that can be offset by relative sea
level rise.
104 J. C. Ellison

Wave-dominated coastlines are concentrated in the mid-latitudes, and feature


intermittently closed lagoons (McSweeney et al. 2017). The strength of wave
processes relative to tidal processes builds shore-parallel barriers (such as a spit or
berm) which allow the sheltered habitats that mangroves favour as demonstrated in
Fig. 4.4C. Analysis of Australian estuaries showed mangroves are most extensive in
tide-dominated estuaries and deltas, and smallest in wave-dominated estuaries
(Mahoney and Bishop 2018). The area increased with tidal rage, estuary entrance
width and catchment area. By contrast, mangroves of New Zealand are mostly in
wave-dominated settings, or river dominated, or a combination of both (Horstman
et al. 2018).
Oceanic marine dominated mangrove settings have no fluvial sediment supply,
and freshwater is limited to rainfall and groundwater outflow at low tide. This
reduces mangrove biodiversity, with inner mangrove zone species of riverine
habitats such as Nypa fruticans absent, and favours full salinity tolerant species
such as Rhizophora stylosa in the Pacific and Rhizophora mangle in the Caribbean.
With the rising sea level these systems can accumulate autochthonous sediment
deterministically (Ellison 2019), sustaining low gradients across the mangroves to
allow extensive areas even in microtidal ranges.

4.5 Tidal Range and Inundation Regimes

Tidal range varies across the mangrove world from microtidal in the Caribbean to
macrotidal in locations such as North West Australia and eastern Africa, and for
mangroves tidal range is a limiting factor in spatial extent (Ellison 2015). Accom-
modation space for the mangrove habitat is only from mean sea level to high tide
levels, which acting against the intertidal gradient limits the spatial expansion of
mangrove areas. This does not however mean that microtidal areas have narrow
mangroves, as the geomorphic setting is a moderator (Table 4.1). Deterministic
mangroves can maintain their own habitat in the tidal range by net accretion during
slow relative sea level rise (Ellison 2019), and with substrate surfaces controlled by
root mat development can achieve very low gradient surfaces in microtidal areas to
allow expansive mangroves.
Micro-topography influences the distribution of mangroves, controlling the criti-
cal periods of tidal inundation and air exposure that governs the health of the forest
(Watson 1928; Kjerfve 1990, Friess 2017). Flooding depths, durations and
frequencies are critical factors in the survival of each mangrove species (Lewis
et al. 2019), resulting in zones of different species, each with preferred habitats
(Fig. 4.5). In the first research on this subject, Watson derived inundation classes for
Port Klang, Malaysia, recording mangroves across a 2.13 m range (Friess 2017)
within a 5.5 m spring tidal range (Port Klang Malaysia Marine Information Hand-
book 2010). This research commenced a mangrove paradigm relating ground surface
elevation to frequency of flooding that result in species zonation (Friess 2017;
Ellison 2019).
4 Factors Influencing Mangrove Ecosystems 105

Fig. 4.5 Typical zonation by elevation of mangroves, with graph showing amount of time each
zone has the ground surface in the air (in yellow) and under water (in blue)

Mangrove species are adapted to tidal inundation with aerial roots architectures;
these include pneumatophores in Avicennia and Sonneratia, stilt roots in
Rhizophora, and knee roots in Bruguiera and Lumnitzera species (Tomlinson
2016). These are illustrated in Fig. 4.5 from right to left. Each species has specific
preferences for timeframes and depths of inundation as controlled by the level of
tidal waters relative to the roots and their substrate, which results in species zonation
by micro-elevation. Figure 4.5 adapts a standard tidal curve from Maritime Safety
Queensland (2019: 118) to demonstrate in the base graph the proportion of a tidal
cycle that mangroves at different elevation zones have their substrate in the air and
underwater. This is 50% at the seaward edge at mean sea level, and increased time in
the air towards the high tide margins, the gradient controlling oxygen availability
and salinity.
Mangroves can experience decline and mortality when their inundation patterns
are altered, and this is part of their vulnerability to rising sea levels (Ellison 2015).
Inundation changes caused by direct human influences that alter hydrological
exchange show clear evidence of this sensitivity. These are well documented from
the decades before mangroves became better appreciated and valued, as they have
become this century.
Hydrological alteration is a major cause of loss of mangroves in the Indo-West
Pacific (Saenger et al. 2019), indicating the importance of the inundation regime as a
key factor for mangrove vitality. Evidence is also well recorded from American
longitudes, with a mortality of 15 hectares of Avicennia germinans and
Laguncularia racemosa occurred in Puerto Rico when dredging caused an abrupt
change to the hydrological regime, resulting in permanent flooding with cause of
death attributed to reduced soil oxygen (Jimenez et al. 1985). Mortality of Avicennia
germinans at Clam Bay, Florida occurred following blockage of tidal channels to
cause excess inundation by accumulated flood water (Turner and Lewis 1997).
Blockage of tidal creeks in NW Australia caused death of 6 hectares of Avicennia
marina and Ceriops tagal (Gordon 1988). Construction of an evaporator pond
caused permanent ponding and mass mortality to about 11 km2 of mangroves
106 J. C. Ellison

(Gordon 1988). In Florida, impoundment of mangroves for mosquito control


provided several early case studies of prolonged inundation causing mangrove
mortality (Brockmeyer et al. 1997). Avicennia germinans trees are killed if their
short pneumatophores are submerged after just weeks. The inundation was of sudden
onset, and then sustained to cause mortality of Rhizophora mangle and Avicennia
germinans at Indian River, East Florida as a result of 4 months flooding of
0.3–0.45 m depth (Harrington and Harrington 1982).
Following closure of the river mouth, water levels in the Kosi Estuary, Natal rose
by about 0.3 m over the succeeding 140 days (Forbes and Cyrus 1992). Mass
mortality to mangroves occurred, primarily Avicennia marina, with the bark of all
trees rotted around 0.2 m above the mud level. Mangroves on higher elevations,
mostly Lumnitzera racemosa survived.
Human-caused changes to hydrological exchange in Columbia resulted in hyper-
salinisation of soils, resulting in mortality of 30,000 hectares of mangroves (Elster
et al. 1999). At Clarence Estuary in NSW, structural flood mitigation works caused
loss of tidal fluctuations in mangrove areas (Pollard and Hannan 1994). This caused
near total loss of mangroves above the flood gates, from reduced tidal inundation.
Natural causes of sustained inundation have also been recorded to have caused
mangrove mortality, further indicating the importance of this factor to system
tolerance. River flooding causing inundation and sediment deposition on Bruguiera
caused 42 hectares of mortality of in Java (Soerianegara 1968). Inundation of 2.5 m
above the mean summer level falling slowly to 1.7–1.5 m two weeks later, and
returning to typical levels by 2 months later caused mortalities to Bruguiera
gymnorrhiza (Forbes and Cyrus 1992).
Flooding after enclosure by a natural sand bar of 50 hectares of Avicennia
mangroves in Brunei-Darussalam, Borneo, brought water depths of 0.5 m above
tidal high water for 8 weeks before the floodwater was released (Choy and Booth
1994). Avicennia marina suffered substantial losses of shrub-sized plants, while
most trees >5 m survived. Necrosis (death of tissue) occurred to bark and surface
wood tissues on mangroves stems below flood level, and death of pneumatophores
of trees occurred.
Experimental flooding resulted in the lower ability of leaves to conduct water, and
increase in stomatal closing in Bruguiera gymnorrhiza, leading to reduced rates of
photosynthesis (Naidoo 1983). When lenticels of aerial roots become inundated,
oxygen concentrations in the plant fall dramatically (Scholander et al. 1955). If
inundation is sustained, low-oxygen conditions and mortalities follow.
Mangrove seedlings are thought to be more susceptible to death from excess inun-
dation than adult trees. During sustained flooding of about 30 cm, all seedlings of
Avicennia germinans and Laguncularia racemosa died within 1 month (Elster et al.
1999). Inundation effects on mangrove seedling growth have been demonstrated
experimentally (Ellison and Farnsworth 1997), showing that normal tidal fluctuation
around a 16 cm raised water level slowed growth at the sapling stage, and by
2.5 years age plants were 10–20% smaller than control plants.
4 Factors Influencing Mangrove Ecosystems 107

Hence, while mangroves are tolerant of tidal inundation, inundation above


normal for prolonged periods brings mangrove tree mortality, and decline is
observed during less extreme flood events. This is due to reduced capability of gas
exchange in the root systems (Jimenez et al. 1985). These case studies indicate the
importance of the inundation factor (Fig. 4.5) in limiting mangroves, bringing a
tipping point to mangrove system equilibrium.

4.6 Wave Action

Exposure to wave action is a limit to mangrove distributions (Duke 2017), and on


wave-dominated coastlines mangroves are limited to sheltered lagoons inside
barriers (Table 4.1, Fig. 4.4C). Wave energy is higher outside of the equatorial
latitudes for mangrove distributions, as prevalent winds are stronger. Swell waves
are generated by the frictional effect of wind as it passes over the ocean surface,
where swell consists of long low waves generated in a major wave generation area
due to the action of prevailing/storm winds on the water. Wave dispersion causes
regularly spaced successions of waves to migrate from the source zone. As waves
move towards land, wave velocity decreases and wave height increases, and offshore
gradient moderates wave energy in approach to shore. Tsunami waves are a rare
series of large waves of extremely long wavelength generated by undersea
earthquakes or disturbance near the coast or in the ocean. Wave energy limits
mangrove distributions, and large waves can cause damage and mortality. Friction
of intertidal root systems and intertidal mud banks trapped under mangroves forces
storm waves to spill not plunge in break. Energy is dissipated as the waves traverse
the mangrove fringe.
An important value of mangroves to human communities is the protection the
forest provides to land from waves and tidal surges, during cyclones and other
storms. The dense foliage of mangroves combined with friction effects of aerial
roots provides some facility in reducing wave power, including reducing damage
from extreme events such as tsunami waves (Mazda et al. 2007). At high tide in a
Rhizophora forest, there is a 50% decline in wave energy within 150 m of the
seaward edge (Brinkman et al. 2007). The degree of protection obviously depends
on the size of the waves, and the maturity and density of the forest (Yanagisawa et al.
2009).
Following the 2004 Asian tsunami it was found that, in some places, human
deaths and loss of property were reduced by the presence of coastal vegetation
shielding coastal villages (Dahdouh-Guebas et al. 2005; Katharesan and Rajendran
2005). However, a tsunami wave of over 6 m resulted in mangroves being mostly
destroyed (Yanagisawa et al. 2009), with total destruction occurring from larger
tsunamis (Koh et al. 2018).
Evidence of damage to mangroves from high energy events is available from
many accounts over the last several decades, such as Cyclone Nargis impacting
Myanmar in 2008, with winds of up to 215 km/h in the area of mangroves for about
12 hours, and most mangroves were damaged (Aung et al. 2011). Most defoliation
108 J. C. Ellison

and uprooting occurred to taller plantations, and Avicennia marina was found to be
the most resistant to defoliation. After Hurricane Mitch hit Honduras in 1998, after
2 days of winds of around 200 km/h, nearly complete defoliation of mangroves
occurred, and taller trees uprooted or broken. Of about 311 hectares of mangrove
forests in Guanaja, only 11 hectares (3%) survived (Cahoon et al. 2003), leading to
rapid sediment elevation loss as peat collapsed.
In moderate events, mangroves have an important role in protecting coasts during
storm and tsunami events, primarily by reduction of the energy of waves (Massel
et al. 1999; Dahdouh-Guebas et al. 2005; Katharesan and Rajendran 2005;
Danielson et al. 2005; Vermaat and Thampanya 2006). During a typhoon wave
height can be reduced by 20% on a mangrove shore of 100 m width (Mazda et al.
1997), and a similar tract of mangroves could reduce tsunami wave height by 50%
(Hirashi and Harada 2003). The survival rate of mangroves to tsunami impacts is
greater where there is a higher stem diameter (Yanagisawa et al. 2009).
Where waves are combined with wind action such as in severe storms, damage to
mangroves has been shown to be severe in many case studies. Mangroves prefer
habitats of low energy shorelines, sheltered from strong wave action. Mangroves can
be impacted or damaged by high energy events such as storms or tsunamis. Consis-
tent wave energy is a factor influencing the natural occurrence of mangroves on
shorelines.

4.7 Tipping Points

Mangroves are complex systems influenced by physical factors of temperature, frost,


wave power, tides, currents, sea level, sediment supply and geomorphic setting.
These control attributes such as biodiversity, net productivity and sedimentation
rates (Fig. 4.2). Such resilient systems are in dynamic equilibrium as shown by (1) in
Fig. 4.6, where a small horizontal mark is a period of steady time (Ritter et al. 2002),
such as represented by observations a scientist or student may make on visiting a
mangrove area. Over time within that equilibrium state, ups and downs occur related
to perturbations or El Niño cycles, these not so great as to break out of the long-term
equilibrium, the system recovers. For example, storms physically impact mangroves,
and recovery can then occur by local recruitment. Large waves cause erosion, and
deposition and mangrove extent adjust. Pests or human cutting may have some
impact, but the system recovers robustness and younger trees grow up in light gaps
caused. The equilibrium steady state is represented in Fig. 4.6 by the regular up and
down waves, such as at (1).
Tipping points have conceptual roots in the palaeoecological literature, with
major geological time periods categorised by extinction events caused by major
perturbations (Alvarez et al. 1984). After mangroves first appeared in the Late
Palaeocene in the Asian/Pacific region, during warmer climates than today (Saenger

1998), they extended up to 50 north and south latitude (Ellison 2008). As
temperatures fell from the end of the Eocene to the Pliocene, mangrove latitudinal
4 Factors Influencing Mangrove Ecosystems 109

Fig. 4.6 Conceptual graph showing system equilibrium states of mangrove ecosystems over time,
and tipping points leading to change and adjustment

ranges reduced, each location reaching a tipping point followed by the loss from the
record. Tipping points are also evident in relation to sea level rise, of large mangrove
expanses in the mid-Holocene that were replaced by offshore communities such as
seagrass or lagoons (Ellison 2008, 2019). The actual causes of these changes by
factors of salinity, inundation during mid-Holocene events, or temperature fall or
frost in the Late Tertiary can only be inferred from stratigraphic records. They can
however be understood through concurrent ecological system studies (Otero et al.
2016; Cavanaugh et al. 2019; Osland et al. 2020).
The interplay of growth factors promoted by beneficial factors balanced against
stress or reduction factors regulates population size and hence ecosystem balance.
Stable or balanced ecosystems have become a concept behind successfully managed
ecosystems in a sustainable manner. Physical factors change to bring habitat stress or
disturbance, to which the ecosystem may respond with resilience recovery or
progress towards a tipping point of change to an altered system. A tipping point is
any influence that causes a threshold change (Brunsden and Thornes 1979; van
Belzen et al. 2017), such that regulatory processes that maintained the system
equilibrium are so sufficiently altered that it cannot be recovered. For ecosystems
a phenomenon of critical slowing down has been recognised (Scheffer et al.
2009; Van Belzen et al. 2017; Eslami-Andergoli et al. 2015) which can give early
warning of an imminent tipping point. It is a state of low resilience, which for
mangroves could be measured within dimensions of ecosystem robustness, magni-
tude of stressors and effectiveness of management actions (Ong and Ellison 2021).
Stressors that may cause critical reduction of mangrove resilience are impacts
from humans, climate becoming significantly drier, increased inundation, reduced
net sedimentation and relative sea level rise. The combination of low sediment
supply and relative sea level rise indicates how factors may apply in combination
to cause a tipping point that the system resilience may endure if only one were
110 J. C. Ellison

applying. A notable factor in mangrove recovery from oil spill impacts is the prior
condition and resilience of the system before impact (Duke 2016). Ecosystem
robustness indicators of critical reduction are mangrove condition and mortality,
spatial loss, low net accretion or reduced adjacent ecosystem resilience such as the
co-benefits of adjacent ecosystems such as coral reefs or seagrass are reduced.
Stressors and reduced robustness apply in parallel, such as a critical threshold for
future mangrove vulnerability is net mangrove accretion rates relative to rates of
relative sea level rise (Ellison 2019). Management actions that may lead to a critical
reduction of mangrove resilience are poor mangrove protection legislation or lack of
enforcement, local community management in need of capacity building, and poor
stakeholder involvement in mangrove conservation. These indicators of mangrove
resilience can each be monitored and evaluated (Ong and Ellison 2021), to help
potentially avoid a tipping point for the mangrove system.
A tipping point (Fig. 4.6) is a sudden change in ecosystem function, a breakdown
of equilibrium. Change or adjustment after a tipping point is schematically shown as
a line in Fig. 4.6, but could be a crash or jagged decline, with loss of the previous
stable equilibrium. It is followed by a time of change/adjustment which may reach a
new stable equilibrium. After impoundment, mangrove areas suffered mortality to
become aquatic lagoons (Gordon 1988; Brockmeyer et al. 1997). Sea level rise
caused conversion of mangrove areas to shallow bays (Ellison and Zouh 2012;
Kemp et al. 2019). Mangrove areas are frequently cleared and transformed into
aquaculture ponds (Spalding et al. 2010; Oh et al. 2017), such as to allow shrimp
farming in Indonesia (Koh et al. 2018). Water exclusion such as by road construction
causes mangrove mortality owing to hyper-salinisation (Elster et al. 1999), and
subsequently mangroves have reduced, ecosystems have altered (de Klerk 2016;
Gómez et al. 2017). These alternative systems become the new equilibrium for the
location, as shown by (2) in Fig. 4.6. From here, a tipping point may further change
the system to a different equilibrium, which may be rehabilitation back to a replanted
mangrove area, these shown by (3) in Fig. 4.6. Rehabilitation can be successful if
ecological guidance on mangrove restoration is followed (Lewis and Brown 2014;
Lewis et al. 2019), particularly topographic positioning with respect to tidal inunda-
tion frequency factors (Ellison 2020).

4.8 Conclusions

Mangroves are sensitive to even minor changes in influencing factors (Fig. 4.1), such
as altered drainage patterns, saltwater intrusion, accretion or erosion in response to
changes in sea level (Ellison 2009). Each factor influences mangroves simulta-
neously (Fig. 4.2), though with varying intensity relative to coastal setting
(Table 4.1), worldwide location (Fig. 4.3) and over time such as with climate change.
The response of mangroves to these factors is observed through variations in the
composition and relative abundance of plant species within the mangrove habitat
(Blasco et al. 1996; Ellison 2005). Although the responses may be gradual, particu-
larly in undisturbed systems, the alterations in coverage and composition of species
4 Factors Influencing Mangrove Ecosystems 111

can be used to assess the effects of climate change and other environmental impacts
on mangrove habitats. This can be demonstrated through palaeo-environmental
reconstruction (Ellison 2019), spatial monitoring (Bunting et al. 2018) or mangrove
ecosystem monitoring (Ellison et al. 2012).
Understanding of the physical factors controlling mangrove ecosystems is essen-
tial to effective management. Factors of temperature, ocean currents and land
barriers influence the spatial range of species, and rainfall can be a limiting factor
in drier areas. Tidal ranges control vertical mangrove extents, across intertidal
gradients and hydroperiod controls the locations of different mangrove zones.
Coastal typology influences sediment supply, and the relative balance between
wave, tide and river processes. Change in factors can bring a tipping point of loss
of ecosystem equilibrium, followed by a change and adjustment to a different system
which may be undesired to management goals. The rehabilitation investment may be
able to return mangrove ecosystem balance, if the factors for mangrove successful
growth are managed.

References
Ajonina GN, Kairo J, Grimsditch G, Sembres T, Chuyong G, Diyouke E (2014) Assessment of
mangrove carbon stocks in Cameroon, Gabon, the Republic of Congo (RoC) and the Demo-
cratic Republic of Congo (DRC) including their potential for reducing emissions from defores-
tation and forest degradation (REDD+). In: Diop S, Barusseau J-P, Descamps C (eds) The land/
ocean interactions in the coastal zone of West and Central Africa, Springer, Cham, p. 177–189.
Alvarez W, Alvarez LW, Asaro F, Michel HV (1984) The end of the Cretaceous: sharp boundary or
gradual transition? Science 223(4641):1183–1186.
Ashbridge EF, Bartolo R, Finlayson CM, Lucas RM, Rogers K, Woodroffe CD (2019) Assessing
the distribution and drivers of mangrove dieback in Kakadu National Park, northern Australia.
Estuar Coast Shelf Sci 228. https://doi.org/10.1016/j.ecss.2019.106353
Aung TY, Than MM, Katsuhiro O, Yukira M (2011) Assessing the status of three mangrove species
restored by the local community in the cyclone-affected area of the Ayeyarwady Delta,
Myanmar. Wetl Ecol Manage 19(2):195–208.
Blasco F, Saenger P, Janodet E (1996) Mangroves as indicators of coastal change. Catena
27:167–178.
Brinkman RM, Massel SR, Ridd PV, Furukawa K (2007) Surface wave attenuation in mangrove
forests. Proc 13th Austral Coastal Ocean Engineer Conf 2:941–949.
Brockmeyer RE Jr, Rey JR, Virnstein RW, Gilmore RG, Ernest L (1997) Rehabilitation of
impounded estuarine wetlands by hydrologic reconnection to the Indian River Lagoon, Florida
(USA). Wetl Ecol Manage 4:93–109.
Brunsden D, Thornes JB (1979) Landscape sensitivity and change. Trans Inst Brit Geogr 4
(4):463–484.
Bunting P, Rosenqvist A, Lucas RM, Rebelo LM, Hilarides L, Thomas N, Hardy A, Itoh T,
Shimada M, Finlayson CM (2018) The global mangrove watch-a new 2010 global baseline of
mangrove extent. Remote Sens 10:1669. https://doi.org/10.3390/rs10101669
Cavanaugh KC, Osland MJ, Bardou R, Hinojosa-Arango G, López-Vivas JM, Parker JD, Rovai AS
(2018) Sensitivity of mangrove range limits to climate variability. Global Ecol Biogeogr 27
(8):925–935.
Cavanaugh KC, Dangremond EM, Doughty CL, Williams AP, Parker JD, Hayes MA,
Rodriguez W, Feller IC (2019) Climate-driven regime shifts in a mangrove–salt marsh ecotone
over the past 250 years. Proc Nat Acad Sci 116(43):21602–21608.
112 J. C. Ellison

Cahoon DR, Hensel P, Rybczyk J, McKee KL, Proffitt CE, Perez BC (2003) Mass tree mortality
leads to mangrove peat collapse at Bay Islands, Honduras after Hurricane Mitch. J Ecol 91
(6):1093–1105.
Choy SC, Booth WE (1994) Prolonged inundation and ecological changes in an Avicennia
mangrove: implications for conservation and management. Hydrobiol 285(1):237–247.
Dahdouh-Guebas F, Jayatissa LP, Di Nitto D, Bosire JO, Lo Seen D, Koedam N (2005) How
effective were mangroves as a defence against the recent tsunami? Curr Biol 15:R443–447.
Danielson F, Soerensen M, Olwig M, Selvam V, Parish F, Burgess N, Hiraishi T, Karunagaran V,
Rasmussen M, Hansen L, Quarto A, Nyoman S (2005) The Asian tsunami: A protective role for
coastal vegetation. Science 310:643.
de Klerk J (2016) The slow death of the Ciénaga Grande de Santa Marta. The City Paper (Bogata)
22 September 2016. https://thecitypaperbogota.com/features/the-slow-death-of-the-cienaga-
grande-de-santa-marta/14235.
de Lange WP, De Lange PJ (1994) An appraisal of factors controlling the latitudinal distribution of
mangrove (Avicennia marina var. resinifera) in New Zealand. J Coast Res 10(3):539–548.
Duke NC (2020) Mangrove harbingers of coastal degradation seen in their responses to global
climate change coupled with ever-increasing human pressures. CHEC Hum Ecol 30:19–23.
Duke NC (2016) Oil spill impacts on mangroves: recommendations for operational planning and
action based on a global review. Mar Pollut Bull 109(2):700–715.
Duke NC (2017) Mangrove floristics and biogeography revisited: further deductions from biodi-
versity hot spots, ancestral discontinuities, and common evolutionary processes. In: Rivera-
Monroy VH, Lee SY, Kristensen E, Twilley RR (eds) Mangrove ecosystems: A global
biogeographic perspective, Springer, Cham, p. 17–53.
Duke NC, Kovacs JM, Griffiths AD, Preece L, Hill DJ, Van Oosterzee P, Mackenzie J, Morning
HS, Burrows D (2017) Large-scale dieback of mangroves in Australia’s Gulf of Carpentaria: a
severe ecosystem response, coincidental with an unusually extreme weather event. Marine
Freshw Res 68(10):1816–1829.
Duke NC, Ball MC, Ellison JC (1998) Factors influencing biodiversity and distributional gradients
in mangroves. Global Ecol Biogeog Let 7:27–47
Ellison AM, Farnsworth EJ (1997) Simulated sea level change alters anatomy, physiology, growth,
and reproduction of red mangrove (Rhizophora mangle L.). Oecologia 112(4):435–446.
Ellison AM, Farnsworth EJ, Merkt RE (1999) Origins of mangrove ecosystems and the mangrove
biodiversity anomaly. Global Ecol. Biogeogr. 8, 95–115.
Ellison JC (2020) Mangrove ecosystem-based adaptation: Advice on improved success. CHEC
Hum Ecol 30:37–40.
Ellison JC (2019) Biogeomorphology of mangroves. In: Wolanski E, Cahoon D, Perillo GME (eds)
Coastal wetlands: An ecosystem integrated approach, Elsevier Science, Amsterdam,
p. 687–715.
Ellison JC (2005) Holocene palynology and sea-level change in two estuaries in Southern Irian
Jaya. Palaeogeogr Palaeoclimatol Palaeoecol 220:291–309.
Ellison JC (2009) Geomorphology and sedimentology of mangrove swamps. In: Wolanski E,
Cahoon D, Perillo GME (eds) Coastal wetlands: an ecosystem integrated approach. Elsevier
Science, Amsterdam, p 564–591.
Ellison JC (2015) Vulnerability assessment of mangroves to climate change and sea-level rise
impacts. Wetl Ecol Manage 23:115–137.
Ellison JC (2008) Long-term retrospection on mangrove development using sediment cores and
pollen analysis. Aquat Bot 89:93–104.
Ellison JC (1997). Mangrove community characteristics and litter production in Bermuda. In:
Kjerfve B, Lacerda LD, Diop ES (eds) Mangrove forests of the Latin America and Africa
regions, UNESCO, Paris, p. 5–17.
Ellison JC, Jungblut V, Anderson P, Slaven C (2012) Manual for mangrove monitoring in the
Pacific Islands Region. Secretariat of the Pacific Regional Environment Program, Apia, Samoa.
4 Factors Influencing Mangrove Ecosystems 113

Ellison JC, Zouh I (2012) Vulnerability to climate change of mangroves: Assessment from
Cameroon, Central Africa. Biology 1:617–638.
Elster C, Perdomo L, Schnetter M-L (1999) Impact of ecological factors on the regeneration of
mangroves in the Cienaga Grande de Santa Marta, Colombia. Hydrobiol 413:35–46.
Eslami-Andergoli L, Dale PE, Knight JM, McCallum H. (2015) Approaching tipping points: a
focussed review of indicators and relevance to managing intertidal ecosystems. Wetl Ecol
Manage 23(5):791–802.
Forbes AT, Cyrus DP (1992) Impact of a major cyclone on southeast African estuarine lake system.
Neth J Sea Res 30:265–272.
Friess DA (2017) JG Watson, inundation classes, and their influence on paradigms in mangrove
forest ecology. Wetl 37(4):603–613.
Gómez JF, Byrne ML, Hamilton J, Isla F (2017) Historical coastal evolution and dune vegetation in
isla salamanca national park, colombia. J Coast Res 33(3):632–641.
Google Earth Pro (2020) Imagery attributed to CNES/Airbus, Terra Metrics, and Maxar
Technologies, used following attribution guidelines https://www.google.com/permissions/
geoguidelines/attr-guide/
Gordon DM (1988) Disturbance to mangroves in tropical-arid Western Australia: hypersalinity and
restricted tidal exchange as factors leading to mortality. J Arid Environ 15:117–145.
Harrington, RW, Harrington ES (1982) Effects on fishes and their forage organisms of impounding
a Florida salt marsh to prevent breeding by salt marsh mosquitos. Bull Mar Sci 32:523–531.
Hirashi T, Harada K (2003) Greenbelt tsunami prevention in South-Pacific region. Rep Port Airport
Res Inst 42:3–25.
Horstman EM, Lundquist CJ, Bryan KR, Bulmer RH, Mullarney JC, Stokes DJ (2018) The
dynamics of expanding mangroves in New Zealand. In: Makowski C, Finkl CW (eds) Threats
to mangrove forests, Springer, Cham, p. 23–51.
Hutchings P, Saenger P (1987) Ecology of Mangroves. University of Queensland Press, St Lucia.
Jimenez JA, Lugo AE, Cintron G (1985) Tree mortality in mangrove forests. Biotrop 17:177–185.
Katharesan K, Rajendran N (2005) Coastal mangrove forests mitigates tsunami. Estuar Coast Shelf
Sci 65:601–606.
Kemp AC, Vane CH, Khan NS, Ellison JC, Engelhart SE, Horton BP, Nikitina D, Smith SR,
Rodrigues LJ, Moyer RP (2019) Testing the utility of geochemical proxies to reconstruct
holocene coastal environments and relative sea level: a case study from Hungry Bay, Bermuda.
Open Quat 5(5):1–8.
Kjerfve B (1990) Manual for investigation of hydrological processes in mangrove ecosystems.
Baruch Institute for Marine Biology and Coastal Research, University of South Carolina.
Koh H, Teh SY, Kh’ng XY, Raja Barizan RS (2018) Mangrove forests: protection against and
resilience to coastal disturbances. J Trop For Sci 30(5):446–460.
Lewis RR, Brown BM, Flynn LL (2019) Methods and criteria for successful mangrove forest
rehabilitation. In: Wolanski E, Cahoon D, Perillo GME (eds) Coastal wetlands: An ecosystem
integrated approach, Elsevier Science, Amsterdam, p 863–887.
Lewis RR, Brown B (2014) Ecological mangrove rehabilitation-a field manual for practitioners.
Mangrove Action Project Indonesia, Blue Forests, Canadian International Development Agency
and OXFAM, Gatineau, Quebec, Canada. http://mangroverestoration.com/.
Lincoln R, Boxshall G, Clark P (1982) Dictionary of ecology, evolution and systematics.
Cambridge University Press, Cambridge.
Lovelock CE, Feller IC, Reef R, Hickey S, Ball MC (2017) Mangrove dieback during fluctuating
sea levels. Sci Rep 7(1):1–8.
Lugo AE, Snedaker SC (1974) The ecology of mangroves. Ann Rev Ecol Syst 5(1):39–64.
Macmillan dictionary (2020) https://www.macmillandictionary.com/thesaurus-category/british/
biological-processes.
Mahoney PC, Bishop MJ (2018) Are geomorphological typologies for estuaries also useful for
classifying their ecosystems? Aquat Conserv 28(5):1200–1208.
114 J. C. Ellison

Maritime Safety Queensland (2019) Queensland tide tables standard port tide times 2020. Depart-
ment of Transport and Main Roads, State of Queensland.
Mazda Y, Magi M, Kogo M, Hong PN (1997) Mangroves as a coastal protection from waves in the
Tong King delta, Vietnam. Mangr Salt Marsh 1(2):127–135.
Mazda Y, Parish F, Danielsen F, Imamura F (2007) Hydraulic functions of mangroves in relation to
tsunamis. Mangr Sci 4(5):57–67.
McSweeney SL, Kennedy DM, Rutherfurd ID, Stout JC (2017) Intermittently closed/open lakes
and lagoons: Their global distribution and boundary conditions. Geomorph 292:142–152.
Naidoo G (1983) Effects of flooding on leaf water potential and stomatal resistance in Bruguiera
gymnorrhiza. New Phytol 93:369–373.
Odum EP (1972) Ecosystem theory in relation to man. In: Wiens JA (ed) Ecosystem structure and
function, Oregon State University Press, Corvallis, p. 11–24.
Oh RRY, Friess DA, Brown BM (2017) The role of surface elevation in the rehabilitation of
abandoned aquaculture ponds to mangrove forests, Sulawesi, Indonesia. Ecol Engineer
100:325–334.
Ong WJ, Ellison JC (2021) A framework for the quantitative assessment of mangrove resilience. In:
Friess D, Sidik F (eds) Dynamic sedimentary environments of mangrove coasts, Elsevier,
Amsterdam, p. 513–538.
Osland MJ, Day RH, Hall CT, Brumfield MD, Dugas JL, Jones WR (2017) Mangrove expansion
and contraction at a poleward range limit: climate extremes and land-ocean temperature
gradients. Ecology 98:125–137.
Osland MJ, Day RH, Michot TC (2020) Frequency of extreme freeze events controls the distribu-
tion and structure of black mangroves (Avicennia germinans) near their northern range limit in
coastal Louisiana. Divers Distrib 00:1–17. https://doi.org/10.1111/ddi.13119
Otero V, Quisthoudt K, Koedam N, Dahdouh-Guebas F (2016) Mangroves at their limits: Detection
and area estimation of mangroves along the Sahara Desert Coast. Remote Sens 8:512. https://
doi.org/10.3390/rs8060512
Polidoro BA, Carpenter KE, Collins L, Duke NC, Ellison AM, Ellison JC, Farnsworth EJ, Fernando
ES, Kathiresan K, Koedam NE, Livingstone SR, Miyagi T, Moore GE, Nam VN, Primavera JH,
Salmo III SG, Sanciangco JC, Sukardjo S, Wang Y, Yong JWH (2010) The loss of species:
Mangrove extinction risk and geographic areas of global concern. PLoS One 5(4): e10095.
https://doi.org/10.1371/journal.pone.0010095
Pollard DA, Hannan JC (1994) The ecological effects of structural flood mitigation works on fish
habitats and fish communities in the lower Clarence River system of south-eastern Australia.
Estuar 17:427–461.
Port Klang Malaysia Marine Information Handbook (2010) https://www.pka.gov.my/
phocadownloadpap/information/marine/MARINE%20HANDBOOK.pdf.
Putman R, Warren SD (1984) Principles of ecology. Chapman and Hall, London.
Ritter DF, Kochel C, Miller JR (2002) Process geomorphology. McGraw Hill, Dubuque.
Saenger P (1998) Mangrove vegetation: an evolutionary perspective. Mar Freshw Res 49:277–286.
Saenger P, Ragavan P, Sheue CR, López-Portillo J, Yong JWH, Mageswaran, T (2019) Mangrove
biogeography of the Indo-Pacific. In: Gul B, Böer B, Khan M, Clüsener-Godt M, Hameed A
(eds) Sabkha ecosystems, Springer, Cham, p. 379–400.
Sasmito SD, Kuzyakov Y, Lubis AA, Murdiyarso D, Hutley LB, Bachri S, Friess DA, Martius C,
Borchard N (2020) Organic carbon burial and sources in soils of coastal mudflat and mangrove
ecosystems. Catena 187:104414.
Scheffer M, Bascompte J, Brock WA, Brovkin V, Carpenter SR, Dakos V, Held H, Van Nes EH,
Rietkerk M, Sugihara G (2009) Early-warning signals for critical transitions. Nature 461
(7260):53–59.
Scholander PF, van Dam L, Scholander SI (1955) Gas exchange in roots of mangroves. Amer J Bot
42(1):92–98.
Soerianegara I (1968) The causes of mortality of Bruguiera trees in the mangrove forest near
Tjilatjap, Central Java. Rimba Ind 13:1–11.
4 Factors Influencing Mangrove Ecosystems 115

Spalding M, Kainuma M, Collins L (2010) World Atlas of Mangroves. Earthscan, London and
Washington, DC.
Srivastava J, Prasad V (2019) Evolution and paleobiogeography of mangroves. Mar Ecol 40(6):
e12571.
Steinke TD, Naidoo Y (1991) Respiration and net photosynthesis of cotyledons during establish-
ment and early growth of propagules of the mangrove, Avicennia marina, at three temperatures.
S Afr J Bot 57:171–174.
Tedford, M, Ellison, JC (2018) Analysis of river rehabilitation success, Pipers River, Tasmania.
Ecol Indic 91:350–358.
Tomlinson PB (2016) The botany of mangroves. Cambridge University Press, Cambridge.
Turner RE, Lewis RR (1997) Hydrologic restoration of coastal wetlands. Wetl Ecol Manage
4:65–72.
Valiela I, Pascual J, Giblin A, Barth-Jensen C, Martinetto P, Otter M, Stone T, Tucker J,
Bartholomew M, Viana IG (2018) External and local controls on land-sea coupling assessed
by stable isotopic signatures of mangrove producers in estuaries of Pacific Panama. Mar Env
Res 137:133–144.
van Belzen J, van de Koppel J, Kirwan M, van der Wal D, Herman PMJ, Dakos V, Kefi S,
Scheffer M, Guntenspergen GR, Bouma TJ (2017) Vegetation recovery in tidal marshes reveals
critical slowing down under increased inundation. Nat Commun 8:15811. https://doi.org/10.
1038/ncomms15811.
Van der Stocken T, Wee AK, De Ryck DJ, Vanschoenwinkel B, Friess DA, Dahdouh-Guebas F,
Simard M, Koedam N, Webb EL (2019) A general framework for propagule dispersal in
mangroves. Biol Rev 94(4):1547–1575.
Vermaat JE, Thampanya U (2006) Mangroves mitigate tsunami damage: A further response. Estuar
Coast Shelf Sci 69:1–3.
Watson JG (1928) Mangrove forests of the Malay Peninsula. Malay For Rec 6:1–275.
Waycott M, McKenzie LJ, Mellors JE, Ellison JC, Sheaves MT, Collier C, Schwarz AM, Webb A,
Johnson JE, Payri CE (2011) Vulnerability of mangroves, seagrasses and intertidal flats in the
tropical Pacific to climate change. In: Bell JD, Johnson JE, Hobday AJ (eds) Vulnerability of
fisheries and aquaculture in the Pacific to climate change, Secretariat of the Pacific Community,
Noumea, p. 297–368.
Yan YB, Duke NC, Sun M (2016) Comparative analysis of the pattern of population genetic
diversity in three Indo-west Pacific Rhizophora mangrove species. Front Plant Sci 30(7):1434.
Yanagisawa H, Koshimura S, Goto K, Miyagi T, Imamura F, Ruangrassamee A, Tanavud C (2009)
The reduction effects of mangrove forest on a tsunami based on field surveys at Pakarang Cape,
Thailand and numerical analysis. Estuar Coast Shelf Sci 81:27–37.
Energy Flux in Mangrove Ecosystems
5
Engku Azlin Rahayu Engku Ariff,
Ahmad Faris Seman Kamarulzaman, and Mohd Nazip Suratman

Abstract

The energy balance at the Earth’s land surface requires that the energy gained
from net radiation be balanced by the fluxes of sensible and latent heat to the
atmosphere and the storage of heat in the soil. Latent and sensible heat are crucial
variables in ecology, hydrology and meteorology because they give influence to a
climate that can be used to determine environmental parameters which alter mass
and energy exchange between the soil and the atmosphere. These energy fluxes
are a primary determinant of surface climate. The annual energy balance at the
land surface differs geographically depending on the incoming solar radiation and
soil water availability. Thus, provides key insight into processes such as photo-
synthesis and respiration. Throughout the days and years, energy flux varies
depending on the diurnal and annual cycles of solar radiation and also soil
water availability. The various terms in the energy budget (net radiation, sensible
heat flux, latent heat flux and soil heat flux) are illustrated for different climate
zones and for various vegetation types. In this review, the energy flux in man-
grove ecosystem in the trophic level is highlighted. We propose that integrating
view point from community and ecosystem ecology in mangrove by quantifying
energy fluxes from solar energy to producers, consumers and decomposers which
can provide vital information for understanding the connections between the

E. A. R. Engku Ariff · A. F. Seman Kamarulzaman


Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM) Pahang Campus, Bandar Tun
Razak Jengka, Malaysia
M. N. Suratman (*)
Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 117
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_5
118 E. A. R. Engku Ariff et al.

diversity of complex multitrophic systems as well as multiple ecosystem


functions in mangrove ecosystem.

Keywords

Mangroves · Energy flux · Photosynthesis · Respiration · Consumer · Decomposer

5.1 Introduction

Mangrove is referring to the trees, large shrubs, ferns and palm that grow in the
intertidal coastal zone of the tropics and subtropics regions of the world between
approximately 30 N and 30 S latitude (Spalding 2010; Giri et al. 2011). There is a
range of 50 to 73 different species of mangrove all over the world as the exact
number of species is still under discussion due to different classification systems
(FAO 2007; Spalding 2010).
Mangrove is considered as vascular halophytes plant as it is able to withstand or
tolerate the high salt condition (Flowers and Colmer 2015). The plant can impres-
sively grow in the harsh environment of oxygen poor intertidal coastal area as the
plant has developed the convergent adaptation in its morphology, physiology, which
includes the roots, leaves and stem anatomy and reproductive strategies. The plant
has an aerial rooting system which functions in respiration during high tide. It also
functions as anchorage and for nutrient uptake in the muddy soil (FAO 2007).
The plant adapted the high salt soil by preventing or reducing the Na+ from the
outside of the root from entering the xylem through apoplastic route by having a
wider Casparian strip and by producing a compatible solute as osmoprotectants
(Naskar and Palit 2015). Some of the adaptation to reduce salt content in the
mangrove is species-specific such as Xylocarpus and Excoecaria channel the salt
into the senescent leaves, while Aegialitis, Aegiceras and Avicennia secrete out the
salt actively using special glands through the leaves (Spalding 2010). Despite
preventing the entering and secreting out the salt from the plant, mangroves also
have succulent leaves with thick cuticle wax and large vacuole to preserve water
(Naskar and Palit 2015).
All mangroves take advantage on the water tides to disperse their seeds. How-
ever, mangrove did not undergo seed dormant stage as the plant reproduces through
vivipary where the seeds produced immediately germinate into a seedling. The
seedling will fall off from the parent tree and float until it sinks, or the tides are
low to take root and lodge in the soil (Spalding 2010; Feller 2018). The different
adaptation of the mangrove towards the regular environmental changes and different
intertidal condition has led to different diversity of mangrove plant which can be
categorized into five types of mangrove forest, namely fringe, basin, riverine,
overwash and dwarf (Spalding 2010; Feller 2018). Each of the categories is resulted
from different ecosystems and will support an incredible diversity of creatures,
including some species unique to mangrove forests.
In 1993, the world total area of mangroves estimated from 91 countries was
19.89 million hectares (Fisher and Spalding 1993) and further estimated by Giri et al.
5 Energy Flux in Mangrove Ecosystems 119

(2011) using the data in the year of 2000 from 118 countries as 13.77 million
hectares. The estimate shows a reduction up to 30.7% of the mangrove’s world
total area in just within seven years period. The reduction of the mangroves will also
leave an impact towards energy flux between the biodiversity in its ecosystem. The
unique geomorphic characteristic of mangrove grows where the land, ocean and
freshwater overlap is a great habitat or hunting ground for a wide range of organisms
(Spalding 2010). The distribution of organisms may vary according to the regions of
the world, salinity and tidal level.
The ecosystems of mangrove may involve the sustainability of some endangered
animals such as proboscis monkey (Nasalis larvatus) which is endemic to mangrove
forests of Sabah and Sarawak states of Malaysia, the pygmy three-toed sloth
(Bradypus pygmaeus), in mangrove forests at the off coast of Panama, and Bengal
tigers (Panthera tigris) in the largest mangrove forests in Sundarban of Bangladesh.
The same is true with the endangered mangrove hummingbirds (Amazilia boucardi)
that feed on the sweet nectar of the rare and vulnerable Pacific mangroves and only
grow in about a dozen patchy forests from Nicaragua to Ecuador.
Mangroves not only provide habitat, but also as spawning grounds and nurseries
which make the diversity ranging from young to adult organisms. The diversity of
organism includes reptiles, amphibian, mammals and birds. However, major parts of
mangrove food web are driven by detritus (Spalding 2010). The dry leaves that fall
from the tree are decomposed by detritivore and saprophyte which recycle the
energy flux to be re-used by the plants. This chapter aimed at highlighting the energy
fluxes in mangrove ecosystems in order to enhance our understanding of the
magnitude and changes of their regulatory mechanisms, which will help to under-
stand the mangrove ecosystem energy balance.

5.2 Energy Fluxes Within and Over Mangroves

Active regulation of water, energy and carbon fluxes between forests and the
atmosphere has been conducted over the past few years, in order to understand
both forest functioning and the role of forests as sinks or sources. Similar to all the
other forest ecosystems, mangrove trees received powerful energy sources from the
sun that supplied the energy to the earth that gave the strongest influence on other
environmental factors.
There are several energy fluxes processes that occur between Earth’s surface and
overlying atmospheres, which include (1) Thermal conduction of heat energy within
the ground, (2) Absorption and emission of ‘natural’ electromagnetic radiation by
the surface, (3) Evaporation of water stored in the soil or condensation of atmo-
spheric water vapour onto the surface and (4) Turbulent transfer of heat energy
towards or away from the surface within the atmosphere.
Therefore, each of these processes is correlated with an energy flux density
(SI unit: J m 2 s 1) that can be referred to as the rate of energy transferred
perpendicular to a surface of the unit area. The SI unit for the energy flux density
is also equivalent to Wm 2.
120 E. A. R. Engku Ariff et al.

5.3 Short-wave Radiation and Long-wave Radiation

The earth’s surface is affected by two main types of radiations, namely short-wave
(280–2800 nm) and long-wave (2800–100,000 nm) components. Short-wave radia-
tion comprises higher energy about 85% as compared to long-wave radiation at the
waveband from 0.1 to 5.0 μm, which includes ultraviolet (0.1–0.4 μm), visible light
(0.4–0.7 μm) and near-infrared (0.7–5.0 μm) spectral regions. Furthermore, short-
wave radiation is the radiation received from the sun and is composed of both the
beams (or direct) and diffuse radiations where beam radiation reaches the earth’s
surface directly from the sun after travelling through space, while diffuse radiation is
subjected to interference from any interspatial matter (Kumar et al. 1997). Short-
wave radiation is the energy source that drives evaporation, transpiration, photosyn-
thesis and many other important processes linked to agricultural systems.
The remaining 15% is long-wave radiation, which falls into the range of approxi-
mately 4–50 μm and reaches the surface of the Earth via contributions from the
atmosphere and the sun’s spectrum (Kiehl and Trenberth 1997; Wild et al. 2013).
The majority of long-wave radiation falling on the surface of the Earth is from the
atmosphere which is emitted by the gasses, especially water vapour and CO2, present
in the atmosphere.

5.4 Energy Balance

The temperature of the Earth–atmosphere system basically depends on its energy


balance on the way of absorption of solar energy and distribution between different
levels of the atmosphere and the ground. Thus, the temperature of any plant organ
such as leaf, stem and root depends on the balance between incoming energy and
also energy loss. ‘Steady state’ is where the rate of energy absorption exactly
balances the rate of energy loss and the temperature of the absorbing tissue stays
constant. The short-wave radiation entering the canopy and meet a leaf or other
elements in the canopy (e.g. root, branch, part of stem) is absorbed. A variety of
processes involve, such as evaporation, transpiration, heat conduction and photo-
synthesis. Some of the energy that absorbed goes to heating the surface and another
some is emitted back to the atmosphere as long-wave radiation or transported as
sensible heat.
The global energy balance considers the energy flows within the climate system
and their exchanges with outer space. The energy balance of a surface layer of finite
depth and unit horizontal area can be written as:

dQ
= Rn H LE G Qp
dt
Where Q is the total heat energy stored in the surface layer and Rn is the net surface
irradiance (commonly referred to as the net radiation flux, Wm 2). The Rn represents
the gain of energy by the surface from radiation and it is a positive number when it is
5 Energy Flux in Mangrove Ecosystems 121

towards the surface. G is the ground heat flux (W 2) and it is the loss of energy by
heat conduction through the lower boundary. The G is a positive number when it is
directed away from the surface into the ground. The value at the surface is denoted as
G0. H is the sensible heat flux (Wm 2) and represents the loss of energy from the
surface through heat transfer to the atmosphere. The H is positive when directed
away from the surface into the atmosphere. LE is the latent heat flux which
represents a loss of energy from the surface due to evaporation, whereas Qp
represents the energy used during photosynthesis, (Wm 2).
Obviously, the largest terms of the fluxes are Rn, H and LE and energy used in
photosynthesis accounts only a few percentages of net radiation. Net radiation, Rn
which involves division of energy into sensible and latent heat fluxes strongly
depends on the surface characteristics, vegetation functioning and weather
conditions.

5.5 Photosynthesis in Mangroves

In life, about 5% of energy reaching a leaf surface is absorbed by photosynthetic


pigments, primarily chlorophyll, and around 2% of this absorbed energy is converted
into chemical energy. The first stage of reactions involved in the driving force of
photosynthesis is called the light dependent reaction stage.
When the reaction centre of the chlorophyll a molecules reaches an excited state,
it loses its electron entirely. The losses of the electron from chlorophyll a is a crucial
for the light energy to be converted into chemical energy within the plant (Malkin
and Fork 1981; Kühlbrandt et al. 1994; Scott 2008). The capture of electron and
transduction of energy occurs between photosystem I (PSI; found mainly in the
intergranal lamellae) and photosystem II (PSII) is an extremely fast reaction of
approximately 200 to 500 picoseconds and almost exclusively independent of
temperature (Whitmarsh and Govindjee 1995). Both PSI and PSII are two
membrane-protein complexes linked by the complexes of antennae.
Besides PSI and PSII, there are two other large membrane-protein complexes
called Cytochrome b6 and ATP synthase in thylakoid membranes that are involved
in light independent reaction of photosynthesis (Moore et al. 1998; Nelson and
Ben-Shem 2004; Dekker and Boekema 2005). The light energy absorbed by photo-
synthetic pigments is first captured in transient forms of chemical energy, namely
(ATP), that along with NADPH, facilitates the enzymatic fixation and reduction of
atmospheric CO2 into sugars in the chloroplasts of the leaves. Of the chemical
energy and mass contained in these sugars, approximately 50% ends up in biomass,
primarily cellulose in tree stems, and the other half is utilized in the growth and
maintenance respiration.
Compared to other plants, the light response curve of mangroves is similar with
steep linear increase up to 300 to 400 μmol photons m 2 s 1 after which saturation
is reached. It has been shown that during low light level, mangrove photosynthesis
reaches saturation due to decrease in stomatal conductance and intercellular CO2
concentrations (Cheeseman 1994; Cheeseman and Lovelock 2004). Among the
122 E. A. R. Engku Ariff et al.

mangrove species, the rate of photosynthesis varies widely depending on the regu-
latory factors such as soil salinity, vapour pressure deficit between leaf and
surrounding which are air and light intensity (Lovelock and Ball 2002).

5.6 Mangrove Respiration

Respiration in plant refers to as the energy produced by oxidation of organic


compound resulted from energy stored during photosynthesis in living cells, which
required for growth and maintenance of tissues, absorption of mineral nutrients and
translocation of organic and inorganic materials (Hoque et al. 2010). Mangroves can
survive in various conditions of coastline ranging from the extremely arid coast of
the Persian Gulf to the west coast of Asia, where the soil types ranging from heavy
consolidated clay to coral rubble and organic peats with different level of salinity
(Clough 1992). Different conditions of the coastline may give different impacts on
energy flux, gross productivity, net productivity and primary productivity of man-
grove forests especially the availability of oxygen for the respiration.
Respirations by the plant were estimated to consume approximately 30–70% of
the total carbon fixed in photosynthesis (Hoque et al. 2010). It is widely stated that
mangroves can survive to grow in flooded anoxic soils. It is due to the mechanism of
the plants which are able to get the access towards oxygen for respiration from the air
through aerial rooting systems (Feller 2018). However, the root respiration can still
be affected by several other factors such as salinity and temperature. High level of
salinity was reported to increase leaf respiration due to the higher energy needed in
secreting out the salt or transporting the oxygen to the submerged root (Ye et al.
2005) but the root respiration decreases due to the limitation of oxygen supply which
resulted with a higher cost of metabolic activity (Burchett et al. 1989).
The changes in temperature can also affect the respiratory enzyme through kinetic
response (Hoque et al. 2010), where the respiration reduces in warm summer and the
respiration increases in cool winter, as more energy is needed in colder temperature
to maintain the metabolic activity. Mangrove forests have higher capacity in fixing
the carbon compared to terrestrial forest and producing a higher value of Qp but at
the same time would have high energy as the metabolic rate used in respiration of the
mangrove forest could be affected by several factors when it grows in different
places. Hence, the different net energy resulted from the different localization of
mangrove forest may provide different energy flux toward the consumer.

5.7 Consumers in Mangrove Ecosystems

Each ecosystem, including mangrove has its own dynamicity which develops from a
tangle of thousand other species of plants, animals, fungi and bacteria. The man-
grove itself is the engineer of the ecosystems by providing and maintaining the
physical structure of the habitat and becomes budget energy as key primary producer
(Spalding 2010; Sahu and Kathiresan 2019). Other than mangrove itself, there are
5 Energy Flux in Mangrove Ecosystems 123

various types of animals that involve in the energy flux of mangrove ecosystem,
including vertebrates (birds, reptiles, amphibians, mammals, fishes) and
invertebrates (insects, crabs, prawns).
Birds which live in mangrove ecosystem can be divided into several types,
namely aerial feeder, surface forager and foliage gleaner. The birds are important
as pollinators and as a control agent on insect pest population. In addition, insect also
was preyed by a few species of mangrove frogs. The existence of birds and frogs
enables the energy flux from insect pests as first consumer to its predator as the next
consumer. Reptiles that inhabited the mangroves including snakes, crocodiles and
alligators also involve in energy flux as they use mangrove areas as the hunting
ground. They become a predator of different animals, such as birds, fishes, mammals
and other reptiles where the energy will be further passed from the prey.
Compared to birds and reptiles, mammals are more significant component in
energy flux of mangrove ecosystems as it functions as a major food source for a
variety of animals. Mammals can become the prey for crocodiles, snakes and even
for bigger mammals such as Bengal tigers (Sandilyan and Kathiresan 2012). The
mammals also can function as seed dispersal agent when frugivore mammals such as
monkeys, bats and squirrels eat the fruits and disperse the seeds.
Other important players in energy flux of mangrove ecosystems are the aquatic
invertebrates, especially the detritivores such as crabs and bivalves (Spalding 2010).
The major part of herbivores on mangrove trees belong to detritivore which devours
the fallen mangrove leaves as it is more palatable than the fresh mangrove leaves
which protected by thick wax and many secondary metabolites such as tannin
(Spalding 2010). They break down the majority of leaf litter into detritus that acts
as fertilizer for mangrove trees. The energy flux is returned back to mangroves as the
mangrove uses its own degraded leaves as a nutrient to grow. The crabs also help to
increase the root respiration by developing a honeycomb of complex tunnel in the
mangrove soil, which leads to increase of dissolved oxygen in the soils.
Aside from the organisms mentioned above, there are also important organisms
that play a critical role in balancing the final stage of energy flux in the mangrove
ecosystems which are often overlooked. These include the microscopic life such as
fungi, protist, bacteria and archaea, which functioned as decomposer (Faridah-
Hanum et al. 2014).

5.8 Decomposer of Mangrove Ecosystems

Mangroves are considered as detritus-based ecosystems where the decomposers are


very important in cycling the nutrients efficiently from the litter of mangrove that are
sedimented in the soil and absorbed by the mangrove trees (Nazim et al. 2013).
Nutrient cycling begins when the litter accumulated are subjected to microbial
degradation. There is a critical dependence of mangrove towards decomposer as it
releases the nutrient from the decomposition of litter near to the root and maintaining
the growth of the mangroves (Spalding 2010). Unlike the herbivory and predation
that occur in the mangrove ecosystem, the cycle of nutrient provided by
124 E. A. R. Engku Ariff et al.

decomposers causes the energy flux to stay unchanged (Potapov et al. 2019). In
addition, the highly productive mangrove ecosystem can provide a high amount and
continuous amount of litter which consist of fallen leaves, branches and other debris
(Nazim et al. 2013). However, the productivity can vary according to various
climatic condition of mangroves forest in various places.
Different localization and conditions of mangroves can also affect the rate of
decomposition. Low intertidal zone, low latitude, high feeding activities of inverte-
brate, leaves with lower tannin and leaves that easier to sink show faster rate of
decomposition (Nazim et al. 2013). As the bacteria and fungi are the organisms that
contribute to the decomposition of mangrove material, high temperature is reported
to result with higher fungi growth rate and lead to a faster decomposition. Besides,
the biodiversity of decomposer and detritivores can also be affected by the quantity
and quality of detritus throughout the decomposition process. Until 2012, a sum of
120 fungi species or also called as manglicolous fungi have been recorded from
29 different mangrove forests around the world (Sandilyan and Kathiresan 2012).
The ability of decomposer to provide nutrients to the mangrove forests to increase
its growth to become a greater habitat for various organisms has brought the
importance of decomposers in mangrove ecosystem to the extent where it can
physically alter the habitat. The alteration of the habitat may affect the transfer
efficiency of energy through different trophic level of consumer and stabilize the
energy flux in mangrove ecosystems (Roy et al. 2008). Complex interactions of
these decomposers maintain the harmony of different biogeochemical processes and
sustain the nutritional status and ecological balance of mangroves (Thatoi et al.
2012).

5.9 Conclusions

An exchange of energy between the land surface and the atmosphere is an important
process in an ecosystem. In the case of the mangrove forests, this process affects the
forest ecosystem which includes photosynthesis, respiration, plant growth, water
transport and many other processes. Understanding on the latent and sensible heat
fluxes is fundamental for mangrove ecological analysis. In this chapter, the energy
flux in the multitrophic level of mangrove ecosystems from solar energy to the
producers, consumers and decomposers is discussed. Based on previous studies,
some pertinent examples of processes were highlighted on how energy from solar is
transferred and lost along the multitrophic level which is expanding previous
findings of ecological scale and complexity. It is crucial to understand the impact
of environmental change factors such as climate change, species invasion, nutrient
deposition at the trophic level and the study of energy effects. In addition, the
regulatory mechanism along with the magnitude and changes of energy flux in
mangrove forests is very important for understanding the climatological processes
at the local, global and regional levels.
5 Energy Flux in Mangrove Ecosystems 125

References
Burchett MD, Clarke CJ, Field CD, Pulkownik A (1989) Growth and respiration in two mangrove
species at a range of salinities. Physiol Planta 75(2): 299–303.
Cheeseman JM (1994) Depressions of photosynthesis in mangrove canopies. In NR Baker, JR
Bowyer, eds, Photoinhibition of Photosynthesis: From molecular mechanisms to the field. Bios
Scientific Publishers, Oxford, UK, p 377–389.
Cheeseman JM, Lovelock CE (2004) Photosynthetic characteristics of dwarf and fringe Rhizophora
mangle in a Belizean mangrove. Plant Cell Environ 27: 769–780.
Clough BF (1992) Primary productivity and growth of mangrove forests. Tropical mangrove
ecosystems. American Geophysical Union, Washington DC complex by electron crystallogra-
phy. Nature 367: 614–621.
Dekker JP, Boekema EJ (2005) Supramolecular organization of thylakoid membrane proteins in
green plants. Biochim Biophys Acta 1706: 12–39.
FAO (2007) The World’s Mangroves 1980–2005, FAO Forestry Paper 153 Rome, Forest
Resources Division, FAO. p 77.
Faridah-Hanum I, Latiff A, Hakeem KR, Ozturk M (Eds.) (2014). Mangrove ecosystems of Asia:
Status, challenges and management strategies. London/New York/Heidelberg: Springer. https://
doi.org/10.1007/978-1-4614-8582-7.
Feller I (2018) Mangroves. Retrieved from https://ocean.si.edu/ocean-life/plants-algae/mangroves.
Fisher P, Spalding MD (1993) Protected areas with mangrove habitat. Draft Report World Conser-
vation Centre, Cambridge, UK, p 60.
Flowers TJ, Colmer TD (2015) Plant salt tolerance: Adaptations in halophytes. Ann Bot 115(3):
327–331.
Giri C, Ochieng E, Tieszen LL, Zhu Z, Singh A, Loveland T, Duke N (2011) Status and distribution
of mangrove forests of the world using earth observation satellite data. Glob Ecol Biogeogr 20
(1): 154–159.
Hoque AR, Sharma S, Suwa R, Mori S, Hagihara A (2010) Seasonal variation in the size-dependent
respiration of mangroves Kandelia obovata. Mar Ecol Prog Ser 404: 31–37.
Kiehl J, Trenberth K (1997) Earth’s Annual Global Mean Energy Budget. Bull Amer Meteor Soc
78: 197–208.
Kühlbrandt W, Wang DN, Fujiyoshi Y (1994) Atomic model of plant light-harvesting complex by
electron crystallography. Nature 367(6464): 614–621.
Kumar L, Skidmore AK, Knowles E (1997) Modelling topographic variation in solar radiation in a
GIS environment. Int J Geogr Inf Sci 11(5): 475–497.
Lovelock CE, Ball MC (2002) Influence of salinity on photosynthesis of halophytes. In: Lauchli A,
Luttge U (eds) Salinity: Environment plants molecules. Kluwer Academic, New York,
p 315–339.
Malkin S, Fork DC (1981) Photosynthetic units of sun and shade plants. Plant Physiol 67: 580–583.
Moore R, Clark WD, Vodopich DS (1998) Botany. McGraw-Hill Companies, Inc.
Naskar S, Palit PK (2015) Anatomical and physiological adaptations of mangroves. Wetl Ecol
Manag 23(3): 357–370.
Nazim K, Ahmed M, Shaukat SS, Khan MU (2013) Seasonal variation of litter accumulation and
putrefaction with reference to decomposers in a mangrove forest in Karachi, Pakistan. Turk J
Bot 37(4): 735–743.
Nelson N, Ben-Shem A (2004) The complex architecture of oxygenic photosynthesis. Nat Rev Mol
Cell Biol 5: 971–982.
Potapov AM, Klarner B, Sandmann D, Widyastuti R, Scheu S (2019) Linking size spectrum, energy
flux and trophic multifunctionality in soil food webs of tropical land-use systems. J Anim Eco
88(12): 1845–1859.
Roy M, Mandal S, Ray S (2008) Detrital ontogenic model including decomposer diversity. Ecol
Model 215: 200–206.
126 E. A. R. Engku Ariff et al.

Sahu SK, Kathiresan K (2019) The age and species composition of mangrove forest directly
influence the net primary productivity and carbon sequestration potential. Biocatal Agric
Biotechnol 20: 101235.
Sandilyan S, Kathiresan K (2012) Mangrove conservation: A global perspective. Biodivers Conserv
21(14): 3523–3542.
Scott P (2008) Physiology and behaviour of plants. John Wiley & Son, Ltd, England.
Spalding M (2010) World atlas of mangroves. Routledge.
Thatoi H, Behera BC, Mishra RR, Dutta SK (2012) Biodiversity and biotechnological potential of
microorganisms from mangrove ecosystems: A review. Ann Microbiol 63(1): 1–19.
Whitmarsh J, Govindjee (1995) Photosynthesis. Encyclopedia of Applied Physics, VCH
Publishers, 13, p 513–532.
Wild M, Folini D, Schär C, Loeb N, Dutton EG, König-Langlo G (2013) The global energy balance
from a surface perspective. Clim Dyn 40: 3107–3134.
Ye Y, Tam NFY, Lu CY, Wong YS (2005) Effects of salinity on germination, seedling growth and
physiology of three salt-secreting mangrove species. Aquati Bot 83(3): 193–205.
Nitrogen and Phosphorus Budget
in Mangrove Ecosystem 6
Raghab Ray, Sandip Kumar Mukhopadhyay, and Tapan Kumar Jana

Abstract

Mangrove biogeochemistry focuses on large, slow moving reservoirs (atmo-


sphere-biosphere-geosphere-hydrosphere), and their smaller but active exchanges
of elements within the reservoirs, mainly driven by biogeochemical activities in
the geosphere or sediment. Other than carbon (C), nitrogen (N) and phosphorus
(P) involve in such active cycling of elements and play essential role in the
biogeochemical processing of organic matter and primary productivity of man-
grove ecosystems. Despite being highly productive, mangroves maintain very
low-nutrient level, which has been emphasized earlier in this chapter. Cycling of
both N and P in mangrove ecosystem has been discussed, followed by their
budget calculations. World mangroves generally act as a sink for essential
elements, storing majority into the live biomass and sediment reservoir. A
major fraction of the nutrient is conserved and recycled within the organic
structure of the ecosystem. Finally, world’s largest deltaic mangrove, the
Sundarbans is recognized as one model site for developing regional N and P
budget in view of the box model approach.

Keywords
Nitrogen · Phosphorus · Organic matter · Budget · Box model · Mangroves

R. Ray (*)
Atmosphere and Ocean Research Institute, The University of Tokyo, Kashiwa, Japan
e-mail: [email protected]
S. K. Mukhopadhyay · T. K. Jana
Department of Marine Science, University of Calcutta, Kolkata, India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 127
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_6
128 R. Ray et al.

6.1 Introduction

There are several excellent reviews and books available on comprehensive under-
standing of nitrogen (N) and phosphorus (P) cycling in the marine ecosystem
(Fenchel and Blackburn 1979). In this chapter, we summarize N, P cycling in
mangrove ecosystems by examining the storage, transformation, and fluxes in
mangrove sediment, biomass, atmosphere, and tidal waters. Specific to mangrove
ecosystems, such synoptic overview comprising both nutrients is very rare except
the pioneering one by Alongi et al. (1992), but that too lacked P budget. However,
before going into the details of mangrove N and P dynamics, an initial background
knowledge on the state of nutrients in various water systems is essential.
Over-enrichment of nutrients or “eutrophication” in lakes, reservoirs, estuaries,
and rivers are widespread all over the world and the severity is increasing. Although
the mechanisms of water eutrophication are not fully understood, yet excessive
nutrient loading via discharge of domestic wastes and non-point pollution from
agricultural practices and urban development into surface water system are consid-
ered to be among the major factors. There is a long history of such human-induced
eutrophication in the coastal waters. The negative feedback of eutrophication has
been very apparent in rivers and estuaries for centuries, as being historically
documented in a detailed landscape paintings by the Dutch artist Salomon van
Ruysdael (ca. 1648) depicting the eutrophic waterways of the Netherlands as early
as the seventeenth century (Fig. 6.1). Key nutrients of concern are nitrogen and
phosphorus because the supply rates of these nutrients most often control or “limit”
aquatic plant primary production and biomass formation (Paerl 2009). From indi-
vidual freshwater basin and coastal zone studies, e.g. the Baltic region, Mississippi
River, Gulf of Mexico, North Sea, Northern Adriatic, the Black sea (Wulff and
Stigebrandt 1989; IGBP 1997), it was observed that coastal eutrophication was a
consequence of elevated levels of waterborne N and P.
Downstream estuarine and coastal waters are physically and biogeochemically
distinct from the freshwater ecosystems and, as a result, their responses to nutrient
inputs and eutrophication may contrast to those observed in the freshwater
ecosystems (Smith 2003; Bianchi 2007). Interactive effects of bathymetry (basin
morphology), hydrology (upstream discharge and tidal mixing), collectively as
“hydrodynamics” are the physical controls, whereas minerals deposition derived
from watershed geological (erosional) and biological (plant production, microbial,
and higher trophic level cycling) processes are the geochemical drivers of nutrient
enrichment to the downstream estuaries and sea.
Coastal vegetations such as mangroves, salt marshes, sea grasses are the efficient
repositories of such nutrients derived in abundance from the external sources.
Mangroves, in particular, are often limited in N and P in order to aid their growth
and sustenance (Alongi 2011). Despite nutrient limitation, how mangroves maintain
high productivity is often considered as a “paradox,” and that unique feature has
been discussed in the next section. Because mangroves are the central focus of this
chapter, other coastal vegetations like seagrass or tidal marshes are excluded from
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 129

Fig. 6.1 “River landscape,” a painting by Salomon van Ruysdael (ca. 1648) depicting the
eutrophic waterways of the Netherlands in the seventeenth century. Note the agricultural and
human sources of nutrients and scums (presumably algal blooms) present on the water surface
(image copyright permitted by Hans Paerl, UNC-Chapel Hill)

discussing despite their importance in nutrient retention capacity and role in ecosys-
tem functioning.

6.2 Mangroves High Productivity in Low-Nutrient


Environment

Mangroves are swamped forests, occurring mostly in the tropics and subtropics
spanning between about 35 north latitude and about 40 south latitude. They
provide many ecosystem services to the coastal communities such as attenuating
inflow of flooded water and acting as a speed-breaker against storm surges,
providing enough resources for fishery and other aquatic lives. They could provide
important ecosystem service in recycling different forms of nutrients, thereby
protecting the coastal ecosystems from negative impacts of eutrophication and
atmospheric pollution. This typical functioning of mangrove ecosystems in
maintaining high productivity and nutrient recycling has particular importance for
understanding biogeochemistry of the adjacent marine ecosystems.
Mangrove are highly productive ecosystems with global carbon stock of
956 Mg C ha1, and sequestration rate of 174 g C m2 yr1 (Donato et al. 2011).
Among the wetland vegetations such as rainforests, peat swamps, salt marshes, and
seagrass meadows, mangrove C stock are much higher (Alongi 2014), and sequester
130 R. Ray et al.

at a faster rate than any other coastal habitats, e.g. sea grasses (138 g C m2 yr1,
Fourqurean et al. 2012) highlighting their importance as most proficient C fixers on
earth (Donato et al. 2011). Despite such high productivity, mangroves are often
limited by the availability of micro- and macronutrients specially N and P, mainly
due to high nutrient use efficiency for developing their cell walls (Holguin et al.
2001; Lovelock et al. 2005; Reef et al. 2010). Nutrient availability depends on
multiple environmental factors such as salinity, temperature, redox reactions
between dissolved organic and inorganic constituents in interstitial water (or pore
water), root uptake, efficiency of metabolic processes, etc. (Alongi 2013). For
instance, in highly reducing mangrove sediment, it is observed that metal sulfide
complexes readily bind to organic nutrients, and limit the amount of nutrients
available for plant uptake (Alongi 2009). Nitrogen or phosphorus limitation also
depends on tidal regime, like in the Caribbean, fringing mangroves are N-limited,
but permanently flooded forests or those deep within the islands are P-limited (Feller
et al. 2002).
Such low-nutrient availability in mangroves implies that in order to satisfy
nutrient demands, recycling rates (or turnover time) should be very fast and transfers
of nutrients efficient within the geosphere and mangrove vegetation. Nutrients are
regenerated by mangrove litter decomposition (Holguin et al. 2001). It is also seen
that increased nutrient delivery to the mangroves via anthropogenic nutrient loading,
can have negative consequences for mangrove forests and their capacity for retention
of nutrients may be limited (Reef et al. 2010). To explain such a wide range of
mangrove ecosystem properties (like nutrient recycling, transfer efficiency, and
conservation), elemental stoichiometry of major macronutrients, i.e. C, N, and P
have become an efficient tool in biogeochemical and ecological studies (Scharler
et al. 2015).

6.2.1 Implication of C:N:P

Marine phytoplankton has an average C:N:P ratio of 106:16:1 (Redfield 1934). In


contrast, terrestrial plants are characterized by C:P ratios ranging from 300 to 1300
and C:N ratios ranging from 10 to 100 for soft tissue, and C:P ratios greater than
1300 and C:N ratios ranging from 100 to 1000 for woody tissue (Hedges et al. 1986;
Ruttemberg and Goñi 1997). For the mangroves, the C:N:P stoichiometry differs
strongly between the living and non-living compartments, such as concentrations in
mangrove tissues are very low (Avicennia marina leaves: N < 1.5%, P < 0.1%, Ray
et al. 2017) and can differ significantly between locations. In the Red Sea, where
growth of mangroves is stunted, C:N:P of Avicennia marina leaves was reported to
be 1918:36:1 (Almahasheer et al. 2016), that is slightly different than the taller
A. marina in the Sundarbans (1200:40:1; Ray et al. 2017) or old growth Rhizophora
mangle in Belize (1547:47:1; Scharler et al. 2015). Furthermore, a comparative
results of stoichiometric C:N:P between the mangroves in Belize and Sundarbans
showed statistically significant differences between plant tissues, microbial mat, and
sediment ( p < 0.005) (Table 6.1). Both mangrove settings are P-limited which is
6

Table 6.1 Comparison of elemental ratio between fringe Rhizophora mangle L. (red mangrove) in Belize (Scharler et al. 2015), and Avicennia marina (black
mangroves) in the Indian Sundarbans (Ray et al. 2017)
Twin Cay, Belize Sundarbans, India
Molar ratio Leaf Litter Microbial mat Leaf Litter Root Sediment
C:N 46.7 58.4 21.7 25.3 34 38.4 11.4
C:P 1570.2 4018.5 480 930 364 900 42.8
N:P 33.6 68.8 22.1 36.8 11 11.7 3.7
C:N:P 1569:47:1 4018:69:1 479:22:1 1200:40:1 940:36:1 1162:26:1 183:8:01
Nitrogen and Phosphorus Budget in Mangrove Ecosystem
131
132 R. Ray et al.

often a common feature for such marine environment. Primary producers like plants
show higher C:N, C:P than the decomposers or heterotrophs, and overall C:N:P
decreases with increasing trophic level (Scharler et al. 2015). It is therefore believed
that nutrients are initially immobilized from the environment by the decomposers,
and as decomposition proceeds, this difference decreases further, as also evident
from a decreasing ratio in the sediment (C:N ¼ 11.4, C:P ¼ 42.8). When they
become lower than the critical level (C:N ¼ 24.5, C:P ¼ 681–979), net release of
nutrients occur (Moore et al. 2006; Parton et al. 2007). For the Sundarbans
mangroves, Ray et al. (2017) calculated carbon-use efficiency of decomposers to
be 0.86–1.0 (when sediment C:N ¼ 10.6–13) and supported the hypothesis that
decomposers adapt to low sediment C:N conditions by enhancing their carbon-use
efficiency during decomposition processes in the sediment.

6.2.2 Transfer and Conservation of N, P

Most mangrove trees are evergreen with sclerophyllous leaves and high root/shoot
biomass ratios (Komiyama et al. 2000). Root/shoot ratios in mangrove vegetation
have been observed to be an order of magnitude higher than the tropical terrestrial
forests. High root biomass in mangroves, especially the abundance of fine roots
(Poungparn et al. 2016), is conducive to nutrient uptake from low-nutrient soils.
Nutrient transfer efficiencies (TE) from one trophic level to the next, and recycling
rates (calculated as Finn’s cycling index or FCI, Finn 1980) are, in general, highest
for P (29.3–45.3%, 15.4–84%, respectively), followed by N (14.5–28%, 12.4–16.8)
and C (7.4–9.5%, 6.5–10.6%) and such difference could be even higher for the dwarf
mangroves like in the Middle-East (Scharler et al. 2015) (Table 6.2). Benthic faunal
activities of larger sizes (crabs, gastropods) and smaller sizes (leaf litter fauna) are
among the key players in such fast recycling process of the limiting nutrients
(Kristensen and Alongi 2006), and contribute to mangrove C budget (Andreetta
et al. 2014). Indeed, slow recycling and turnover rate of C (several years in the
sediment) result into their conservation within the biomass pool and sustenance of
high productivity despite thriving in nutrient-poor conditions (months of 1–2 years

Table 6.2 Finn cycling Settings Elements FCI% TE%


index (FCI or % recycled),
Fringe Mangroves C 10.6 7.8
transfer efficiency (NTE)
for three mangrove forest N 12.4 14.5
zones and nutrients (C, N, P 51 34.1
P) in Belize Transition Mangroves C 9.3 7.4
N 15 15.6
P 15.4 29.3
Dwarf Mangroves C 6.5 9.5
N 16.8 28.2
P 84.3 45.3
Data Source: Scharler et al. (2015)
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 133

of N, P turnover rates, Alongi et al. 2005). Depending on nature of nutrient recycling


within the forest, mangroves may serve as either a source or sink of nutrient to the
adjacent coastal waters (Twilley and Day 1999).

6.3 Mangroves as Source or Sink of Nutrient

Export of nutrients to the adjacent coastal environment have been well recognized by
a number of studies (Dittmar and Lara 2001; Leopold et al. 2016). This idea came
from the “Outwelling Theory” by Odum and Heald (1975) that initially was put
forward for the salt marshes showing their ability to sustain high productivity of the
adjacent water systems through a supply of nutrients and organic matter (OM). Later
this theory has been supported by the recent studies that suggest export of nutrients/
OM to be an important feature of the mangrove forests (Lee 1995). Furthermore,
porewater seepage has been suggested as one of the most important processes in the
outwelling of C, N, and P (Dittmar et al. 2006; Gleeson et al. 2013; Sippo et al.
2017). The outwelling of nutrients would occur only if export from porewater is
higher than the consumption by the benthic community and trees (Dittmar and Lara
2001). Moreover, the capacity of mangroves to release or retain nutrients depends on
geomorphological settings, such as the current velocity, frequency of inundation,
and topographic elevation (Adame and Lovelock 2011). Especially at low tide, water
body in the mangrove creek is enriched with high concentrations of ammonium
(NH4+), phosphate (PO43), and dissolved organic C and nitrogen (DOC, DON) due
to the strong influence by groundwater or pore water flow (Lara and Dittmar 1999).
Mangrove types can also cause changes in the source/sink pattern, e.g. it is seen that
fringe forests primarily act as sink for dissolved inorganic N (DIN) and a source of
DON, while basin forests may exhibit the reverse trend (Rivera Monroy et al. 1995).
Mangrove forests can also import particulate nutrients and C associated with
suspended sediment and organic debris. The capacity of mangroves to import
nutrients and C has been proposed to be an important mechanism in maintaining
the health of the adjacent seagrass communities (Valiela and Cole 2002). Adame
et al. (2010) showed that geomorphological setting could determine mangroves to
retain nutrients with some riverine site receiving more nutrients than the tidal sites.
With regard to biogeochemical flux estimates, DIN exchange between the man-
grove estuaries and coastal waters can be highly variable ranging from an export flux
of 357.14 mmol m2 yr1 to the coastal waters (Caete River, Brazil; Dittmar and
Lara 2001) to an import of 2621 mmol m2 yr1 to the mangrove basin (Taylor river,
US, Davis et al. 2001), resulting in a net global export of 42.8  117 g N m2 yr1
(Adame and Lovelock 2011). Dissolved inorganic P (DIP) exchange ranges from an
export of 20.72 mmol m2 year1 (Sundarbans, India; Ray et al. 2017) to an import
of 45.2 mmol m2 yr1 (Taylor River, US; Davis et al. 2001) resulting in a net global
import of 1.0  11 mmol m2 yr1. In the world’s largest deltaic mangroves, the
Sundarbans, net N and P export was estimated to be 264 mmol m2 yr1 and
188 mmol m2 yr1 (Indian part only; Ray et al. 2014, 2017). Mean N and P
exchange as particulate matter were reported to be 94.3  99.3 and
134 R. Ray et al.

124  285 mmol m2 yr1, respectively (Adame and Lovelock 2011) with import
rate cited for the SE Asian mangroves. Results of N and P concentrations (μmol L1)
and their exchange fluxes (mmol m2 yr1) in various mangrove locations are
shown in Table 6.3.

6.4 Mangrove N Cycle

6.4.1 Nitrogen Stock in Biomass and Sediment

Mangroves uptake N via atmospheric and belowground sources leading to its storage
within the different compartments of the live plant. Allometric regression equations
are generally used as a non-destructive method to estimate mangrove above ground
and below ground biomass (AGB and BGB) (Kauffman and Donato 2012) which
can be converted into C and N stocks. However, compared to C, very few studies
provided direct estimates of N stocks in mangrove biomass. AGB-N data
(Mg N ha1) are sparsely available from the Oceania and Asian countries such as
New Zealand (15.4; Bulmer et al. 2016), Australia (12.2, Alongi et al. 2003), Indian
Sundarbans (1.28, Ray et al. 2014), Japan (35, Khan et al. 2007), Micronesia
(56, Fujimoto et al. 1999). Very few below ground allometric functions exist for
the mangroves due to hard labor needed for extracting mangrove roots and careful
sieving (Komiyama et al. 2000). The N stock in root biomass or BGB ranges from
0.08 to 0.69 Mg N ha1 (northern Australia: 0.08–0.3; New Zealand: 0.69  0.17;
Sundarbans: 0.36  0.03; Alongi et al. 2003; Bulmer et al. 2016; Ray et al. 2014).
Similarly very few direct measurements of sediment N stock range from 0.04 to
24 Mg N ha1 (up to1.2 m depth from the surface) with maximum observed for the
Micronesian mangroves (20–24, Fujimoto et al. 1999) and minimum for the Indian
Sundarbans. Such differences in N stock in sediment generally arise from different
sampling depths considered for stock estimation, external sources of N (such as
anthropogenic input), supply of OM and decomposition rates. The AGB-N tends to
be 1.4 times as large as that in the BGB, and the sediment N stock is 3.3 times as
large as the biomass N stock (Purvaja et al. 2008). Therefore, global mean of the
ecosystem N stock (AGB + BGB + Sediment) in mangroves is calculated to be
~20 Mg N ha1. Global N cycle comprising stocks and major biogeochemical fluxes
in mangrove ecosystems are summarized schematically in Fig. 6.2.

6.4.2 Nitrogen Transformation Processes and Fluxes

6.4.2.1 Nitrogen Fixation


The mangrove biogeochemistry of plant and sediment-derived nutrients (mainly N,
P) focus on the large, slow moving chemical reservoirs and their smaller but active
exchange or cycling driven by biogeochemical activities in the mangrove reservoirs.
In particular, N cycle in mangrove benthic system is mediated predominantly by
microbial rather than chemical processes (Alongi et al. 1992). Biological N fixation
6

Table 6.3 Average N and P concentrations (μmol L1) and their net exchange fluxes (mmol m2 yr1) of their dissolved inorganic and particulate forms in
various mangrove locations, negative and positive fluxes meaning export and import, respectively
Mangrove N conc. P conc. Net N flux Net P flux
Location settings μmol L1 μmol L1 mmol m2 yr1 mmol m2 yr1 References
Dissolved inorganic fraction
Klong Ngao, River 0.43 0.13 35.00 0.19 Wattayakorn et al.
Thailand dominated 1990
Lobos bay, Mexico Tide 24.14 1.55 128.57 5.81 Sánchez-Carrillo et al.
dominated 2009
Caete River, Brazil Tide 15.00 0.52 357.14 6.13 Dittmar and Lara 2001
dominated
Sundarbans, India Tide 20.00 0.65 264.00 20.72 Ray et al. 2017
dominated
Conn creek, Tide 0.21 0.00 85.71 19.68 Ayukai et al. 1998
Australia dominated
Hinchinbrook, Tide 0.07 0.10 114.29 10.97 Boto and Wellington
Australia dominated 1988,
Ayukai et al. 1998
Nitrogen and Phosphorus Budget in Mangrove Ecosystem

Taylor River, USA Carbonate 6.00 0.10 2621.43 45.16 Davis et al. 2001
setting
S. Everglades, USA Carbonate 6.00 0.10 2.14 0.03 Sutula et al. 2003
setting
Particulate fraction
Tapi River, River – – 3592.8 60 Wattayakorn et al.
Thailand dominated 2001
Red River, Vietnam River – – 8071.4 435 Wösten et al. 2003
dominated
Data Source: Review by Adame and Lovelock (2011)
135
136 R. Ray et al.

Fig. 6.2 Global mangrove N cycle. N stocks are given in Mg N ha1, and biogeochemical fluxes of
N are presented in mmol m2 d1 except for nitrification rates** given as mmol g1 d1. BNF:
biological nitrogen fixation, AGB: above ground biomass, BGB: below ground biomass. Data are
taken from 123 mangrove sites comprising areas in the Atlantic Ocean, Caribbean Sea, Gulf of
Mexico, and Indo-Pacific coasts (Alongi et al. 1992; Purvaja et al. 2008; Reis et al. 2017). Image has
been modified from the global syntheses by Reis et al. (2017)

(BNF) is the reduction of nitrogen gas (N2) to NH4+ carried out by the Eubacteria
and Archaea that possess a required enzyme, nitrogenase (termed as diazotrophs).
The BNF has been detected in mangrove stands associated with plant roots,
sediments (free-living), microbial mats, leaf litter, pneumatophores, and
cyanobacterial crusts (free-living). Low rates of BNF in mangrove sediments were
reported than those in seagrass and salt marsh communities with rates varying from
0 to 4.9 mmol m2 d1. It has often been suggested that high DOC present in
mangrove sediment pore water could limit BNF contrast to sea grasses which
stimulate N fixation otherwise, suggesting that the N-fixing communities of seagrass
and mangroves may be dominated by different bacterial groups. Sengupta and
Chaudhuri (1991) isolated diazotrophic bacteria associated with root samples of
several mangrove species from the Indian Sundarbans. It was observed that regard-
less of mangrove species, root samples from tidally inundated mangroves sustained
greater BNF rates than the samples from occasionally inundated or drier sites,
attributing to the presence of a larger number of diazotrophs belonging to a greater
number of O2 response groups in the tidally inundated mangrove sites. Despite the
dominance of variable groups of microorganisms during N fixation (e.g., heterotro-
phic bacteria in sediment and roots, cyanobacteria in pneumatophores, mixture of
both in leaf litter), mangrove sediment, roots, pneumatophores, litter debris, and
cyanobacterial mats tend to show similar BNF rates, attesting to their insignificantly
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 137

different contributions towards the total N input in mangrove benthic system


(H ¼ 2.84, df ¼ 3,51, p ¼ 0.416; Reis et al. 2017).

6.4.2.2 Nitrogen Mineralization


N mineralization is the microbial mediated process by which organic N is converted
to inorganic N forms through a series of reactions like NH4+ production or ammoni-
fication, and NH4+ oxidation to nitrite (NO2) and nitrate (NO3) or nitrification.
Nitrifying bacteria or chemolithotrophic organisms motivate nitrification process
(e.g. species of the genera Nitrosomonas, Nitrosococcus, Nitrobacter, and
Nitrococcus). In mangrove N cycle, high magnitude of difference between the
gross and net ammonification (NH4+ production minus NH4+ utilization by
microbes) implies efficient microbial immobilization of NH4+ that may constrain
net ammonification rates in sediment. Chen and Twilley (1999) found a very strong
positive correlation of net ammonification rates with P availability in mangrove
stands in Florida, suggesting a P-limitation of microbial activities. Nitrification rates,
on other hand, could be limited by many factors, such as the NH4+ availability and
microbial immobilization, absence of root activity in rather anoxic non-vegetated
tidal flat (Kristensen et al. 1998), Mn and OC availability (Krishnan and
LokaBharathi 2009). Generally slow rates of nitrification in mangrove forests are
associated with extensive uptake of NH4+ by the mangrove below ground part and
microbes (Purvaja et al. 2008).

6.4.2.3 Nitrogen Removal


The microbial mediated N-removing pathway in mangrove sediment is important
not only because it acts to mitigate N-loading but also because it means a loss of
fixed nitrogen from such an ecosystem where N is frequently a limiting nutrient
(Holguin et al. 2001). The process, called denitrification is one of such mechanisms
that involves dissimilatory reduction of nitrate ion to nitrous oxide (N20) and N2 by
the denitrifying bacteria (e.g. some species of Serratia, Pseudomonas). In reducing
environment like mangrove sediment, denitrification step is energetically second
most favorable during the diagenesis of organic matter. In this step, NO3 is utilized
as an alternative electron source of oxygen within few centimeters of top-soil. Global
mean denitrification is higher than the BNF in sediment (0.7 versus
0.5 mmol m2 d1, Fig. 6.2) that generally suggests the N2 source of mangrove
sediment to the atmosphere. However, in most regional studies in pristine
conditions, denitrification is slower than the N mineralization, that is about 15% of
total N input to mangrove soils is denitrified. In other coastal ecosystems, the
percentage of N lost via denitrification ranges from 20% to 70% (Seitzinger 1988).
That is due to the low NO3 level in sediment and pore water, whereas mangrove
forests receiving large NO3 discharges from sewage treatment plants show rela-
tively high denitrification rates of 1 to 2-order of magnitude higher than that usually
observed (Nedwell 1975; Corredor and Morell 1994).
N removal as N2O is one of the key N transformation processes in the mangrove
benthic environment. Nitrous oxide, a major greenhouse gas, is produced as a
by-product between the microbial pathways of denitrification and nitrification.
138 R. Ray et al.

Many studies have highlighted the impact of biogenic activities and seasonal
changes on N2O emissions from mangrove sediment (Corredor and Morell 1994;
Allen et al. 2007; Chen et al. 2012; Chauhan et al. 2015). Lower N2O fluxes in
natural crab-bioturbated areas are observed in Brazilian mangroves (0.007–-
4.5 mmol m2 d1) due to constant soil oxidation by macrofauna, whereas higher
N2O fluxes (0.01–0.08 mmol m2 d1) in crab-exclusion mangrove areas are due to
wet/anaerobic soil conditions that favor denitrification (Otero et al. 2020). Global
mean N20 emission flux from mangrove sediment is ~0.007 mmol m1 d1
(Fig. 6.2).
Anaerobic ammonium oxidation (anammox) is another important microbial
mediated N removal mechanism that oxidizes NH4+ anaerobically coupled to
NO2 reduction to N2 (Dalsgaard et al. 2005). The presence of anammox activity
in mangrove sediment was noted for the first time in the north-eastern Australia
where rates of anammox were low (<2μmol N2 m2 d1) compared to other marine
environments (Meyer et al. 2005). Later in Vietnam, Amano et al. (2011) measured
anammox rate on sediment volume basis (0–0.7 nmol N2 cm3 h1) that
corresponded to only ~2% of denitrification. Anammox reactions are believed to
be inhibited by the soluble tannins or sulfides in the interstitial water (Alongi 2014).
Net tidal exchange represents the largest loss of N via mangrove waterways. This
pattern is consistent with the idea that tidal pumping and pore water seepage
transports high dissolved concentrations of nutrients into adjacent waters after the
hydrostatic pressure gradually declines towards low tide (“outwelling concept”
discussed before). Studies that quantified dissolved N fluxes between the tidal
creek or estuarine waters in mangrove areas and coastal or ocean waters indicated
that mangroves can act as a source of dissolved N for the adjacent water bodies
(Adame and Lovelock 2011). A study by Rivera Monroy et al. (1995) in the fringe
mangrove forest in Mexico observed net import of NH4+ and NO3 to the tidal creek
and basin forest, and net export of DON to the sea. Hence fringe forests might
primarily act as sink for DIN and a source of DON, while basin forests may exhibit
the opposite pattern. Higher DIN fluxes were reported for the dwarf mangroves in
Everglades (Davis et al. 2001).

6.4.3 Mangrove N Budget

Only two complete N budgets exist for the mangroves, one is the Missionary Bay
mangroves of Hinchinbrook Island in Australia, and the second is the Sundarbans
mangrove, India. The former was based on the research done by D.M Alongi and
others in early 90s (Alongi et al. 1992), and latter was a recent one by Ray et al.
(2014). In this section, N budget will be discussed based on the Bay ecosystem
which is in balance considering many extrapolations and systematic and relative
errors involved in a large number of individual measurements made over time
(synthesized by Alongi 2013). For the Sundarbans, the comprehensive mass budget
was more regional but robust, and would be discussed in detail later in this chapter.
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 139

Fig. 6.3 A mass-based


global N budget in mangrove
ecosystem, assuming total
area as15.2 million ha
(Spalding et al. 2010). Annual
flux data are retrieved from
Alongi (2013)

The existing mangrove N budget shows two main N inputs to the ecosystem and
four N outputs from the ecosystem (Fig. 6.3). The mass balance indicates that
2687 Gg N yr1 is required to sustain global mangrove net primary production
(combining wood, litter, and root) and only ~5% of the total N input is contributed
by N2 fixation. Less than 10% of mangrove N is lost via denitrification and N2O
emissions, while the majority of N loss occurs through tidal export (~60%). N fluxes
are typically well balanced with unaccounted sink of only 351 Gg N yr1 which is
very small compared to the total inputs, outputs, and sources of error. Most of the
allochthonous N (mainly tidally imported or anthropogenic) is efficiently recycled
via plant-soil-microbe pathways and 75% of mangrove N is either stored within the
organic structure (~1000–1300 Gg N yr1 in leaf litter plus root, 740 Gg N yr1 as
sediment burial) or exported to the sea (1496 Gg N yr1). In the long run, N status as
a source or sink is dependent on the balance between inputs and outputs of nutrient,
and the biogeochemical coupling between different reservoirs of a mangrove
ecosystem.

6.5 Mangrove P Cycle

The basic distinction between the N and P cycle is the absence of gaseous phases in
latter, which makes P cycle relatively simple in nature, although the relationship
between microbial activities and changes in P geochemistry can be highly complex
and difficult to measure. Furthermore, despite rapid flowing of dissolved P through
plants and animals, the processes governing their movement through the soil or sea
are very slow and make the P cycle overall one of the slowest biogeochemical cycles
(Oelkers 2008). General aspects of P cycling in estuarine and marine environments
can be found in the paper by Nixon (1980).

6.5.1 Storage and Bioavailability of Phosphorus

In mangrove ecosystem, major pools of P are live biomass (AGB and BGB) and
sediment. Mangrove sediments act as a sink for P with high retention capacity (Tam
140 R. Ray et al.

and Wong 1996). For instance, in Australia, mangrove sediment have been reported
to have adsorption maxima in the range 8.1–22.6 mmol P kg1 dry wt of sediment,
that is ~50% of the total concentration in dry sediment (Clough et al. 1983). It has
been estimated that up to 88% of the forest P pool is retained within the system in
tropical mangroves. Furthermore, total P in sediment tends to be not easily
influenced by the degradation or restoration of the wetlands, owing to its more
conservative cycling process than those of C and N (due to the lack of exchange with
the atmosphere).
In the domain of pH that is relevant to most mangrove soils (i.e. generally
between 5 and 7.5), H2PO4 and HPO42 are the dominant orthophosphate ions
(Lindsay 1979). Although organic P is the major fraction, phosphate-P represents the
largest potential pool of plant-available, soluble reactive form (Boto and Wellington
1988). Differing from soil total P concentration, plant-available soil-P plays an
important role in controlling mangrove species distributions even though compara-
tively few data exist on this topic. It has been seen that invasion of mangrove
associates (Spartina alterniflora) heavily decreases plant-available P, but exhibited
only a slight influence on the sediment total P (Feng et al. 2018). Occurrences of
phosphatase enzyme in sediment and phosphate-solubilizing bacteria associated
with mangrove roots serve an important role in providing enough phosphate to
plant biomass (Das et al. 2014).

6.5.2 Input and Output of Phosphorus

Biogeochemical input fluxes of P within the mangrove ecosystem is driven by


various processes (Pomeroy et al. 1965), such as

Input
• Atmospheric (dry and wet deposition),
• Mangrove (canopy nutrient transfer and litter fall),
• Mineralization from soil,
• Anthropogenic sources (sewage, agriculture, aquaculture, etc.).

Output
• Mangrove plant assimilation,
• Microbial uptake,
• Uptake by macro-feeder,
• Tidal exchange
• Soil immobilization/sedimentation.

A comprehensive overview of P cycle has been provided for the Malaysia peninsular
mangroves which has very impressive forest coverage globally (~3.7%). These
specific mangroves located in Merbok stand for an ideal site for P cycle where
above mentioned input and output sources of P are present, and anthropogenic inputs
of nutrient are heavy due to intensive aquaculture practices (Fig. 6.4). However, for
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 141

Fig. 6.4 Schematic representation of P cycle with stock and fluxes in a anthropogenic mangrove
environment in Malaysia. Phosphorus stock (in Mg) present in biomass (AGB and BGB) is order of
magnitude lower than the sediment. Various biogeochemical fluxes are given in Mg P yr1. (Data
Source: Yeok 2002; Gong and Ong 1990)

more pristine mangroves, Indian Sundarbans is chosen as model site in this chapter
to revisit the P budget (Ray et al. 2017, discussed in Sect. 6.6.3.3).
Primary producers mainly depend upon the internal input of P which is mobilized
from the sediments. Deposition of P in the sediment and below ground biomass takes
place through litterfall and streamflow, respectively. However, remobilization of P is
not very smooth in such sediment conditions due to their occurrence in geochemi-
cally protected forms, i.e. P either associated to Ca, Fe or Al-hydroxides or can be
adsorbed onto mineral surfaces or protected within the mineral matrices or present in
organic compounds (Ruttenberg 1992). Chemical speciation of P largely controls
biogeochemistry of this element. Studies of S, Fe, and P dynamics in wetlands
indicate a strong sulfide/reactive Fe dependency controlling the P solubility under
reduced conditions. For instance, in North Brazil mangroves, it was found that the
speciation of P with Fe/Al (P-Fe/Al) was the main chemical bound species in the
sediment (0.35  0.09 to 0.56  0.26 mg g1) compared to Ca-bound P (P-Ca)
(0.03  0.01 mg g1) (Ursula 2007). Sedimentation and subsequent immobilization
of P within the geosphere is a greater flux term than its mineralization.
Despite temporary sink of P in the mangrove estuarine sediments, river water-
driven point sources like agricultural/aquacultural run-off and wastewater discharges
are always a dominant source of P to their estuary and coastal zone (Fig. 6.4). Direct
runoff containing dissolved PO43 is also very significant, particularly when rainfall
follows the application of fertilizers in upland (Kleinman et al. 2009). During
estuarine exchange, the pore- and groundwater P can be leached by rainfall, tidal
inundation or drainage (Dittmar and Lara 2001).
142 R. Ray et al.

Phosphate enters the atmosphere from a variety of sources such as continental-


derived dust, sea spray, and plant pollen (Mahowald et al. 2005). Among these
external atmospheric sources, dry and wet deposits in aerosol system also make a
significant contribution to the total input flux. The sea-air exchange of aerosol
particles plays an important role in the global biogeochemical cycle of phosphorus
(Graham and Duce 1982). Dry deposition mangroves in the Indian Sundarbans was
estimated to be higher than the marine environment in the North Island of
New Zealand (0.12 to 27 versus 0.14 μg P cm2 yr1; Ray et al. 2017; Chen et al.
1985).

6.5.3 Mangrove P Budget

Despite the importance of P cycle in maintaining mangrove productivity, a complete


ecosystem scale budget is still very rare. A holistic scientific approach on an overall
P budget is needed to elucidate the relative importance of various pools as a source or
sink. Based on the results of P fluxes from the Merbok mangroves in Malaysia, a
budget has been calculated (Fig. 6.5).
Analysis of data from various sources, balancing the input, output, and fluxes
between the major pools of P leads to a conclusion that the mangrove ecological role
as P source/sink has been greatly masked by the huge input of P from human
activities. It is evident from the budget that municipal sewage loads in the Merbok
catchment is the largest input flux of total P (>85% of the total TP input). Agricul-
ture runoff contains high dissolved PO43 concentration (~20-fold higher than the
levels in more pristine region) that has been increased in recent decades when inputs
of anthropogenic nutrients into the coastal sea have resulted in eutrophication,
modifying aquatic food webs and causing severe hypoxic events in the coastal
environment (review by Ramesh et al. 2015).
In the budget, tidal export is the largest loss term despite over-estimation of P
output from tidal exchange due to the inherent variability of the tidal system on both
temporal and spatial scales (Yeok 2002). According to the author, this is “unrealisti-
cally high” flux, but still makes a point given the riverine mangroves, unlike tide-
dominated settings, always act as a net source of P to the estuary and sea despite their

Fig. 6.5 A mass-based P


budget in the anthropogenic
mangrove ecosystem in
Malaysia
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 143

temporary storage. A similar trend for the riverine Sundarbans is highlighted in Sect.
6.6.3.3.
Overall budget for the anthropogenically perturbed mangrove locations in
Malaysia results into a net P gain of 15.9 Mg P yr1. Growing human population
and industrializations around Asia-Pacific settings post an immediate major concern
for the overall health of the mangrove ecosystem, but still these they manage to
conserve P at net flux basis, and assign as a potential solution in hosting man-made
nutrients through conservation mechanism.

6.6 Nutrient Budget in the Sundarbans

6.6.1 Overview of the Mangrove System

Our focus here is the Sundarbans, a UNESCO heritage site and the largest del-
taic mangrove ecosystem in the world. The Sundarbans is bounded by 21 320 –
22 400 N latitude and 88 050 –89 510 E longitude covering an area of around
10,000 km2, of which approximately 60% lies in Bangladesh and 40% in India.
The Harinbhanga River forms a natural demarcation, separating the Bangladeshi and
Indian Sundarbans (Fig. 6.6). The mangrove ecosystem is characterized by high
biodiversity, monsoonal rains, flooding, delta formation, tidal influences, exposure
to super cyclones (Mandal and Hosaka 2020).
This unique ecosystem hosts a large number of flora and fauna. The forest is
particularly rich in floral biodiversity, such as Avicennia alba, Avicennia marina,
Avicennia officinalis are the predominant ones followed by Ceriops decandra,
Excoecaria agallocha, Bruguiera gymnorrhiza, Aegialitis rotundifolia, Sonneratia
apetala, Aegiceras corniculatum, Xylocarpus granatum, Heritiera fomes and man-
grove associates like Porteresia coarctata, Phoenix paludosa, Acanthus ilicifolius.
Average canopy height rarely exceeds 10 m.
Most of the rivers at the Sundarbans biosphere region flow north to south and are
influenced by the tides from the Bay of Bengal. The main estuaries from west to east
are Hooghly, Saptamukhi, Thakuran, Matla, Bidya, Ajmalmari, Bidyadhari, Gosaba,
Kalindi, and Raimangal. These rivers, apart from the Hooghly, have no direct
connection with the Ganges. Therefore, eastern part of Indian Sundarbans is more
of tide-dominated settings over river influence.
Geologically the area is a result of extensive fluvio-marine deposits of the river
Ganges and Bay of Bengal and the character of the sediment is silty clay with
composition of quartzo-feldspathic minerals (quartz, albite, microcline) contributed
from the eroded rocks of acidic composition of the drainage basin.
Climate in the region is characterized by the southwest monsoon (June–
September), north east monsoon or post-monsoon (October–January) and
pre-monsoon (February–May); 70–80% of annual rain fall occurs during the sum-
mer monsoon (South west monsoon), resulting in high river discharge (2952 and
11897 m3s1), which gradually diminishes to 900–1500 m3s1 during
non-monsoonal months (Mukhopadhyay et al. 2006).
144 R. Ray et al.

Fig. 6.6 Map of the Sundarbans mangrove covering India and Bangladesh (modified from Mandal
and Hosaka 2020)

6.6.2 State of Nutrients in the Sundarbans

Mangrove dominated estuaries in the Sundarbans are the main sources of nutrients to
the coastal water of Bay of Bengal, with significantly higher concentrations during
the monsoonal run-off periods (53% for DIN, 31% for DIP) compared to the
non-monsoon seasons (Mukhopadhyay et al. 2006). Such monsoonal enhancement
of nutrient loads is due to the anthropogenic sources derived from upland aquacul-
ture farms, waste discharge from Industry from adjacent Haldia port and domestic
sewage discharge-points of the Kolkata mega city. Despite these anthropogenic
impacts, a large section of the mangroves are typically pristine where human
interferences are minimal in the protected areas (especially the tiger reserve forest
in the east). High litter fall and its degradation and re-mineralization are the major
biogenic sources of N and P in the Sundarbans where significant positive correlation
was observed between OC and nutrients indicating (R2 ¼ 0.80; p < 0.05) sediment
in situ processes to control N, P dynamics (Ray et al. 2014, 2017). In Pichavaram
mangroves of southern India, high concentration of total N and its weak correlation
with OC was observed owing to ex situ sources (Bala Krishna Prasad and
Ramanathan 2008).
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 145

Concentrations of dissolved and sedimentary N and P in the Sundarbans are


comparable with the global mangroves and estuaries (Table 6.4). Both nitrate and
phosphate concentrations in the mangrove estuaries remained almost stable varying
between ~14 and 20 μmol L1 for DIN and between 0.5 and 1.0 μmol L1 for
dissolved PO43 during the period between 1990 and 2011 (Ghosh et al. 1992;
Mukhopadhyay et al. 2006; Chowdhury et al. 2012; Nandy et al. 2017). These
concentrations are generally found to be higher during the neap tide than the spring
tide. Monsoonal run-off increase significant amount of nutrients concentration (53%
for dissolved inorganic nitrate, 31% for dissolved inorganic phosphate) compared to
the pre-monsoon from riverine and estuaries surroundings of the Sundarbans
(Mukhopadhyay et al. 2006).

6.6.3 Nutrient Budget in the Sundarbans

Standard allometric models have been developed by Ray et al. (2011) for estimating
AGB and BGB in the Sundrabans. Plant N and P stock and their accrual to live
biomass were estimated from their concentration in dry biomass stock of different
plant parts (leaf+wood+root) and its monthly increment in the AGB and BGB during
the study period between 2009 and 2011 (details about model and method, refer to
Ray et al. 2014, 2017)

6.6.3.1 Box Model Approach


The basic characteristics of carbon and nutrient cycle are often described in terms of
the content in various reservoirs and the fluxes between them (Lerman et al. 1975).
Box model approach has been used for budgeting various biogeochemical processes
in the ocean (Frost and Franzen 1992) and estuaries (Mukhopadhyay et al. 2006).
Box models are representations of a system where quantities of materials are
depicted as uniform within each box, and the flux between them is shown with
arrows depending on their net concentrations (Rodhe and Bjorkstrom 1979). Wide
classes of natural processes like radioactive decay, many forms of chemical decom-
position, advective transport increase in a rate proportional to the number of
molecules available and in many cases the increase could be smaller than propor-
tional; for example, carbon-dependent photosynthesis in the sea is limited by
nitrogen and phosphorus. Despite mass-based box model approach being only
indicative, but it is very much instructive in pinpointing the magnitude of sinks
and sources of elements. The implications of such model could also be used to guide
management decisions with respect to a global carbon sequestration program and
nutrient state (enrichment or limitation) in the marine ecosystem.

6.6.3.2 Nitrogen Budget


A schematic box model based N cycle has been developed for the Sundarbans
considering multiple fluxes and stocks associated to the cycle (Fig. 6.7). Mangrove
ecosystem N stock (i.e. total biomass and sediment up to 60 cm depth) is 720 Gg and
most of it is in the biomass pool (98%). Major fraction of the available N in the
146

Table 6.4 Comparison of N, P concentrations in dissolved and sediment forms of world’s selective mangrove locations
Mangrove Dissolved Dissolved Sedimentary Dissolved Sedimentary
location NO3 + NO2μmol L1 NH4+μmol L1 Nμmol g1 PO43μmol L1 Pμmol g1 Reference
Sundarbans, 12–15 2–5 134–225 0.5–1 0.07–0.3 Ray et al. (2017)
India
Amazon, 2–3 5–30 n.d. 1–4 n.d. Dittmar and Lara
Brazil (2001)
SE Brazil 0.5–5.5 n.d. 207 0.1–1.1 13 Sanders et al.
(2014)
French 1.5–3.5 2.5–6 50–450 0.25–0.8 1–10 Ray et al.
Guiana (unpublished)
Vietnam 6.8–11.4 2–8 n.d. 1.2–3.1 n.d. Taillardat et al.
(2019)
Malaysia 4.5 2.5 190–415 0.2–0.8 n.d. Tanaka and Choo
(2001)
n.d: no data available
R. Ray et al.
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 147

Fig. 6.7 Box model of N cycle in the left side, and mass-based budget (right side) calculating net
sink/source of the Sundarbans. Stock are in Gg N (colored boxes, mean  SD) and fluxes in
Gg N yr1 (solid arrow). (Data Source: Ray et al. 2014)

Sundarbans mangrove ecosystem is at all times in the biomass and is recycled within
the organic structure of the biosphere-geosphere system (short N residence time
~58 days in sediment; Ray et al. 2014). Biological mechanisms help to retain and
conserve N in the forest.
The major losses or sources of N from the sediment system occur via plant
assimilation and, to a lesser extent, by N emission (probably denitrification). N
emission appears to be very low compared with N2 fixation for the Sundarbans,
although in situ experiments of these fluxes were not performed except for the
bioassay experiments of fixation (12 nmol cm2 hr1), but it appears to follow
same trend of the Missionary Bay mangroves (0–18 nmol cm2 hr1, Alongi 2009).
However, this is in contrast to the global N budget where N loss via denitrification
was a greater flux component than the fixation (Fig. 6.3), revealing diverse geomor-
phic features of the mangrove sediment conditions from location to location.
After N2 fixation, litter N input is the second largest input flux to the mangrove
sediment. Net uptake of both NOx and NH3 by the Sundarbans mangrove forest was
observed, but they altogether could account for only 2% of N required for mangrove
net production. After summing up all those input fluxes, total N gain by the sediment
is estimated to be 110.3 Gg N yr1. Total output flux or removal of N is
108.2 Gg N yr1, resulting into a net sink of 2.12 Gg N yr1. Considering
extrapolations made from the measurements over a small area of such a huge
ecosystem and also unaccounted sources (like sedimentation, volatilization, river
run-off, etc.), the input and output sums are very close with the net loss well within
the range of probable error (~5% difference). The net N sink is 1.9% of the total N
input flux, that is in line with the global percentage (12.6%, refers to Fig. 6.3).
Therefore, potentiality of such budget for the Sundarbans should be accepted with
148 R. Ray et al.

confidence, and included in the existing global data so that the more comprehensive
N budget for the mangroves could be achieved.

6.6.3.3 Phosphorus Budget


Box model representation of P cycle in the Sundarbans mangrove is presented in
Fig. 6.8. There are six coupled reservoirs of P in the model (Ray et al. 2017): P in the
form of aerosol in the atmosphere, P in the form of organic matter in AGB and BGB,
P in the form of organic and inorganic matter in soil, land estuary, and ocean.
Fundamental processes involved in the model are: (1) dust aerosol deposition from
the atmosphere and aerosol emission from the forest and surrounding mangrove
water, (2) litter input from AGB, (3) Root uptake as BGB from sediment, (4) break-
down of plant litter to inorganic P followed by their advective and diffusive
exchange between sediment and water, (5) deposition of particulate P on mud floor,
(6) export of P from the mangrove system to the coastal water.
Forest biomass and sediment P stock (up to 60 cm depth) was 49.67 Gg P, out of
which 97% was in the biomass pool. The residence time of P (¼ total mass of P in
reservoir/rate of P removal from reservoir) incorporated into the sediment is very
short (51 days) compared to the mangrove biomas (7.9 years; Ray et al. 2017)
suggesting conservation of P in the biological reservoir and its rapid recycling within
the biosphere-geosphere system.
Total P input fluxes to the sediment after summing up all sources, i.e., litter fall
(6.1), sedimentation of particulate matter (0.8), and net atmospheric deposition (9.1)
is 16.06 in Gg P yr1. The total output fluxes of P combining plant uptake (7.4),
advective transport (3.6), and export (3.7) are 14.7 Gg P yr1. The net forest

Fig. 6.8 Box model of P cycle in the left side, and mass-based budget (right side) calculating net
sink/source of the Sundarbans. Stock are in Gg P (colored boxes, mean  SD) and fluxes in
Gg P yr1 (solid arrow). (Data Source: Ray et al. 2017)
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 149

deposition of P from atmospheric sources is the largest input flux (dry deposition–
emission ¼ 10.51.4 ¼ 9.1 Gg P yr1) that also compares well with the P deposition
of long range African dust reported for the Amazon forest (6–37 Gg P yr1; Yu et al.
2015). Plant uptake accounted for 50.3% of the total output or P removal from the
sediment. It was found that in contrast to N, concentration of P in pore water was
roughly double the tidal water value and loss of P through sediment pores
characterized by advective dispersal was about sixty two-fold as large as that of
dispersal due to molecular diffusion.
The budget results into a net P gain or sink as 1.36 Gg P yr1 that corresponds to
8.4% of the total input flux. This percentage in the pristine Sundarbans is very
similar to the perturbed mangroves in Malaysia (7.4%, Fig. 6.5), suggesting an
excellent agreement of budget formulation with high confidence, and consideration
of these data in the global budget.

6.7 Conclusion and Perspective

With the rapidly changing world caused by human-induced pressures, atmospheric


temperature, CO2, and sea level rise (collectively as Climate Change), balancing the
budgets of nutrients in coastal and mangrove environment are subject to hamper.
However, at present, there are still substantial limitations on the impacts of global
change on mangroves and their relationships with the nutrients. Although there are
only few articulations available on pros and cons of global changes to nutrient
dynamics in mangrove ecosystems, experimental approaches to test the effects of
such relationship are largely descriptive than empirical.
In this context, an excellent review by Alongi (2018) is available for further dis-
cussion, where Author described in detail about how sources or sinks of N and P in
mangroves could be influenced by the global alterations, such as land-use change,
intensification of tropical storms and flooding, elevated CO2 and temperature in the
atmosphere. Author hinted that the effects of land-use change and the resulting
eutrophication would lead to changes in rates and pathways of nutrient transforma-
tion processes that in turn affect rates of net primary productivity and survival of
specific mangrove species. Recent frequency of intense storms and flooding could
result in pulses of freshwater and sediment loads to the downstream coastal water,
and relieve nutrient limitation in the mangroves by the excessive upstream discharge.
While monsoon could lower the rate of denitrification (Fernandes et al. 2013),
flooding has been suggested to release plant-available phosphate, hence affecting
nutrient limitation and availability in mangroves (Mendoza et al. 2012). Temperature
increases are likely to result in faster cycling of N and P transformation processes
because microbial growth and rates of transformation processes are closely and
positively linked to changes in temperature (Alongi 1988). Increased CO2 concen-
tration has shown positive feedback on terrestrial plant productivity (known as “CO2
fertilization”; Norby et al. 2005), but the consequences of negative feedback on
atmospheric CO2 is uncertain because of the expectation that feedbacks through N
and P cycles would reduce CO2 fertilization effect (Thornton et al. 2007). For
150 R. Ray et al.

example, C cycle model predicted 1.17- and 2-fold increase of C storage in live
biomass and sediment, respectively, in response to the hypothetical atmospheric
CO2 increment from 364 to 580 ppmv (Ray et al. 2013). Under such high CO2
condition, enhanced storage of N and P in the long-lived reservoirs can significantly
reduce their bioavailable fractions in the tight N and P budget of the Sundarbans, and
that could eventually induce a negative feedback on mangrove productivity increase
in high CO2 world.
All these arguments and assumptions on the relationship between mangrove
nutrients and global changes can be ascertained through more results from regional
and global surveys, and also by comparing time series data over the decades. A better
resolution mangrove N and P budget can only be achieved after that.

Acknowledgement Authors sincerely thank Sundarbans Biosphere Reserve and Divisional forest
office, South 24 Parganas, Govt. of West Bengal, for giving permissions to carry out the consistent
mangrove research. Authors are indebted to Prof. Hans Paerl (UNC-Chapel Hill) for providing the
image used in Fig. 6.1.

References
Adame MF, Neil D, Wright SF, Lovelock CE (2010) Sedimentation within and among mangrove
forests along a gradient of geomorphological settings. Estuar Coast Shelf Sc 86:21–30.
Adame ME, Lovelock CE (2011) Carbon and nutrient exchange of mangrove forests with the
coastal ocean. Hydrobiol 663:23–50.
Almahasheer H,Duarte CM, Irigoien X (2016) Nutrient Limitation in Central RedSea Mangroves.
Front. Mar.Sci.3:271. https://doi.org/10.3389/fmars.2016.00271
Alongi DM (1988) Bacterial production and microbial biomass in tropical mangrove sediments.
Microb Ecol 15:59–79.
Alongi DM, Boto KG, Robertson AI (1992) Nitrogen and phosphorus cycles. In: Tropical Man-
grove Ecosystems, obertson, A.I. and D.M. Alongi, eds., Coastal & Estuarine Studies
No. 41, American Geophysical Union, Washington D.C., pp. 251–292.
Alongi D, Clough B, Dixon P, Tirendi F (2003) Nutrient partitioning and storage in arid-zone
forests of the mangroves Rhizophora stylosa and Avicennia marina. Trees 17:51–60.
Alongi, D.M., B.F. Clough and A.I. Robertson (2005a) Nutrientuse efficiency in mangrove forests
along the arid tropical coast of Western Australia. Aquatic botany, 82: 121–131
Alongi, D.M. (2009) The energetics of mangrove forests. Springer, Dordrecht
Alongi, D.M. (2013) Cycling and Global Fluxes of Nitrogen in Mangroves. Global Environmental
Research, 17/2013: 173–182
Alongi DM (2014) Carbon cycling and storage in mangrove forests. Ann Rev Mar Sci 6:195–219.
Alongi DM (2018) Impact of Global Change on Nutrient Dynamics in Mangrove Forests. Forests
9:596; https://doi.org/10.3390/f9100596
Alongi DM (2011) Early growth responses of mangroves to different rates of nitrogen and
phosphorus supply. J Exp Mar Biol Ecol 397:85–93.
Allen DE, Dalal RC, Rennenberg H, Meyer RL, Reeves S, Schmidt S (2007) Spatial and temporal
variation of nitrous oxide and methane flux between subtropical mangrove sediments and the
atmosphere. Soil Biol Biochem 39:622–631.
Amano, T, Yoshinaga, I., Yamagishi, T. et al., 2011. Contribution of Anammox Bacteria to Benthic
Nitrogen Cycling in a Mangrove Forest and Shrimp Ponds, Haiphong, Vietnam. Microbes and
Environment, 26, 1–6. https://doi.org/10.1264/jsme2.ME10150
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 151

Andreetta A, Fusi M, Cameldi I, Cimò F, Carnicelli S, Cannicci S (2014) Mangrove carbon sink.
Do burrowing crabs contribute to sediment carbon storage? Evidence from a Kenyan mangrove
system. J Sea Res 85:524–533.
Ayukai T, Miller D, Wolanski E, Spagnol S (1998) Fluxes of nutrients and dissolved and particulate
organic carbon in two mangrove creeks in northeastern Australia. Mangroves and Salt Marshes
2:223–230.
Bianchi T (2007) Biogeochemistry of estuaries, 700. Oxford: Oxford University Press.
Bala Krishna Prasad M, Ramanathan AI (2008). Sedimentary nutrient dynamics in tropical estua-
rine mangrove ecosystem. Estuar Coast Shelf Sci 80:60–66.
Boto KG, Wellington JT (1988) Seasonal variation in concentrations and fluxes of dissolved
organic and inorganic materials in a tropical, tidally-dominated, mangrove waterway. Mar
Ecol Prog Ser 50:151–160.
Bulmer RH, Schwendenmann L, Lundquist CJ (2016) Carbon and Nitrogen Stocks and Below-
Ground Allometry in Temperate Mangroves. Front Mar Sci 3:150. https://doi.org/10.3389/
fmars.2016.00150.
Corredor JE, Morell JM (1994) Nitrate depuration of secondary sewage effluents in mangrove
sediments. Estuaries 17:295–300.
Chen GC, Tam NFY, Ye Y (2012) Spatial and seasonal variations of atmospheric N2O and CO2
fluxes from a subtropical mangrove swamp and their relationships with soil characteristics. Soil
Biol Biochem 48:175–181.
Chen L, Arimoto R, Duce RA (1985) The sources and forms of phosphorus in marine aerosol
particles and rain from Northern New Zealand. Atmos Environ 19:779–787.
Chen R, Twilley RR (1999) A simulation model of organic matter and nutrient accumulation in
mangrove wetland soils. Biogeochem 44:9–118.
Chauhan R, Datta A, Ramanathan A, Adhya TK (2015) Factors influencing spatio-temporal
variation of methane and nitrous oxide emission from a tropical mangrove of eastern coast of
India. Atmos Environ 107:95–106.
Chowdhury C, Majumder N, Ray R, Jana TK (2012) Inter-annual variation in some genera of
diatom and zooplankton in a mangrove ecosystem. Biodiv Conserv 21:2029–2043
Clough BF, Boto KG, Attiwill PM (1983) Mangrove and sewage: A re-evaluation. In Biology and
Ecology of Mangroves (ed. Teas, H. J.), Tasks for Vegetation Science Series, Dr W. Junk
Publishers, Lancaster, UK, pp. 151–162.
Dalsgaard T, Thamdrup B, Canfield DE (2005) Anaerobic ammonium oxidation (anammox) in the
marine environment. Res Microbiol 156:457–464.
Das S, De TK, Jana TK (2014) Vertical profile of phosphatase activity in the Sundarban mangrove
forest, North East Coast of Bay of Bengal, India. Geomicrobiol J 31:716–725.
Davis SE, Childers DL, Day JWJ, Rudnick D, Sklar F (2001) Nutrient dynamic in vegetated and
unvegetated areas of a southern Everglades mangrove creek. Estuar Coast Shelf Sci
52:753–776.
Dittmar T, Lara RJ (2001) Do mangroves rather than rivers provide nutrients to coastal
environments south of the Amazon River? Evidence from long-term flux measurements. Mar
Ecol Prog Ser 213:67–77.
Dittmar T, Hertkorn N, Kattner G, Lara RJ (2006) Mangroves, a major source of dissolved organic
carbon to the oceans. Glob Biogeochem Cy 20:GB1012. https://doi.org/10.1029/
2005GB002570.
Donato DC, Kauffman JB, Murdiyarso D, Kurnianto S, Stidham M, Kanninen M (2011)
Mangroves among the most carbon-rich forests in the tropics. Nature Geosci 4:293–297.
Feller I, McKee K, Whigham D, O’Neill J (2002) Nitrogen vs. phosphorus limitation across an
ecotonal gradient in a mangrove forest. Biogeochem 62:145–175.
Fenchel T, Blackburn TH (1979) Bacteria and Mineral Cycling, Academic Press, London, p. 225.
Feng W, Wu F, He Z, Song F, Zhu Y, Giesy JP, Wang Y, Qin N, Zhang C, Chen H, Sun F (2018)
Simulated bioavailability of phosphorus from aquatic macrophytes and phytoplankton by
152 R. Ray et al.

aqueous suspension and incubation with alkaline phosphatase. Sci Total Environ
616:1431–1439.
Fernandes SO, Gonsalves, M-J, Michotey VD, Bonin P, Loka Bharathi PA (2013) Denitrification
activity is closely linked to the total ambient Fe concentration in mangrove sediments of Goa,
India. Estuar Coast Shelf Sci 131:64–74.
Finn JT (1980) Flow analysis of models of the Hubbard Brook ecosystem. Ecology 61:562–571.
Fourqurean JW, Duarte CM, Kennedy H, Marbà N, Holmer M, Mateo MA, Apostolaki ET,
Kendrick GA, Krause-Jensen D, McGlathery KJ, Serrano O (2012) Seagrass ecosystems as a
globally significant carbon stock. Nature Geosci 5:505–509.
Frost BW, Franzen NC (1992) Grazing and iron limitation in the control of phytoplankton stock and
nutrient concentration: A chemostat analogue of the Pacific equatorial upwelling zone. Mar Ecol
Prog Ser 83:291–303.
Fujimoto K, Imaya A, Tabuchi R, Kuramoto S, Utsugi H, Murofushi T (1999) Below ground
carbon storage of Micronesian mangrove forests. Ecol Res 14:409–413.
Ghosh SK, De TK, Chowdhury A, Jana TK (1992) Distributions of nutrients in estuarine waters of
Hooghly River. Trop Ecol 23:72–77.
Gong WK, Ong JE (1990) Plant biomass & nutrient flux in a managed mangrove forest in Malaysia.
Estuar Coast Shelf Sci 31:519–530.
Graham WF, Duce RA (1982) The atmospheric transport of phosphorus to the western North
Atlantic. Atmos Environ 16:1089–1097.
Hedges JI, Clark WA, Quay PD, Richey JE, Devol AH, Santos U de M (1986) Compositions and
fluxes of particulate organic material in the Amazon River. Limnol Oceanogr. 31:717–738.
Holguin G, Vazquez P, Bashan Y (2001) The role of sediment microorganisms in the productivity,
conservation, and rehabilitation of mangrove ecosystems: An overview. Biol Fertil Soils
33:265–278.
IGBP (1997) Modelling the transport and transformation of terrestrial materials to fresh water and
coastal ecosystem, Stockholm work shop Report No. 39.
Kauffman JB, Donato DC (2012) Protocols for the measurement, monitoring and reporting of
structure, biomass and carbon stocks in mangrove forests Working Paper 86, CIFOR, Bogor,
Indonesia.
Khan MNI, Suwa R, Hagihara A (2007) Carbon and nitrogen pools in a mangrove stand of Kandelia
obovata (S., L.) Wetlands Ecol Manage (2009) 17:585–599 597 123 Yong: vertical
distributionin the soil-vegetation system. Wetlands Ecol Manage 15:141–153
Kleinman PJA, Sharpley AN, Saporito LS, Buda AR, Bryant RB (2009) Application of manure to
no-till soils: Phosphorus losses by sub-surface and surface pathways. Nutr Cycl Agroecosyst
84:215–227.
Komiyama A, Havanond S, Srisawatt W, Mochida Y, Fujimoto K, Ohnishi T, Ishihara S, Miyagi T
(2000) Top/root biomass ratio of a secondary mangrove (Ceriops tagal (Perr.) C. B. Rob.) forest.
Forest Ecol Manag 139:127–134.
Kristensen E, Jensen MH, Banta GT, Hansen K, Holmer M, King GM (1998) Transformation and
transport of inorganic nitrogen in sediments of a Southeast Asian mangrove forest. Aquat
Microb Ecol 15:165–175.
Kristensen E, Alongi DM (2006) Control by fiddler crabs (Uca vocans) and plant roots (Avicennia
marina) on carbon, iron, and sulfur biogeochemistry in mangrove sediment. Limnol Oceanogr
51:1557–1571.
Krishnan KP, LokaBharathi PA (2009) Organic carbon and iron modulate nitrification rates in
mangrove swamps of Goa, south west coast of India. Estuar Coast Shelf Sci 84:419–426.
Lara RJ, Dittmar T (1999) Nutrient dynamics in a mangrove creek (North Brazil) during the dry
season. Mangroves Salt Marshes 3:185–195.
Lee SY (1995) Mangrove outwelling: A review. Hydrobiol 295:203–212.
Lerman A, MacKenzie FT, Garrels RM (1975) Modeling of geochemical cycles: Phosphorus as an
example. Geological Society of America 142:205–218.
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 153

Leopold A, Marchand C, Deborde J, Allenbach M (2016) Water Biogeochemistry of a Mangrove-


Dominated Estuary under a Semi-Arid Climate (New Caledonia). Estuar Coast 40:773–791.
Lindsay WL (1979) Chemical equilibria in soils. John Wiley and Sons: New York, USA. p. 449.
Lovelock CE, Feller IC, Mc Kee KL, Thompson R (2005) Variation in mangrove forest structure
and sediment characteristics in Bolas del Toro, Panama. Caribb J of Sci 41:456–464.
Mahowald NM, Artaxo P, Baker AR, Jickells TD, Okin GS, Randerson JT, Townsend AR (2005)
Impacts of biomass burning emissions and land use change on Amazonian atmospheric phos-
phorus cycling and deposition, Glob Biogeochem Cy 19, GB4030. https://doi.org/10.1029/
2005GB002541.
Mandal, M.S.H. & Hosaka, T. Assessing cyclone disturbances (1988–2016) in the Sundarbans
mangrove forests using Landsat and Google Earth Engine. Nat. Hazards 102, 133–150 (2020)
Mendoza UN, da Cruz CC, Menezes MP, Lara RJ (2012) Flooding effects on phosphorus dynamics
in an Amazonian mangrove forest, Northern Brazil. Plant Soil 353:107–121.
Meyer RL, Risgaard-Petersen N, Allen DE (2005) Correlation between anammox activity and
microscale distribution of nitrite in a subtropical mangrove sediment. Appl Environ Microbiol
71:6142–6149.
Moore TR, Trofymow JA, Prescott CE, Fyles J, Titus BD (2006) Patterns of carbon, nitrogen and
phosphorus dynamics in decomposing foliar litter in Canadian forests. Ecosystems 9:46–62.
Mukhopadhyay SK, Biswas H, De TK, Jana TK (2006) Fluxes of nutrients from the tropical River
Hooghly at the land–ocean boundary of Sundarbans, NE Coast of Bay of Bengal, India. J Mar
Syst 62:9–21.
Nandy T, Mandal S, Deb S, Ghosh M, Nath T, Chatterjee M (2017) Short-term variations in surface
water properties in the Sundarban Estuarine System, India. Sustain. Water Resour Manag
4:559–566.
Nedwell DB (1975) Inorganic nitrogen-metabolism in a eutrophicated tropical mangrove estuary.
Water Res 9:221–231.
Nixon SW (1980) Between coastal marshes and coastal waters – A review of twenty years of
speculation and research on the role of salt marshesin estuarine productivity and water diversity.
In: Hamilton, R. and MacDonald, K.B. (Eds), Estuarine and Wetland Processes, pp. 437–525,
Plenum, New York.
Norby RJ, Delucia EH, Gielen B, Calfapietra C, Giardina CP, et al. (2005) Forest response to
elevated CO2 is conserved across a broad range of productivity. Proc Natur Ac Sci
102:18052–18056.
Odum WE, Heald EJ (1975) The detritus-band food web on an estuarine mangrove community. In:
Cromin, L.E. (Ed.), Estuarine Research. Academic Press, New York (265–286 pp).
Oelkers EH (2008) Phosphate mineral reactivity: From global cycles to sustainable development.
Mineralogical Magazine 72: 337–340.
Otero XL, Araújo JMC Jr, Barcellos D, Queiroz HM, Romero DJ, Nóbrega GN, Siqueira Neto M,
Ferreira TO (2020) Crab Bioturbation and Seasonality Control Nitrous Oxide Emissions in
Semiarid Mangrove Forests (Ceará, Brazil). Appl Sci 10:2215.
Paerl HW (2009) Controlling Eutrophication along the Freshwater–Marine Continuum: Dual
Nutrient (N and P) Reductions are Essential. Estuar Coast 32:593–601.
Parton W, Silver WL, Burke IC, Grassens L, Harmon ME, Currie WS, King JY, Adair EC, Brandt
LA, Hart SC, Fasth B (2007) Global-scale similarities in nitrogen release patterns during long-
term decomposition. Science 315:361–364.
Pomeroy LR, Smith EE, Grant CM (1965) The exchange of phosphate between estuarine water &
sediments. Limnol Oceanogr 10:167–172.
Poungparn S, Charoenphonphakdi T, Sangtiean T et al. (2016) Fine root production in three zones
of secondary mangrove forest in eastern Thailand. Trees 30:467–474.
Purvaja R, Ramesh R, Ray AK, Rixen T (2008) Nitrogen cycling: A review of the processes,
transformations and fluxes in coastal ecosystems. Current Sc. 94:1419–1438.
Ramesh R, Robin RS, Purvaja R (2015) An inventory on the phosphorus flux of major Indian rivers.
Current Sci 108 (10):1294–1299.
154 R. Ray et al.

Ray R, Ganguly D, Chowdhury C, Dey M, Das S, Dutta MK, Mandal SK, Majumder N, De TK,
Mukhopadhyay, SK, Jana, TK (2011) Carbon sequestration and annual increase of carbon stock
in a mangrove forest. Atmos Environ 45:5016–5024.
Ray R, Chowdhury C, Majumdar N, Dutta MK, Mukhopadhyay SK, Jana TK (2013) Improved
model calculation of atmospheric CO2 increment in affecting carbon stock of tropical mangrove
forest. Tellus B: Chem Phys Meteorol 65:18981.
Ray R, Majumder N, Das S, Chowdhury C, Jana TK (2014) Biogeochemical cycle of nitrogen in a
tropical mangrove ecosystem, east coast of India. Mar Chem 167:33–43.
Ray R, Majumder N, Chowdhury C, Das S, Jana TK (2017). Phosphorus budget of the Sundarban
mangrove ecosystem: Box model approach. Estuar Coast 41:1036–1049.
Redfield AC (1934) On the proportions of organic derivatives in sea water and their relation to the
composition of plankton. In: Laboratory LS-F (ed) James Johnstone memorial volume. Univer-
sity Press, Liverpool, pp. 176–192.
Reef R, Feller IC, Lovelock CE (2010) Nutrition of mangroves. Tree Physiology 30:1148–1160.
Reis, CRG, Nardoto, GB, Oliveira, RS. 2017. Global overview on nitrogen dynamics in mangroves
and consequences of increasing nitrogen availability for these systems. Plant Soil (2017)
410:1–19. https://doi.org/10.1007/s11104-016-3123-7
Rivera Monroy VH, Day JW, Twilley RR, Vera Herrera F, Coronado Molina C (1995) Flux of
nitrogen and sediment in a fringe mangrove forest in Terminos Lagoon, Mexico. Estuar Coast
Shelf Sci 40:139–160.
Rodhe H, Bjorkstrom A (1979) Some consequence of nonproportionality between fluxes and
reservoir contents in natural systems. Tellus 31:269–278.
Ruttenberg KC (1992) Development of a sequential extraction method for different forms of
phosphorus in marine sediments. Limnol Oceanogr 37:1460–1482.
Ruttemberg KC, Goñi MA (1997) Phosphorus distribution, C:N:P ratios, and 13 Coc in artic,
temperate, and tropical coastal sediments: Tools or characterizing bulk sedimentary organic
matter. Mar Geol 139:123–145.
Sánchez-Carrillo S, Sánchez-Andrés R, Alatorre LC, Angeler DG, Arreola-Lizárraga JA, Álvarez-
Cobelas M (2009) Nutrient fluxes in a semi-arid microtidal mangrove wetland in the Gulf of
California. Estuar Coast Shelf Sci 82: 654–662.
Sanders, C. J., et al. (2014), Elevated rates of organic carbon, nitrogen, and phosphorus accumula-
tion in a highly impacted mangrove wetland, Geophys. Res. Lett., 41, 2475–2480. https://doi.
org/10.1002/2014GL059789.
Scharler UM, Ulanowicz RE, Fogel ML et al. (2015) Variable nutrient stoichiometry (carbon:
nitrogen:phosphorus) across trophic levels determines community and ecosystem properties in
an oligotrophic mangrove system. Oecologia 179:863–876.
Seitzinger SP (1988) Denitrification in freshwater and coastal ecosystems: Ecological and geo-
chemical significance. Limnol Oceanogr 33:702–724.
Smith VH (2003) Eutrophication of freshwater and coastal marine ecosystems: A global problem.
Environ Sci Pollut Res Inter. 10:126–139.
Sengupta A, Chaudhuri S (1991) Ecology of heterotrophic dinitrogen fixation in the rhizosphere of
mangrove plant community at the Ganges River estuary in India. Oecologia 87:560–564.
Sippo JZ, Maher DT, Tait DR, Ruiz-Halpern S, Sanders CJ, Santos IR (2017) Mangrove outwelling
is a significant source of oceanic exchangeable organic carbon. Limnol Oceanog Lett 2:1–8.
Spalding, M., Kainuma, M., Collins, L., 2010. World atlas of Mangroves. Earthscan.
Sutula MA, Perez ABC, Reyes E, Childers DL, Davis S, Day JWJ, Rudnick D, Sklar F (2003).
Factors affecting spatial and temporal variability in material exchange between the Southern
Everglades wetlands and Florida Bay (USA). Estuar Coast Shelf Sci 57:757–781.
Taillardat P, Ziegler AD, Friess DA, Widory D, David F et al (2019). Assessing nutrient dynamics
in mangrove porewater and adjacent tidal creek using nitrate dual-stable isotopes: a new
approach to challenge the outwelling hypothesis? Mar. Chem. 214:103662. https://doi.org/10.
1016/j.marchem.2019.103662
6 Nitrogen and Phosphorus Budget in Mangrove Ecosystem 155

Tanaka K, Choo P-S (2001) Influences of Nutrient Outwelling from the Mangrove Swamp on the
Distribution of Phytoplankton in the Matang Mangrove Estuary, Malaysia. J Oceanograph
56:69–78. https://doi.org/10.1023/A:1011114608536
Tam NFY, Wong YS (1996) Retention of wastewater-borne nitrogen and phosphorus in mangrove
soils. Environ Technol 17:851–859.
Thornton PE, Lamarque JF, Rosenbloom NA, Mahowald NM (2007) Influence of carbon-nitrogen
cycle coupling on land model response to CO2 fertilization and climate variability. Glob
Biogeochem Cy 21:GB4018.
Twilley RR, Day JW Jr. (1999) The productivity and nutrient cycling of mangrove ecosystems,
pp. 127–152. In: A. Yáñez-Arancibia y A. L. LaraDomínguez (eds.). Ecosistemas de Manglar
en América Tropical. Instituto de Ecología A.C. México, UICN/ORMA, Costa Rica, NOAA/
NMFS Silver Spring MD USA, p. 380.
Ursula MNM (2007) Dynamics of phosphorus and sulphur in a mangrove forest in Bragança, North
Brazil. Dissertation thesis, Zentrum für Marine Tropenökologie, der Universität Bremen.
Valiela I, Cole M (2002) Comparative Evidence that Salt Marshes and Mangroves May Protect
Seagrass Meadows from Land-derived Nitrogen Loads. Ecosystems 5:92–102.
Wattayakorn, G., E. Wolanski & K. Björn, 1990. Mixing, trapping and outwelling in the Klong
Ngao Mangrove Swamp, Thailand. Estuarine, Coastal and Shelf Science 31: 667–688.
Wattayakorn G, Prapong P, Noichareon D (2001) Biogeochemical budgets and processes in
Bandon Bay, Suratthani, Thailand. J Sea Res 46:133–142.
Wösten JH, Willigen MP, Tri NH, Lien TV, Smith SV (2003). Nutrient dynamics in mangrove
areas of the Red River estuary in Vietnam. Estuar Coast Shelf Sci 57:65–72.
Wulff F, Stigebrandt A (1989) A time dependent budget model for nutrients in the Baltic Sea. Glob
Biogeochem Cy 3:53–78.
Yeok FS (2002) Phosphorus Budget in a Mangrove Ecosystem. Ph.D thesis, Universiti Sains
Malaysia.
Yu HB, Chin M, Bian HS, Yuan TL et al. (2015) Quantification of trans-Atlantic dust transport
from seven-year (2007–2013) record of CALIPSO lidar measurements. Remote Sensing Envi-
ron 159:232–249.
Mangroves as a Carbon Sink/Stocks
7
Tengku Mohd Zarawie Tengku Hashim and Mohd Nazip Suratman

Abstract

Mangroves are recognized as ecosystem that grow and dominate the coastal areas
of tropical and sub-tropical regions across the world. The high adaptability
properties of these halophytic trees enable them to thrive in harsh conditions
such as the intertidal zones. They not only provide ecological and socio-economic
support, but also play pivotal role in ecosystem function, especially in offsetting
an excess of carbon from the atmosphere. Recently, the global climate change
scenario has generated interest in understanding the carbon storage of mangroves.
Despite the crucial roles provided by mangroves, the ecosystem has degraded at
an alarming rate mainly due to climate change and anthropogenic activities. The
existence of mangroves in the coastal areas where they are considered as the most
biogeochemically active area makes them potential to store/sink a large amount of
carbon. The ability of mangroves to sink excessive carbon is reported to be more
superior from other terrestrial forests, and this could hold the key component in
mitigating global climate change. However, there is still uncertainty in
quantifying the biomass and characterizing carbon dynamics in mangroves.
Therefore, it is important to understand the functions of mangroves in reducing
the impact of climate change. Moreover, an understanding the productivity of
mangroves such as biomass, primary productivity and carbon accumulation could
have a significant impacts to this uncertainty. In this chapter, recent advancements
on the determination of mangroves carbon sinks are highlighted. Apart from that,
this paper also reviews on future challenges that are faced by the mangroves to
maintain their status as a blue carbon area.

T. M. Z. Tengku Hashim · M. N. Suratman (*)


Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 157
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_7
158 T. M. Z. Tengku Hashim and M. N. Suratman

Keywords

Mangroves · Carbon sink · Carbon stocks · Biomass · Climate change

7.1 Introduction

Since ancient time the word mangrove is believed to originate from Malay languages
mangi-mangi which mean above soil. Mangroves can be defined as an assortment of
salt tolerant plants such as trees, palms, shrubs and ferns that form a community and
flourish within transitional or intertidal zones of coastal, estuary and riverine areas of
tropical and sub-tropical regions across the globe. There is unanimity over the fact
that this halophytic ecosystem is architecturally much simpler compared to the
terrestrial forest, usually harbouring trees, shrubs, palms and scare ground ferns
with height generally exceeding one half-metre and can easily be spotted across the
coastlines (Duke 2011). These evergreen trees are a true ecotones, where it com-
monly found on mudflats and banks and can be easily identified by stands with
rooted in salty sediments where the area are frequently submerged by daily ocean
tides.
Mangroves develop numerous special adaptation capabilities to facilitate their
survival against the harsh condition of the coastal climate environment. Constant
inundation by frequent tidal action has transformed sediment in the area to become
soft and muddy, the muddy condition makes mangroves to possess a bottom-heavy
form that not only produce a high amount of biomass, but also to ensure tree can
stand still, grow straight and strong in those kinds of condition (Naidoo 2016).
According to Suratman (2008), to encounter the anoxic condition due to the water-
logged soils, mangroves adapt by having an aerial roots to aid them in gas exchange,
water uptake and give support to older trees. Certain species of mangroves such as
Avicennia and Sonneratia developed a special root system known as
pneumatophores that act as a medium to overcome the low amount of oxygen
(Naidoo 2016). This special root like pencil is equipped with lenticels that can be
found a few centimetres above the anaerobic soil, enabling the diffusion of oxygen.
One of the distinguish features uniquely possess by the mangroves that enable them
to successfully inhabit the coastal area is their unique reproductive trait known as
viviparous embryos. The mass production of viviparous propagules for species such
as Bruguiera, Rhizophora and Ceriops can maintain and produce a mass number of
individual trees by enabling the seed to germinate and mature into seedling while
still attached to the mother tree before descending into the ground floor (Kathiresan
and Bingham 2001; Naidoo 2016).
Despite their existence in the restricted coastal zone, mangroves are considered as
a natural treasure in the coastal area as it plays a vital role in supporting socio-
economics needs such as providing timber products, home to many important
commercial fisheries resources and suitable area for aquaculture activity (Hutchison
et al. 2014; Abdul Aziz et al. 2015; Venkatachalam et al. 2018). Mangroves also
prove essentials in its ability to provide crucial ecological functions that effecting
7 Mangroves as a Carbon Sink/Stocks 159

both upland communities and oceanic resources. For example, mangroves act as a
first line of defence in anticipating a storm surge and tsunami while shielding the
coastal community (Ahmadun et al. 2020). The presence of mangroves in coastal
area can be termed as a natural ecosystem engineer in reducing coastal erosion and
provide soil stabilization by binding sedimentation with their complex root system
(Horstman et al. 2015; Gracia et al. 2018). Another ecological benefit that offered by
the mangroves is by providing sound, suitable and safe nursery ground for many
high commercial aquatic inhabitants such as bream, snapper, barramundi, grouper,
banana prawns and mangroves mud crabs (Hutchison et al. 2014; Nanjo et al. 2014).
Among all of the benefits provided by the mangroves, one particular important trait
that possesses by this marine community that given less attention and always
underestimates compare to upland forests is their ability to sink excessive carbon
from the atmosphere. Mangroves have an enormous capacity for carbon storage and
considered as earth’s blue carbon sinks (Donato et al. 2011; Kauffman and Donato
2012).
Although mangroves are considered to be one of the highly productive biotopes,
have vibrant, rich and endemic biodiversity while offered so many benefits both
ecological and socio-economic, this ecosystem continue to experience losses at the
highest degree. A previous study has shown that mangroves are the most threatened
ecosystems in the whole world that caused by the calamity of global climate change
and become more susceptible when uncontrollable anthropogenic activity
intertwined (Ahmed and Glaser 2016; Das and Mandal 2016; Richards and Friess
2016). While mangroves receive a constant threat from the global climate change
factors such as sea level rises, storms and tsunamis from the past decade, the rapid
anthropogenic activities seem to be the new menace that could be the catalyst to the
destruction of mangroves worldwide. According to Alongi (2012), mangroves
around the globe experiencing high degradation about 1–3% annually, which are
driven by dense human population and poverty in the coastal area. Furthermore,
persistent hunger for more advance civilization, the human race has pushed the
mangroves to the brink of extinction as demand for aquaculture farming, human
settlement, illegal logging, agricultural activities and land development loomed large
in the coastal area.
The uncertainty of the global climate change factors nowadays that have moved
permanently outside the range of historical variation has come to the point that needs
to be given serious defining by the human race. From shifting weather patterns to
escalating combustion of fossil fuel has resulted increase in the concentration of
carbon dioxide (CO2) in the atmosphere suggest that the impact of climate change is
global in scope and unprecedented in scale. The fluxes of the CO2 concentration will
cause perturbation in the global CO2 reservoir and can have a significant impact on
the global carbon cycle and sequestration (Le Quéré et al. 2017). Furthermore, the
changes of carbon storage in the land and ocean reservoirs in response to increasing
atmospheric CO2 can be an additional fuel that could accelerate the global climate
change scenario to a whole new level. Recent concern regarding climate change and
increasing CO2 in the atmosphere has generated interest among the researchers in the
capabilities of mangroves to sink the excessive carbon concentration in the
160 T. M. Z. Tengku Hashim and M. N. Suratman

atmosphere. Despite account, only 2.4% of tropical forests, the capacity of


mangroves to store carbon are four times greater than most other tropical forests
around the world (Donato et al. 2011). Furthermore, the rapid carbon cycle, sediment
and organic material that takes place between land and seas at the coastal area
provide an opportunity for carbon sequestration potential in mangroves (Hashim
et al. 2015). With the great hype surrounding their capacity in sinking excessive
carbon, this fringe coastal community might hold the possible answer in ameliorate
the impact of climate change.
This chapter aims to provide insights regarding the potential role of mangroves as
a medium for sinking excessive carbon from the atmosphere. The perspective
function of this halophytic plant in the global carbon cycle and mitigating the global
climate change is underlined. Additionally, recent advancements on the determina-
tion of mangroves carbon sinks will be further discussed.

7.2 Ecosystem Services by Mangroves

7.2.1 Ecological Role of Mangroves

Mangroves form one of the unique wetland ecosystems that said to be most
productive and biodiverse on the Earth, which comprise both living and non-living
things. The unique adaptation ability of the mangroves not only enables them to cope
with the harsh condition of the coastal area, but also provide a tremendous ecosystem
function for the organism that lives in the area. For ages, mangroves serve as a
frontline protector in terms of their position relative to many coastal hazards. They
provide ecological support in buffering the impact of storms surge, wave activities
and tsunamis. The existence of mangroves that possess complex root structure,
width of mangroves zone and density in the coastal area acts as a blockage and
reflects part of the wave current backward to offshore will reduce the destructive
impact of the storm surge and tsunami (Zhang et al. 2012; Gracia et al. 2018).
According to Krauss et al. (2009), during the Hurricanes Charley in 2004 and Wilma
in 2004, the impact of these hurricanes when come in contact with mangroves,
reduce the storm surge height and help in reducing the destruction to the coastal area
and human society. Furthermore, the study also found out that the destruction by
Hurricane Wilma could extend more than 70% further inland without the mangroves
in the coastal area. Another study in Andaman Island during the catastrophic tsunami
in 2004 indicated that existence of mangroves in the coastal area helped to reduce the
damage caused by the tsunami only to 7%, meanwhile for the area that mangroves
have degraded from the coastal area, it was estimated that the area received 80% to
100% of the damage (Dahdouh-Guebas and Koedam 2006).
As a coastal engineer in terms of reducing the soil erosion unique root systems of
mangroves in limiting sediment exposure to wave energy and binding the sediment
help to consolidate soil that will reduce the effect of erosion and in turns promoting
soil stabilization (Horstman et al. 2015; Gracia et al. 2018). According to
Thampanya et al. (2006), Southern Thailand which had lost almost 50% of its
7 Mangroves as a Carbon Sink/Stocks 161

total mangrove since 1961 had promoted shorelines erosion by 0.01 to 0.32 km2/year
from 1967 to 1998, whereas the coastal area that dominated by the mangroves has
experienced less erosion. Cabral et al. (2017) in their study found out that the rapid
clearing of mangroves has resulted in 10% of the Mozambique coastal area is being
highly exposed to erosion.
One of the major functions of the mangroves to the aquatic habitat is by providing
sound and suitable place to life and survives. The unique structure of mangroves
such as the complex prop roots and canopy shades that exists from the combination
of branches and leaves not only form a strategic hideout spot from predators, but also
provide shelter for the aquatic organism from predation thus create more microhabi-
tat availability and increase the amount of food (Hutchison et al. 2014). Mangroves
play a crucial role as the basis of the food chains that support a wide range of marine
habitats in the coastal area. The high level of primary productivity from the man-
grove vegetation such as litter, branches and trunks and other primary producers are
important in establishing a complex food web (Hutchison et al. 2014).

7.2.2 Socio-Economic Role of Mangroves

The most tangible products from mangroves that can be commercialized are timbers.
With a characteristic of mangroves that often grow as an almost pure stand in the
coastal area made the mangroves suitable to be harvested for its timber. Timbers
extracted from mangroves are used for many commercial goods such as charcoal
production, poles for construction, industrial tannin, dye and Nipah products. Char-
coal production is the most popular industry from mangrove trees, especially
R. apiculata to produce high quality charcoal (Kridiborworn et al. 2012). Mangroves
are also sources of poles for export and local house buildings purposed. As indicated
by Ong and Gong (2013), poles from the mangroves especially from Rhizophora and
Bruguiera make excellent materials for house pillars.
The roles of mangroves in providing breeding, feeding and nursery ground for
many high values of commercial fish, shrimp and crab can be valuable assets for the
commercial fishing industry. The availability of marine fish species such as the
bream, snapper, barramundi, grouper, banana prawns that spend some part of their
life cycle in the mangroves ecosystem can contribute to sources of income to the
coastal households (Hutchison et al. 2014). According to Mohammad Abdullah et al.
(2016), in Sundarbans India, the availability of rich resources of marine and fish
species are fully utilized by the people in the area for their source of livelihoods. A
study conducted in the coastal area of Madagascar found that 87% of adults in the
area work in fisheries industry where most of the men employed in fishing and
woman actively gleaning along the shorelines (Barnes-Mauthe et al. 2013).
The existence of mangroves throughout the coastal areas of the Gulf of California
provided significant contribution to related fish species and total fisheries catch in the
range of 10–32% (Aburto-Oropeza et al. 2008). In a study, an estimate annual values
per sq. km of mangroves to fisheries was valued around USD 0.14 million to USD
6.1 million for offshore prawn and for USD 34 to USD 2.7 million for inshore
162 T. M. Z. Tengku Hashim and M. N. Suratman

coastal fisheries (Christensen et al. 2008). Das (2017) mentioned that the increased
of mangroves in the coastal area boost the annual catchment of commercial fisheries
and contributed to 15% of the total annual landing of the India, which saw the annual
monetary gained of USD 0.57 billion.

7.3 Destruction of Mangroves

Despite the potential roles of mangroves to give benefit to the environment and
surrounding coastal community, this ecosystem has degraded rapidly as compared to
the upland forests throughout the decade. Table 1 reveals global extent of previous
total world mangroves area estimate. With a constant deforestation rate of 1%
annually (Alongi and Mukhopadhyay 2015), mangrove forests experience down-
ward pattern in terms of global coverage from 18,100,000 ha in 1997 (Spalding et al.
1997) to 8,349,500 ha in 2016 (Hamilton and Casey 2016) (Table 7.1). Constant
losses of dense mangroves throughout the globe to a smaller and fragmented area
prove to have a devastating impact to the mangroves environment. This factor will
result in mangroves long essential ecosystem service and survival decline
tremendously and at great risk and this prompted the idea that mangroves may be
wiped out from its existence in the near future.
As mentioned before, the loss of mangroves around the globe is due to the impact
of global climate change and human encroachment activity. The significant losses of

Table 7.1 Global extent of previous total world mangroves area estimate
References Number of countries included Estimated total area (ha)
FAO (1981)a 51 15,642,673
FAO (1994) 65 16,221,000
Groombridge (1992) 56 16,500,000
ITTO/ISME (1993)b 87 19,847,861
Fisher and Spalding (1993) 54 12,429,115
Spalding et al. (1997) 91 18,100,077
Aizpuru and Blasco (2000)a 112c 17,075,600
FAO (2003) 112 14,653,000
FAO (2007) 124 15,231,000
Spalding et al. (2010) 123 15,236,100
Giri et al. (2011) 118 13,776,000
Hamilton and Casey (2016) 105 8,349,500
Sources: FAO (2003, 2007), Hamilton and Casey (2016) and Giri et al. (2011)
a
Except for FAO (1981) and Aizpuru and Blasco (2000), the reference year is the year of the
publications in which the estimate is cited, not the weighted average of all the national area
estimates
b
Combined figure from three publications by Clough (1993), Diop (1993), and Lacerda and Diop
(1993)
c
New estimates were provided for 21 countries, and for the remaining countries the study relied on
Spalding et al. (1997)
7 Mangroves as a Carbon Sink/Stocks 163

mangroves due to these factors are reported higher in Southeast Asia region as
compared to any part of the world. According to Hamilton and Casey (2016), five
countries in the Southeast regions out of ten in the world are accounted to be the most
countries in the world that experienced losses of its mangrove areas, where Indonesia
top the ranks (6240 ha), followed by Malaysia (2020 ha), Myanmar (1960 ha),
Thailand (390 ha) and Philippines (220 ha).

7.3.1 Anthropogenic Activities

Globally, Asian region representing the largest mangrove forests where 42% of
world totals mangrove areas are located in this region. However, rapid development
that occurs in the regions in recent decade has resulted mangrove forests in Asia
experienced 30% of reduction since 1980, which is believed to be the highest rate of
global mangrove area loss (Hamilton and Casey 2016). During the period between
the years 2000 and 2012, Southeast Asia has lost more than 100,000 ha of
mangroves at an average of 0.18% yearly where Myanmar is believed to be the
highest nation that lost their mangroves in recent decade (Richards and Friess 2016).
Meanwhile, Indonesia has stripped almost 74,900 ha or 3.11% of its mangrove area
largely in the Provinces of Kalimantan Timur and Kalimantan Selatan, largely due to
the land development, overexploitation of resources and aquaculture expansion
(Hamilton and Casey 2016).
According to Truong and Do (2018), a total of 72,825 ha of mangroves in
Mekong River Delta, Vietnam was cleared during 1980–1995 with an annual rate
of 4855 per ha for shrimp aquaculture. Malaysia also faced losses of its mangroves,
according to Hamdan et al. (2012), it was estimated that the rate of loss of mangrove
in Malaysia to be about 1% or 1282 ha/year. Furthermore, according to Romañach
et al. (2018), mangrove areas in Malaysia have lost 17% of its coverage in the coastal
area to 570,516 ha from 695,000 ha in the 1970s. For the last decade, the exponential
growth of aquaculture activities and rapid coastal development are becoming synon-
ymous with the mangrove forest in Malaysia as it becomes the primary drivers that
contribute to the loss of mangroves. In Kedah, Malaysia alone a total of 1041 ha of
mangroves permanent reserve forests was converted into an aquaculture site in order
to satisfy the demand for the aquaculture products (Ahmad and Mohammad 2005).
With the nation gunning towards the status to become the number one exporter of oil
palm product, oil palm crops expansion emerged as a new contender in Malaysia for
the destruction of mangroves. According Richards and Friess (2016), it was
estimated that 38% of mangrove loss in Malaysia was resulted from rapid expansion
of oil palm plantation.
The loss of mangroves around the world not only can be seen in Asia continent,
but also reported in other various continents. In the African continent, it was
estimated that about 13.8% equates to 500,000 ha of mangrove forest has
disappeared over the last 25 year with major losses are reported to be occurred in
the Democratic Republic of the Congo, Gabon, Sierra Leone, Guinea-Bissau and
Senegal (FAO 2007). According to FAO (2007), over the past two decades, the
164 T. M. Z. Tengku Hashim and M. N. Suratman

North and Central America has lost almost 700,000 ha of their mangrove forests,
meanwhile in the Caribbean, approximately 24% of mangrove area has degraded for
the past quarter-century due to the land development. In South America, 90% of the
mangroves are found in five countries which include Columbia, the Bolivarian
Republic of Venezuela, Ecuador, Suriname and Brazil. During the periods 1980s
and 1990s, 250,000 ha of the mangroves have lost all over region due to idea that
considers the mangrove is unproductive ecosystem, where Brazil alone in the past
25 years has lost at least 50,000 ha (around 4%) of its total mangroves (Ferreira and
Lacerda 2016).

7.3.2 Climate Change

One of unprecedented events in this decade that can have a significant impact to the
survival of mangroves is the global climate change. The changes in climate condition
at the coastal area such as an increase in sea level rise, storminess/tsunami and
precipitation not only threaten the mangrove ecosystem, but it also disturbs their
valuable ecosystem and socio-economic functions. One of the major threats that
triggered by the changing of the global climate to the coastal ecosystem is the sea
level rise. According to Church and White (2011), it was estimated that over the past
decades the global sea level rise has risen by 3.2 mm per year due to the constant
rising of heat content and continues melting of land ice. With the current trend of sea
level rise, Nicholls et al. (2011) predicted that by 2100 it could rise to more than 4 m
per year. The increase in water column resulted from sea level rise will eventually
promote constant coastal flooding thus resulted in the high occurrence of submerged
of mangroves under the seawater. The inability of mangroves to commensurate with
this situation will may cause death to the mangroves due to its sensitivity for long
inundation duration and frequency in the water (Ward et al. 2016). According to
Rahman et al. (2011), Sundarbans, India has lost almost 17,000 ha mangroves since
1970 that largely been impacted by sea level rise. Using spatial analysis, Ellison and
Zouh (2012) during the period 1957–2007 found that over two third of the shoreline
edge of mangroves in Cameron suffered dieback and 89% of Mangrove Island
located offshore was destroyed due to sea level rise. A study conducted in Guanxi,
China using the Sea Level Affecting Marshes Model (SLAMM), indicated that an
increase in sea level rise (2.9 mm per year), has resulted loss of mangrove habitat by
9.3%, 9.6% and 18.2% in 2005, 2050 and 2100, respectively (Li et al. 2015).
The continuous increase of temperature in the atmosphere resulted by the global
warming phenomenon is likely to increase in tropical cyclone and storms activities.
Coastal vegetation, especially the mangroves that inhabit the coastal area are vulner-
able to the extreme weather event such as a storm, cyclone and tsunami that will
result in loss of the mangroves. According to Long et al. (2016), the impact of
Typhoon Haiyan that struck the Philippines coast has removed about staggering
8568 ha or 3.5% of the Philippines total mangrove area. Meanwhile, similar pattern
of destruction was reported during the catastrophic Cyclone Sidr in Bangladesh.
Trees were bent and removed from the soil, large number of trees were uprooted and
7 Mangroves as a Carbon Sink/Stocks 165

overturn and the bark and twig were broken (Tanaka 2007). The catastrophic
tsunami in 2004 in the Indian Ocean has resulted mangroves loss in South Andaman
Island of India between 3825–10,000 ha followed by Aceh province, Indonesia
(300–750 ha) and Andaman Coast (306 ha) (Alongi 2008).
The global climate change is predicted to have tremendous influence in terms of
precipitation rates mainly through runoff and it was estimated that rainfall will
increase 25% more by 2050 (Gilman et al. 2008). Variable precipitation will be
further complicated by changes in temperature, influence both evaporation and
transpiration rate (Ward et al. 2016). As mentioned by Gilman et al. (2008), changes
in rainfall pattern have the potential influence on the growth, distribution and
potential extent of the mangroves. The distribution of the world’s mangroves
showed that they are productive forests with complex structure exist in an area
that receive a high amount of rainfall and high runoff as compared to the area that
have a low amount of rainfall and runoff inputs (Kumara et al. 2010). In the drier
condition, the decreased precipitation and increased evaporation will result in an
increase in soil salinity that creates saline flats, which eventually will reduce the
survival rate of seedling and growth rate and consequently resultant to mangrove
loss (Ward et al. 2016). A study conducted in the Gulf of Carpentaria suggested that
drier condition increased the salinities in the soil and resulted in destruction of
A. marina species (Conacher et al. 1996).

7.4 Mangroves as a Carbon Sink

Since the beginning of the pre-industrial era, fossil fuels extraction as a source of
energy, combustion and transportation fuel, triggered a rise of CO2 concentration
from approximately 277 ppm in 1750 to 402 ppm in 2016 (Le Quéré et al. 2017).
The radiative properties and excessive of CO2 in the atmosphere will influence the
direction of the earth’s global climate condition. With the increasing trend of
greenhouse gasses, mangroves are disproportionately important component for the
ecosystem as a medium in the global carbon cycle. Although mangroves represent
only a small fraction of the tropical forests, their unique location in the coastal area
might be crucial toward carbon biogeochemistry. Furthermore, this halophytic
plant’s capacity to sink carbon is believed to be among the best as compared to
the tropical forests (Donato et al. 2011). According to Simard et al. (2019),
mangroves were estimated to contribute to approximately 10–15% of total global
carbon storage in the coastal regions. According to Suratman (2008), the carbon
cycle that occurs in the coastal community was influenced by environmental factors
such as temperature and precipitation while for the rate of carbon cycle is determined
by the primary productivity and decomposition.
Mangroves stored carbon that is acquired from the photosynthesis process in
stems, leaves, roots and branches as a part of their biomass. In terms of primary
production in the coastal area, mangroves are believed to have higher and rapid
carbon production than their adjacent marine primary producer ecosystem (Duarte
et al. 2005). However, the rate of photosynthesis in the mangroves is varied
166 T. M. Z. Tengku Hashim and M. N. Suratman

according to species composition. In a study conducted in Tamil Nadu, India, it was


found that E. agallocha has the highest net canopy photosynthesis (21.65 gC/m2/
day) followed by R. apiculata (21.05 gC/m2/ day), while B. cylindrical (15.99
gC/m2/ day) has the lowest canopy photosynthesis rate (Sahu and Kathiresan 2019).
Mangroves are considered highly productive forests. Dead parts of the mangroves
that fall to the floor do not completely decompose. This material slowly decomposes
into a much simpler chemical substance that producing CO2, water and energy
(Fernando and Bandeira 2009). The breaking down of this organic component is a
part of the coastal community’s unique condition that comprise of waterlogged soil
condition, flora and fauna not only create a unique food web in the mangroves area
but also responsible for the carbon cycle process. Furthermore, the strategic location
of the mangroves in the coastal area, not only enable some of the excess carbon
transfer and subsidize to the adjacent ecosystems such as the ocean and beach
ecosystems, but also stored the majority of organic carbon in the soil (Bouillon
et al. 2008). The potential for storage of carbon in this ecosystem may be an
important carbon sink candidate in order to combat the climate change scenario.
With evidence from the previous literature, mangroves may hold as one of the
important candidates for carbon cycle that always been underestimated and
overlooked. With the capability of mangroves in removing CO2 from the atmosphere
and stored it as a part of their plant materials and in soils, mangroves is undoubted
contain the largest blue carbon in the coastal area. Therefore, to better understanding
the dynamic of carbon sinks of this halophytic ecosystem, it is crucial to know the
productivity of mangrove forests mainly in terms of biomass and primary
productivity.

7.4.1 Biomass

Biomass can broadly be defined as weight or mass of its total above ground living
organic matter in trees and usually expressed in a unit of a metric ton which can be
divided into two different parts, i.e., the above ground biomass (AGB) and below
ground biomass (BGB) (Walker et al. 2011). Mangroves can be categorized as an
ecosystem that has higher biomass as compared to terrestrial forests. Several studies
have been conducted around the world to quantify the biomass in mangroves. A
study conducted in the Matang Mangrove Reserve, Malaysia which was dominated
by R. apiculata species to be 480 Mg/ha (Putz and Chan 1986). Meanwhile, Hashim
et al. (2015), in their study to estimate the AGB in a dominant R. apiculata forest in
Merbok Mangrove Reserve, Malaysia found out that the AGB was 179 Mg/ha.
Although the estimation was done in an area dominated by the same type of species,
the AGB reading might be different due to the size and age of the mangroves in the
area. A plant species such as R. apiculata can do photosynthesis and rapidly store
carbon as a part of its biomass for almost 20 years during its lifespan and when the
process level is off the storage of AGB and carbon does not decline (Alongi 2012).
The mangroves that exist at lower latitudes have higher biomass as compared to
mangroves in higher latitudes. For example, a study that was conducted in Thailand
7 Mangroves as a Carbon Sink/Stocks 167

found that the AGB of the riverine mangrove forests to be 449 Mg/ha (Jachowski
et al. 2013). Similar results were also found by Kauffman et al. (2011), the estima-
tion of AGB in Micronesia yielded about 514 Mg/ha. Meanwhile, in high latitude,
the AGB estimation in Biscayne National Park, Florida, USA estimated that for
fringe mangrove areas the AGB is about 56 Mg/ha while for dwarf mangroves
22 Mg/ha (Ross et al. 2001). Komiyama et al. (2008) highlighted that low latitudes
area have greater AGB as compared to the temperate area. The influence of different
climates might explain the variation of mangroves AGB between lower and higher
latitudes. For lower latitude area, the amount of precipitation received and
favourable climate condition enables the mangroves to grow bigger and have mature
stands which contribute to higher AGB. Meanwhile different climate conditions,
frequency of storm, precipitation and temperature affect the growth of mangroves
that grow as a dwarf mangroves and yield lower AGB estimation. Distribution of
mangroves that largely in Asia region and having a suitable climatic condition could
be the most important area for carbon sinks.
To quantify the BGB in mangroves is not an easy task, with little literature
discussed in estimating it. Even though the mangroves BGB component is consid-
ered higher than their AGB and important criteria for the estimation of mangroves
biomass (Kauffman et al. 2011; Alongi 2012), the challenging task in sampling this
type of biomass such as labour intensive, time consuming and costly makes it less
popular among researchers. For example, Mackey (1993) estimated the BGB for
A. marina in the secondary forests of Australia was 121.0 Mg/ha. Meanwhile,
another study conducted in Australia for the same species in a primary forests
indicated that the BGB was about 147.3 Mg/ha while the AGB were 144.5 Mg/ha
(Briggs 1977). Other study conducted in Western and Central Sundarbans
Mangroves, India, estimated that BGB for a natural forest consisting of S. apetala,
A. alba and Excoecaria agallocha are 32.84 Mg/ha and 27.46 Mg/ha, respectively
(Banerjee et al. 2013). While a study in Bangladesh to quantify the biomass and net
primary productivity of mangroves recorded that the mean value of BGB was
84.2 Mg/ha with A. officinalis contributed the highest in terms of BGB reading
(Kamruzzaman et al. 2017).
Even though carbon is stored in various parts such as stems, shoots, roots and
down woods, soils are considered the most important parts where 50% of total
carbon stocks are stored (Kauffman and Donato 2012). It was estimated that
mangroves are carbon rich ecosystem with a total combination of tree and soil
carbon containing up to 1023 Mg/ha in the tropics, which are higher than any
other wetland ecosystems (Kauffman et al. 2011). In order to survive in the harsh
coastal climate, mangroves possess a bottom-heavy structure where much of their
biomass allocated into the root systems (Komiyama et al. 2008). According to Reef
et al. (2010), mangroves invest more fixed carbon than any plant for adaptation
purposes, such as to maximize water uptake, increase in stability and to transport
oxygen. As a result of constant carbon fixation in the root system, it is fair to assume
that mangroves store a disproportionate of their carbon underground. Furthermore,
the amount of carbon that store in the soil increased with forest age. A study
168 T. M. Z. Tengku Hashim and M. N. Suratman

conducted in the Matang Mangrove Reserve, Malaysia suggested three-fold of soil


carbon stock ranging from 385.2–545.0 Mg/ha (Adame et al. 2018).

7.4.2 Primary Productivity

The primary productivity occurs in mangroves through the photosynthesis process.


The conversion of atmospheric CO2 into organic compound is essential to producing
new mangroves parts (stems, leaves, branches and root tissue) and maintaining the
existing tissues (Alongi 2009). The main reason for the study of primary productiv-
ity in mangroves is to evaluate their carbon stocks. However, due to the mangroves
that exists in the coastal area, physical and chemical factor such as solar radiation,
salinity, fresh water sources, tides, soil type and temperature influence this process
(Twilley et al. 1992, 2017). According to Bouillon et al. (2008), it was estimated that
the global mangroves primary productivity is about 218 TgC yr. 1. In a study
conducted in the Sundarbans Mangrove forest, Bangladesh, the primary production
was estimated to be 17.2 Mg/ha yr. 1 (Kamruzzaman et al. 2017). The study
indicated that mangroves location in the oligohaline zone that frequently flushes
by tidal action might influence the results of primary productivity. Furthermore, it
was concluded that mangroves in an oligohaline zone might have high productivity
as compared to the mangroves that exist in other ecological zone (Kamruzzaman
et al. 2017).
In another study, Ross et al. (2001) reported that the net primary production for
fringe and dwarf mangroves in Florida, USA was 26.1 Mg/ha yr. 1 and 8.1 Mg/
ha yr. 1, respectively. The study concludes that the high primary production in
fringe mangroves is due to the rapid development of woody tissue compared to
dwarf mangroves which slightly slower. In a study conducted in Tamil Nadu, India,
it was reported that age plays an important role in mangroves primary productivity
(Sahu and Kathiresan 2019). The studies estimated that for young mangrove trees in
the area, the primary productivity was 30.80 Mg/ha yr. 1 while for mature stand it
was 17.04 Mg/ha yr. 1. Another study conducted by Putz and Chan (1986) in
estimating the tree growth and productivity in mature R. apiculata stands in Matang
Mangrove Forests, Malaysia, found that the net primary productivity was about
17.7 Mg/ha yr. 1. The best possible explanation for this trend might be when
mangroves reach its maturity, their ability to sequester carbon declines, as for
young mangroves the rate of sequestration increased.

7.5 Recent Advancements in Mangrove Carbon Sinks Studies

The exponential growth of remote sensing technology in recent decades has driven
many forest ecologists and stakeholders to incorporate the usage of imagery data for
many forest applications, especially in biomass monitoring in mangroves. Even
though the popular traditional method such as destructive and allometric functions
can yield better and accurate biomass results, it can measure only to small scale areas
7 Mangroves as a Carbon Sink/Stocks 169

and cannot be applied to a larger area. Moreover, the treacherous mangrove areas
that are almost impossible to access, time consuming, labour intensive and can be
costly made the traditional method less relevance in today inventorying and moni-
toring purposes for biomass and carbon study. Therefore, remote sensing technology
provides better spatially explicit, and can be efficiently combined forest biomass
estimates and has the potential to give input at the large range of spatial and temporal
scales (Galidaki et al. 2017).
Over recent decades, many literatures and publication have grown and applied the
relationships between remote sensing parameters and mangroves attributes in deter-
mining the mangrove biomass. A relationship between the biomass and
multiimagery satellite wavelength data was established during the study conducted
in a mangrove replanting site in Jiulong River Estuary, China (Wang et al. 2018).
The study found out that there is a significant correlation between AGB and Landsat
bands 3 and 4. This study is an agreement with an earlier study that was conducted
by Wu et al. (2016), where they conclude that the individual Landsat 8 band showed
a promising correlation with AGB and carbon stocks where shortwave infra-red was
the highest correlated ( 0.57). A similar finding was also reported in the study in
estimating the biomass and carbon stocks of the Matang Mangrove Forests Malaysia
that utilized the Landsat TM and SPOT 5 data, where it was found that there are
significant relationships between the vegetation indices and the forest variables
(Hamdan et al. 2013). This concludes that satellite image data can have enormous
potential in determining the biomass and carbon stocks in the mangroves which
eventually help in understanding the potential of carbon sinks in mangroves.
A medium resolution sensor such as the Landsat can prove to be a good medium
in estimating the potential of carbon sinks in mangroves. A study that was conducted
in the Merbok Mangrove Reserves, Malaysia using Landsat 8 data to predict the
carbon stocks of mangroves found out that the estimation in the range of 16.88 Mg/
ha to 138.20 Mg/ha with R2 0.56 and RMSE 22.24 Mg/ha (Hashim et al. 2020). A
study conducted in West Kalimantan, Indonesia in their attempt to study the standing
biomass of mangroves in the area using Landsat 8 data found out that the biomass in
the region was 45 Mg/ha to 100 Mg/ha (Yusandi et al. 2018). A recent study
conducted in the Matang Mangrove Forests Malaysia using the SPOT-5 sensor for
AGB modelling estimated that AGB ranged between 33.65 and 437.46 Mg/ha with
an average of 133.97 Mg/ha while for carbon stocks 16.86–218.73 Mg/ha (Muhd-
Ekhzarizal et al. 2018). A study to estimate the mangroves AGB using a traditional
field data collection and Unmanned Aerial Vehicle (UAV) method in two separate
areas which are categorized as productive and protective area were done in the
Matang Mangrove Forests, Malaysia (Otero et al. 2018). The study reported that for
productive area yielded an estimate of AGB 217 Mg/ha for UAV and 238 for field
inventory. Meanwhile, in the protective area, the estimate of AGB was 210 Mg/ha
using UAV and 147 Mg/ha for field inventory.
According to Hamdan et al. (2013), during the study of carbon stocks in the
Matang Mangrove Forests, Malaysia using Landsat 5 TM and SPOT-5 with different
year interval found out that carbon stocks in the area ranged from 1.03 to 263.65 Mg/
ha for year 1991 and 1.01–259.68 Mg/ha for the year 2011. Meanwhile Hamdan
170 T. M. Z. Tengku Hashim and M. N. Suratman

et al. (2014) predicted that using L-band ALOS PALSAR sensor in estimating the
value of AGB in the Matang Mangrove Forests, Malaysia was ranging from 2.98 to
378.32 Mg/ha meanwhile the carbon stocks was ranging from 1.49 to 189.16 Mg/ha.
Another study that conducted in a mangrove plantation area in North Vietnam to
estimate the AGB of mangroves using a combination of ALOS-2 PALSAR-2,
Sentinel-2A data and machine learning approach yielded AGB in the range of
36.22 Mg/ha to 230.14 Mg/ha with an average of 87.67 Mg/ha (Pham et al. 2018).
Furthermore, the study concluded that the combination of satellite data and machine
learning approach can be useful tools in estimating the mangroves AGB. Biomass in
the mangroves can also be determined using the Multifrequency Radar data. The
unique ability of radar to penetrate the vegetation canopies and to interact with tree
parts and underlying water surface make radar an interesting tool to gather three-
dimensional information in mangrove forests (Proisy et al. 2003).

7.6 Conclusions

The main objectives of this chapter were to review the role of mangroves as a carbon
sink ecosystem as it might be the potential answer to mitigate the global climate
change scenario. Studies from the past indicated that mangrove ecosystems could
contribute to important blue carbon sinks in the coastal area as the carbon stored as
part of their AGB and BGB. Furthermore, the potential of high primary productivity
in young mangroves stands while enormous storage of carbon in mature stands
suggested this ecosystem have a role to play in the global carbon budget.
The destruction of mangrove ecosystems due to the impacts of the uncertainty of
global climate change and encroachment by human activity are supposed to be the
main catalyst for the destruction of mangroves worldwide. This trend is expected to
increase in the near future and put more stress on mangrove ecosystems as coastal
development, human population and aquaculture industry expend further into the
mangrove areas. The destruction of mangroves could turn the mangroves to become
a source of carbon that might further accelerate global climate change. Therefore,
strategic mitigation such as reducing forest degradation, reforestation and sustain-
able management of existing mangrove can increase their capacity in the global
carbon cycle. Furthermore, the usage of remote sensing technology in monitoring
and studying the mangroves carbon sinks and ecosystem dynamics can be a crucial
solution to maintain their status as the largest blue carbon reservoir in the
coastal area.

References
Abdul Aziz A, Dargusch P, Phinn S, Ward A (2015) Using REDD+ to balance timber production
with conservation objectives in a mangrove forest in Malaysia. Ecol Econ 120: 108–116.
Aburto-Oropeza O, Ezcurra E, Danemann G, Valdez V, Murray J, Sala E (2008) Mangroves in the
Gulf of California increase fishery yields. Proc Nat Acad Sci 105: 10456–10459.
7 Mangroves as a Carbon Sink/Stocks 171

Adame MF, Zakaria RM, Fry B, Chong VC, Then YHA, Brown CJ, Lee SY (2018) Loss and
recovery of carbon and nitrogen after mangrove clearing. Ocean Coast Manag 161: 117–126.
Ahmad Y, Mohammad A (2005) Management and conservation of mangrove: Kedah experience.
sustainable management of Matang Mangroves: 100 years and beyond. Forestry Department
Peninsular Malaysia, p. 81–90.
Ahmadun FL-R, Wong MMR, Mat SA (2020) Consequences of the 2004 Indian Ocean Tsunami in
Malaysia. Safety Science 121: 619–631.
Ahmed N, Glaser M (2016) Coastal aquaculture, mangrove deforestation and blue carbon
emissions: Is REDD+ a solution? Mar Pol 66: 58–66.
Aizpuru M, Blasco F (2000) Global assessment of cover change of the mangrove forests using
satellite imagery at medium to high resolution. In: EEC research project no 15017-1999-05
FIED ISP FR. Joint Research Center, Ispra.
Alongi DM (2009) The energetics of mangrove forests, Springer Science & Business Media.
Alongi DM (2008) Mangrove forests: Resilience, protection from tsunamis, and responses to global
climate change. Estuar Coas Shelf Sci 76: 1–13.
Alongi DM (2012) Carbon sequestration in mangrove forests. Carbon Manag 3: 313–322.
Alongi DM, Mukhopadhyay SK (2015) Contribution of mangroves to coastal carbon cycling in low
latitude seas. Agric For Meteorol 213: 266–272.
Banerjee K, Sengupta K, Raha A, Mitra A (2013) Salinity based allometric equations for biomass
estimation of Sundarban mangroves. Biomass Bioenergy 56: 382–391.
Barnes-Mauthe M, Oleson KL, Zafindrasilivonona B (2013) The total economic value of small-
scale fisheries with a characterization of post-landing trends: An application in Madagascar with
global relevance. Fish Res 147: 175–185.
Bouillon S, Borges AV, Castañeda Moya E, Diele K, Dittmar T, Duke NC, Kristensen E, Lee SY,
Marchand C, Middelburg JJ (2008) Mangrove production and carbon sinks: A revision of global
budget estimates. Global biogeochemical cycles 22.
Briggs S (1977) Estimates of biomass in a temperate mangrove community. Austr J Ecol 2:
369–373.
Cabral P, Augusto G, Akande A, Costa A, Amade N, Niquisse S, Atumane A, Cuna A, Kazemi K,
Mlucasse R, Santha R (2017) Assessing Mozambique’s exposure to coastal climate hazards and
erosion. Int J Disaster Risk Reduct 23: 45–52.
Christensen SM, Tarp P, Hjortsø CN (2008) Mangrove forest management planning in coastal
buffer and conservation zones, Vietnam: A multimethodological approach incorporating multi-
ple stakeholders. Ocean Coast Manag 51: 712–726.
Church JA, White NJ (2011) Sea-level rise from the late 19th to the early 21st century. Surv
Geophys 32: 585–602.
Clough B (1993) The economic and environmental values of mangrove forests and their present
state of conservation in the South-East Asia/Pacific Region. Mangrove Ecosystems Technical
Reports. International Society for Mangrove Ecosystems, Okinawa, Japan.
Conacher C, O’brien C, Horrocks J, Kenyon R (1996) Litter production and accumulation in
stressed mangrove communities in the Embley River estuary, North-eastern Gulf of Carpentaria,
Australia. Mar Freshw Res 47: 737–743.
Dahdouh-Guebas F, Koedam N (2006) Coastal vegetation and the Asian tsunami. Science 311:
37–38.
Das CS, Mandal RN (2016) Coastal people and mangroves ecosystem resources vis-à-vis manage-
ment strategies in Indian Sundarban. Ocean Coast Manag 134: 1–10.
Das S (2017) Ecological restoration and livelihood: Contribution of planted mangroves as nursery
and habitat for artisanal and commercial fishery. World Dev 94: 492–502.
Diop E (1993) Conservation and sustainable utilization of mangrove forests in Latin America and
Africa Regions. Part II. Africa. Mangrove Ecosystems Technical Reports No. 3, ITTO/ISME
Project PD114/90. Nishihara, Japan, International Society for Mangrove Ecosystems (ISME) &
International Tropical Timber Organization (ITTO).
172 T. M. Z. Tengku Hashim and M. N. Suratman

Donato DC, Kauffman JB, Murdiyarso D, Kurnianto S, Stidham M, Kanninen M (2011)


Mangroves among the most carbon-rich forests in the tropics. Nat Geosci 4: 293.
Duarte CM, Middelburg JJ, Caraco N (2005) Major role of marine vegetation on the oceanic carbon
cycle. Biogeosciences 2: 1–8.
Duke NC (2011) Mangroves. In: HOPLEY, D. (ed.) Encyclopedia of Modern Coral Reefs:
Structure, Form and Process. Dordrecht: Springer Netherlands.
Ellison J, Zouh I (2012) Vulnerability to climate change of mangroves: Assessment from
Cameroon, Central Africa. Biology 1: 617–638.
FAO (1981) Tropical forest resources assessment project. Forest resources of tropical Africa. Part
II: Country Briefs. FAO, UNEP.
FAO (1994) Mangrove forest management guidelines. FAO Forestry Paper 117, Rome, 319 pp.
FAO (2003) FAO’s database on mangrove area estimates. Forest Resources Assessment Working
Paper No. 62. Forest Resources Division, FAO, Rome (Unpublished).
FAO (2007) The world’s mangroves 1980–2005. FAO Rome, Italy.
Fernando SM, Bandeira SO (2009) Litter fall and decomposition of mangrove species Avicennia
marina and Rhizophora mucronata in Maputo Bay, Mozambique. West Indian Ocean J Mar Sci
8:139–145.
Ferreira AC, Lacerda LD (2016) Degradation and conservation of Brazilian mangroves, status and
perspectives. Ocean Coast Manag 125: 38–46.
Fisher P, Spalding M (1993) Protected areas with mangrove habitat. Unpublished report to WCMC,
Cambridge, UK.
Galidaki G, Zianis D, Gitas I, Radoglou K, Karathanassi V, Tsakiri-Strati M, Woodhouse I,
Mallinis G (2017) Vegetation biomass estimation with remote sensing: Focus on forest and
other wooded land over the Mediterranean ecosystem. Int J Remote Sens 38: 1940–1966.
Gilman EL, Ellison J, Duke NC, Field C (2008) Threats to mangroves from climate change and
adaptation options: A review. Aquat Bot 89: 237–250.
Giri C, Ochieng E, Tieszen LL, Zhu Z, Singh A, Loveland TR, Masek J, Duke NC (2011) Status
and distribution of mangrove forests of the world using earth observation satellite data. Glob
Ecol Biogeogr 20: 154–159.
Gracia A, Rangel-Buitrago N, Oakley JA, Williams AT (2018) Use of ecosystems in coastal erosion
management. Ocean Coast Manag 156: 277–289.
Groombridge B (1992) Global biodiversity: Status of the earth’s living resources. Chapman & Hall.
Hamdan O, Aziz HK, Hasmadi IM (2014) L-band ALOS PALSAR for biomass estimation of
Matang Mangroves, Malaysia. Remote Sens Environ 155: 69–78.
Hamdan O, Khairunnisa M, Ammar A, Hasmadi IM, Aziz HK (2013) Mangrove carbon stock
assessment by optical satellite imagery. J Trop For Sci 25: 554–565.
Hamdan O, Khali Aziz H, Shamsudin I, Raja Barizan R (2012) Status of mangroves in Peninsular
Malaysia. Forest Research Institute Malaysia, Kepong.
Hamilton SE, Casey D (2016) Creation of a high spatiotemporal resolution global database of
continuous mangrove forest cover for the 21st century (CGMFC21). Glob Ecol Biogeogr 25:
729–738.
Hashim T, Suratman M, Jaafar J, Hasmadi IM, Abu F (2015) Field assessment of above ground
biomass (agb) of mangrove stand in Merbok, Malaysia. Malaysian Appl Biol 44: 81–86.
Hashim T, Suratman M, Singh H, Jaafar J, Bakar A (2020) Predictive model of mangroves carbon
stocks in Kedah, Malaysia using remote sensing. IOP Conference Series: Earth and Environ-
mental Science, 2020. IOP Publishing, 012033.
Horstman EM, Dohmen-Janssen CM, Bouma TJ, Hulscher SJMH (2015) Tidal-scale flow routing
and sedimentation in mangrove forests: Combining field data and numerical modelling. Geo-
morphology 228: 244–262.
Hutchison J, Spalding M, Ermgassen P (2014) The role of mangroves in fisheries enhancement. The
Nature Conservancy and Wetlands International: 54.
Itto/Isme (1993) Project PD71/89.Rev.1. International Society for Mangrove Ecosystem (ISME),
Okinawa, Japan.
7 Mangroves as a Carbon Sink/Stocks 173

Jachowski NRA, Quak MSY, Friess DA, Duangnamon D, Webb EL, Ziegler AD (2013) Mangrove
biomass estimation in Southwest Thailand using machine learning. Appl Geogr 45: 311–321.
Kamruzzaman M, Ahmed S, Osawa A (2017) Biomass and net primary productivity of mangrove
communities along the Oligohaline zone of Sundarbans, Bangladesh. Forest Ecosys 4: 1–9.
Kathiresan K, Bingham BL (2001) Biology of mangroves and mangrove ecosystems.
Kauffman JB, Donato DC (2012) Protocols for the measurement, monitoring and reporting of
structure, biomass, and carbon stocks in mangrove forests. Working Paper 86, CIFOR, Bogor.
Kauffman JB, Heider C, Cole TG, Dwire KA, Donato DC (2011) Ecosystem carbon stocks of
Micronesian mangrove forests. Wetlands 31: 343–352.
Komiyama A, Ong JE, Poungparn S (2008) Allometry, biomass, and productivity of mangrove
forests: A review. Aquat Bot 89: 128–137.
Krauss KW, Doyle TW, Doyle TJ, Swarzenski CM, From AS, Day RH, Conner WH (2009) Water
level observations in mangrove swamps during two hurricanes in Florida. Wetlands 29:
142–149.
Kridiborworn P, Chidthaisong A, Yuttitham M, Tripetchkul S (2012) Carbon sequestration by
mangrove forest planted specifically for charcoal production in Yeesarn, Samut Songkram. J
Sustain Energ Environ 3: 87–92.
Kumara M, Jayatissa L, Krauss K, Phillips D, Huxham M (2010) High mangrove density enhances
surface accretion, surface elevation change, and tree survival in coastal areas susceptible to
sea-level rise. Oecologia 164: 545–553.
Lacerda LDD, Diop E. (1993) Published. Conservation and sustainable utilization of mangrove
forests in Latin America and Africa regions. Workshop on Conservation and Sustainable
Utilization of Mangrove Forests in Latin America and Africa Regions., 1993. International
Society for Mangrove Ecosystems.
Le Quéré C, Andrew RM, Friedlingstein P, Sitch S, Pongratz J, Manning AC, Korsbakken JI, Peters
GP, Canadell JG, Jackson RB (2017) Global carbon budget 2017. Earth System Science Data
Discussions: 1–79.
Li S, Meng X, Ge Z, Zhang L (2015) Evaluation of the threat from sea-level rise to the mangrove
ecosystems in Tieshangang Bay, southern China. Ocean Coast Manag 109: 1–8.
Long J, Giri C, Primavera J, Trivedi M (2016) Damage and recovery assessment of the Philippines’
mangroves following Super Typhoon Haiyan. Mar Poll Bullet 109: 734–743.
Mackey A (1993) Biomass of the mangrove Avicennia marina (Forsk.) Vierh. near Brisbane, south-
eastern Queensland. Mar Freshwater Res 44: 721–725.
Mohammad Abdullah AN, Stacey N, Garnett ST, Myers B (2016) Economic dependence on
mangrove forest resources for livelihoods in the Sundarbans, Bangladesh. For Policy Econ
64: 15–24.
Muhd-Ekhzarizal M, Mohd-Hasmadi I, Hamdan O, Mohamad-Roslan M, Noor-Shaila S (2018)
Estimation of aboveground biomass in mangrove forests using vegetation indices from SPOT-5
image. J Trop For Sci 30: 224–233.
Naidoo G (2016) The mangroves of South Africa: An ecophysiological review. South Afr J Bot
107: 101–113.
Nanjo K, Kohno H, Nakamura Y, Horinouchi M, Sano M (2014) Effects of mangrove structure on
fish distribution patterns and predation risks. J Exp Mar Biol Ecol 461: 216–225.
Nicholls RJ, Marinova N, Lowe JA, Brown S, Vellinga P, De Gusmão D, Hinkel J, Tol RSJ (2011)
Sea-level rise and its possible impacts given a ‘beyond 4 C world’ in the twenty-first century.
Philos Trans A Math Phys Eng Sci 369: 161–181.
Ong JE, Gong WK (2013) Structure, function and management of mangrove ecosystems, Interna-
tional Society for Mangrove Ecosystems.
Otero V, Van De Kerchove R, Satyanarayana B, Martínez-Espinosa C, Fisol MaB, Ibrahim MRB,
Sulong I, Mohd-Lokman H, Lucas R, Dahdouh-Guebas F (2018) Managing mangrove forests
from the sky: Forest inventory using field data and Unmanned Aerial Vehicle (UAV) imagery in
the Matang Mangrove Forest Reserve, Peninsular Malaysia. For Ecol Manag 411: 35–45.
174 T. M. Z. Tengku Hashim and M. N. Suratman

Pham TD, Yoshino K, Le NN, Bui DT (2018) Estimating aboveground biomass of a mangrove
plantation on the Northern coast of Vietnam using machine learning techniques with an
integration of ALOS-2 PALSAR-2 and Sentinel-2A data. Int J Remote Sens 39: 7761–7788.
Proisy C, Mitchell A, Lucas R, Fromard F, Mougin E. (2003) Estimation of Mangrove Biomass
using Multifrequency Radar Data. Application to Mangroves of French Guiana and Northern
Australia. Proceedings of the Mangrove 2003 Conference, 2003, p. 20–24.
Putz FE, Chan HT (1986) Tree growth, dynamics, and productivity in a mature mangrove forest in
Malaysia. For Ecol Manag 17: 211–230.
Rahman AF, Dragoni D, El-Masri B (2011) Response of the Sundarbans coastline to sea level rise
and decreased sediment flow: A remote sensing assessment. Remote Sens Environ 115:
3121–3128.
Reef R, Feller IC, Lovelock CE (2010) Nutrition of mangroves. Tree Physiol 30: 1148–1160.
Richards DR, Friess DA (2016) Rates and drivers of mangrove deforestation in Southeast Asia,
2000–2012. Proc Nat Acad Sci 113: 344–349.
Romañach SS, Deangelis DL, Koh HL, Li Y, Teh SY, Raja Barizan RS, Zhai L (2018) Conserva-
tion and restoration of mangroves: Global status, perspectives, and prognosis. Ocean Coast
Manag 154: 72–82.
Ross MS, Ruiz PL, Telesnicki GJ, Meeder JF (2001) Estimating above-ground biomass and
production in mangrove communities of Biscayne National Park, Florida (USA). Wet Ecol
Manag 9: 27–37.
Sahu SK, Kathiresan K (2019) The age and species composition of mangrove forest directly
influence the net primary productivity and carbon sequestration potential. Biocatal Agric
Biotechnol 20: 101235.
Simard M, Fatoyinbo L, Smetanka C, Rivera-Monroy VH, Castañeda-Moya E, Thomas N, Van Der
Stocken T (2019) Mangrove canopy height globally related to precipitation, temperature and
cyclone frequency. Nat Geosci 12: 40–45.
Spalding M, Blasco F, Field C (1997) World mangrove Atlas.
Spalding M, Kainuma M, Collins L (2010) World Atlas of Mangroves, Taylor & Francis Group.
Suratman MN (2008) Carbon sequestration potential of mangroves in Southeast Asia. Managing
forest ecosystems: The challenge of climate change. Springer.
Tanaka K (2007) Effectiveness and limitation of the coastal vegetation for storm surge disaster
mitigation. Investigation report on the storm surge disaster by Cyclone Sidr.
Thampanya U, Vermaat J, Sinsakul S, Panapitukkul N (2006) Coastal erosion and mangrove
progradation of Southern Thailand. Estuar Coast Shelf Sci 68: 75–85.
Truong TD, Do LH (2018) Mangrove forests and aquaculture in the Mekong River Delta. Land Use
Policy 73: 20–28.
Twilley R, Chen R, Hargis T (1992) Carbon sinks in mangroves and their implications to carbon
budget of tropical coastal ecosystems. Water Air Soil Pollut 64: 265–288.
Twilley RR, Castañeda-Moya E, Rivera-Monroy VH, Rovai A (2017) Productivity and carbon
dynamics in mangrove wetlands. Mangrove ecosystems: A global biogeographic perspective.
Springer.
Venkatachalam S, Kathiresan K, Krishnamoorthy I, Narayanasamy R (2018) Survival and growth
of fish (Lates calcarifer) under integrated mangrove-aquaculture and open-aquaculture systems.
Aquac Rep 9: 18–24.
Walker W, Baccini A, Nepstad M, Horning N, Knight DE, Braun BA (2011) Field guide for forest
biomass and carbon estimation. Version 1.0. Woods Hole Research Center, Falmouth,
Massachusetts, USA.
Wang M, Cao W, Guan Q, Wu G, Wang F (2018) Assessing changes of mangrove forest in a coastal
region of southeast China using multi-temporal satellite images. Estuar Coast Shelf Sci 207:
283–292.
7 Mangroves as a Carbon Sink/Stocks 175

Ward RD, Friess DA, Day RH, Mackenzie RA (2016) Impacts of climate change on mangrove
ecosystems: A region by region overview. Ecosystem Health and Sustainability 2(4): e01211.
Wu C, Shen H, Wang K, Shen A, Deng J, Gan M (2016) Landsat imagery-based above ground
biomass estimation and change investigation related to human activities. Sustainability 8: 159.
Yusandi S, Jaya INS, Mulia F (2018) Biomass estimation model for mangrove forest using
medium-resolution imageries in BSN Co Ltd concession area, West Kalimantan. Int J Remote
Sens Earth Sci 15: 37–50.
Zhang K, Liu H, Li Y, Xu H, Shen J, Rhome J, Smith TJ (2012) The role of mangroves in
attenuating storm surges. Estuar Coast Shelf Sci 102: 11–23.
Estimation of Blue Carbon Stock
of Mangrove Ecosystem and Its Dynamics 8
in Relation to Hydrogeomorphic Settings
and Land Use-land Cover

Karuna Rao, Prabhat Ranjan, and AL. Ramanathan

Abstract

This chapter reviews blue carbon stock of different mangrove ecosystems across
globe through published literature. It also tries to evaluate its dynamics with
different land use and land cover changes. The study reveals that mangroves have
a high potential to store carbon compared to other terrestrial and coastal
ecosystems. Indian Sundarbans stores 160–360 tC/ha based on salinity and
vegetation types. The emission of carbon from the degradation of above-ground
biomass in Indian Sundarbans was 427,242 tons between 1975 and 2013. Defor-
estation of Bangladesh Sundarbans causes loss of 8500, 1800, 670, 290, 133, and
104 hectares of mangroves along Chakaria, Naf river estuary and offshore Island,
Naf river, Maiskhali Island, Jaliardwip Island, and Matabar Island, respectively.
There is a rapid decline in plantations (-58.2%), mangrove swamps (-49.3%), and
mangrove forests (-21.3%) during 2000–2017 due to their conversion to aquacul-
ture farms in Bangladesh. Along Indian coastlines, Andhra Pradesh is the most
affected area and shows significant decrease in paddy fields due to their conversion
to aquaculture farms between 1980–81 and 2000–01. It also indicates massive
increase in shrimp production from 1990-2017. Bangladesh shows a dramatic rise
in shrimp production from 56,569 to 75,274 tons from 2010-2011 to 2014-2015.
Destruction of mangroves releases carbon dioxide in the atmosphere and can
reverse mangroves’ role from a sink to source. Since aquaculture farming helps
in high revenue generation but negatively affects the coastal ecosystem, it is

K. Rao · A. Ramanathan (*)


School of Environmental Sciences, Jawaharlal Nehru University, New Delhi, India
P. Ranjan
Central Pollution Control Board, New Delhi, India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 177
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_8
178 K. Rao et al.

essential to maintain a balance. Hence Integrated Multi-Trophic Aquaculture


(IMTA) has been introduced along with mangrove forest restoration, and REDD+.

Keywords

Mangroves · Deforestation · Aquaculture · Shrimp farming · Land use land cover


change

8.1 Introduction

The carbon stored and sequestered by the coastal ecosystems from the atmosphere
and oceans in the organic-rich sediments or in the form of biomass could be termed
as “Blue Carbon.” These ecosystems include various tidal wetlands like salt
marshes, seagrasses, and mangroves. Among them, mangroves have a significant
capacity to store carbon compared to seagrasses and salt marshes. Mangroves
contain almost 1,023 Mg C/ha, and its soil has been estimated to contribute around
49-98% of total stored carbon in estuarine ecosystems (Donato et al. 2011). Due to
its high potential to store carbon, this ecosystem plays a significant role in offsetting
the increased atmospheric carbon dioxide (CO2) (Mcleod et al. 2011; Siikamäki et al.
2012), one of the main challenges at a present scenario.
Disturbances created in the mangroves and coastal systems by anthropogenic
activities, especially fossil fuel combustion, release a high amount of carbon in the
atmosphere (in the form of methane (CH4), carbon dioxide (CO2), or other species of
carbon). Other major factors include land use activities; mainly deforestation
accounts for almost 8-20% of all global greenhouse gas emissions (GHGs) (van
der Werf et al. 2009) and releases carbon stored in the living and dead biomass and
deep sediments. Loss of this vegetated ecosystem may introduce a considerable
amount of carbon dioxide in the atmosphere known as “pulse” release. It may have
the largest and most instant effect on GHGs release and is estimated to be 50 times
greater than the annual net carbon sequestration rate (Eong 1993; McLeod et al.
2011). The release of carbon dioxide is due to destabilization or exposure of the
mangrove sediments, thereby, increasing the rate of microbial activities, which in
turn increases the emission of GHGs in a significant amount to the water column or
the atmosphere (Eong 1993; Sjöling et al. 2005; Kristensen et al. 2008; Granek and
Ruttenberg 2008; Strangmann et al. 2008; Sweetman et al. 2010). Carbon emission
from mangroves is relatively unacknowledged or ignored in most climate change
mitigation policies. Various studies reveal that land clearing has reduced sediment
carbon and increased CO2 emission to a large extent, e.g. the potential emission of
carbon dioxide from the global loss of mangrove vegetation biomass and near-
surface carbon stock (a few meters from surface soil) is around 1492 MgCO2/ha
(Pendleton et al. 2012). Land clearing in Panamanian mangrove has reduced sedi-
ment carbon by 50% within eight years (Granek and Ruttenberg 2008). The mean
potential carbon dioxide emission from the degradation of above-ground biomass in
Indian Sundarban mangroves was 1567.98  551.69 Gg between 1975 and 2013
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 179

(Akhand et al. 2017). Hence factors like land clearing and deforestation and degra-
dation of these coastal ecosystems significantly impact very large pools of previ-
ously sequestered carbon.
Keeping in mind the importance of mangroves in capturing carbon, the review
from the various available literature has been carried out to characterize the sedi-
mentary carbon stock of different mangroves and its potential emission due to
several anthropogenic disturbances like land use land cover changes which include
mangrove degradation, deforestation, agriculture, and aquaculture activities. In this
study, an attempt has been made 1) to assess the carbon stock of various mangroves
across the world 2) to estimate the potential carbon loss as a result of various
anthropogenic disturbances. This assessment might support the conservation, pres-
ervation, and management of existing carbon stock in mangrove forests under
immense pressure and threat.

8.2 Significance of Mangrove in Storing Carbon Over Other


Ecosystems

8.2.1 Mangroves and Other Forest Ecosystems

Soil carbon pool within 1m depth from mangroves and other forest ecosystems like
deciduous and evergreen needle leaf forest, permanent wetlands, open shrublands,
mixed forest, grasslands open shrublands, evergreen and deciduous broadleaf forest,
croplands, savannah, and closed shrublands were studied by Sanderman et al. 2018.
Upon comparison with them, mangroves found to store the maximum amount of
carbon (361 ton/ha) than others, as shown in (Fig. 8.1). This signifies that mangrove
ecosystems are the most efficient ecosystem in storing and fixing carbon in their
sediments and biomass (Kristensen et al. 2008; Donato et al. 2011).

8.2.2 Mangroves and Tidal Marshes and Seagrasses

In a study by Mcleod et al. 2011, carbon burial rates of mangroves, tidal marshes,
and seagrasses were compared. It is found that mangroves and tidal marshes have
approximately similar carbon burial rates of 226 g C/m2/yr and 218 g C/m2/yr,
respectively, stating that they have almost the same potential to capture carbon
(Fig. 8.2). At the same time, seagrass shows 138 gC/m2/yr of burial rate. This
indicates the significance of mangroves in capturing carbon over other blue carbon
ecosystems.
In another study, in Victoria, Southeast Australia, Lewis et al. 2018 estimated the
stock in the top 30cm of sediments across the station, namely Glenelg, Corangamite,
Port Phillip Western port ecosystem, West, and East Gippsland. The lowest carbon
stock (170 ha) in Glenelg might be due to the lower organic carbon value relative to
other regions. In this study, Lewis revealed that carbon stored by mangrove ecosys-
tem was 65.64.17 Mg C/ha, and tidal marshes were 87.14.90 MgC/ha
180 K. Rao et al.

400
350
300
Blue Carbon

250
(ton/ha)

200
150
100
50
0

Ecosystem

Fig. 8.1 Soil carbon pool in 1 m depth, indicating the significance of mangrove over other forest
ecosystems in sequestering carbon (Source: Sanderman et al. 2018)

250
Carbon burial rate (gC/m2/yr)

200

150

100

50

0
Mangrove Tidal marshes Seagrass

Fig. 8.2 Carbon burial rate of various blue carbon ecosystems (Data from Mcleod et al. 2011, and
references therein)

(AverageSD) which were not significantly different. Overall, tidal marshes were
shown to have the most extensive carbon stock due to their high average value of
organic carbon and a large area covered by them (Fig. 8.3). Seagrasses showed the
lowest average organic carbon stock (24.31.82 Mg C/ha) of the three ecosystems.
Further, the study done by Mcleod et al. 2011 and Lewis et al. 2018 proposed that the
potential carbon losses lead to the potential losses of monetary values related to these
ecosystems, especially when they are vanishing at a faster rate.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 181

20000 900000
Mangrove Seagrass Tidal marsh
Mangroves & Seagrass (MgCorg)

18000 800000
16000

Tidal Marsh (MgCorg)


700000
14000
600000
12000
500000
10000
400000
8000
300000
6000
4000 200000
2000 100000
0 0
Glenelg Hopkins Corangamite Port Phillip West Gippland East Gippsland
Westernport
Ecosystem

Fig. 8.3 Sedimentary carbon stock (in the top 30 cm) across various catchments areas of Victoria,
southeast Australia (Source: Lewis et al. 2018)

8.3 Blue Carbon Sequestration

Blue carbon sequestration is around 53 million tons annually around the globe, out
of which, approximately 16 million tons (i.e., about 30%) are sequestered by
mangroves alone (Siikamäki et al. 2012). On a global scale, mangrove stores
approximately 44.6 TgC/yr with a mean sequestration value of 210 g CO2 m-2 yr-
1
. The value of the carbon stock ranged from 441.76  120.76 to 1267.00  872.72
t C ha 1 with a global average value of 78.0  64.5 t C ha 1 and carbon sequestra-
tion rate of 2.9  2.2 t C ha 1 year 1 (Murdiyarso et al. 2015; Estrada and Soares
2017). Hamilton and Friess 2018 estimated that global carbon stock in a mangrove in
2012 is around 4.190.62 Pg, out of which 2.960.53 Pg is captured in the soil and
1.230.06 Pg in standing and living biomass.
The sedimentary carbon stock across different mangrove settings in the Indone-
sian archipelago is depicted in Table 8.1. It can be observed that marine mangroves

Table 8.1 Sedimentary carbon stock across different mangrove settings in the Indonesian
archipelago
Mangroves Mg ha-1m-1 References
Seaward oriented marine mangroves 354-377 Kauffman et al. 2011
Interior marine mangroves 380-424 Kauffman et al. 2011
Landward marine mangroves 480-503 Kauffman et al. 2011
Yap 465 Donato et al. 2012
Palau 465 Donato et al. 2012
Marine mangroves 542 Murdiyarso et al. 2015
Tanjung Puting National Park 1059.2 Murdiyarso et al. 2009
Segara Anakan 571 Murdiyarso et al. 2009
Bunaken National Park 822.1 Murdiyarso et al. 2009
182 K. Rao et al.

stored more carbon than estuarine mangroves, which could be due to their
differences in organic carbon content and bulk densities in their soils. The bulk
densities and carbon content of estuarine mangroves are around 0.33–0.80 g cm-3
and 11 to 85 mg SOC g -1, and in marine mangrove soil, it is 0.18 0.27 g cm-3 and
170 to 260 mg SOC g-1, respectively. The bulk density of estuarine soil is about four
times higher than the marine soils, while the carbon concentration of marine
mangroves is up to 25 times more than the estuarine mangroves. This reveals that
marine mangroves have higher soil carbon stock despite low bulk density, making
soil organic carbon the determining factor of soil carbon stock.
Mangroves are considered to be a crucial carbon sink. When these ecosystems are
disturbed directly or indirectly, either by natural or anthropogenic means, these
stored carbons get disturbed and released in the atmosphere. Thus, they act as a
sink in natural conditions but as sources in the degraded and disturbed conditions.
Since the last few decades, various researches have been carried out to reduce or
minimize the emission of CO2 and maximize the sink capacity of mangrove forests.
Hence, they may provide an essential contribution to low-cost mitigation techniques
for climate change.

8.4 Indian Mangrove Ecosystems

In South Asian mangroves, India occupies the second largest position (3400 km2)
after Bangladesh. Total mangrove in India covers 0.15% of the country’s land, and
around 3% of the global mangrove cover; around them, 8% of global mangrove area
belongs to Asia’s mangroves (Sahu et al. 2015). The carbon pool in Indian mangrove
ecosystems and its partitioning into above-ground and below ground is represented
in (Fig. 8.4). The overall assessment reveals that maximum carbon is stored in
Sundarban, followed by Bhitarkanika, Kadalundi, Mahanadi, and Thalassery, etc.
Sundarban has higher stock as it has the most extensive coverage area than any other
mangroves. The carbon stock of Thalassery estuarine wetland was 153.64 t C ha-1
which was approximately equivalent to 536.86 t CO2 ha-1. The area covered by
Thalassery mangroves was about 5.8 ha. Thus, this mangrove’s potential carbon
sequestration is 891.11 t C, and the amount of equivalent CO2 is 3270.37 t CO2
(Vinod et al. 2019). Further, total area of Mahanadi and Bhitarkanika mangrove
(141,589+672) is considered to be 142,261 km2. The equivalent mean CO2 is
estimated to be 455.47  110.56 tones, which comes out to be approximately
64.80 TgC (Banerjee et al. 2020). The variation in carbon stock across different
mangroves might be due to differences in the extent of mangrove cover, geomorphic
settings, land use land cover patterns, different hydrodynamic conditions, etc. These
studies support the notion that mangrove forests act as an essential carbon sink and
the necessity to conserve and protect these critical ecosystems in light of climate
change mitigation.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 183

AGC SOC CO2 equivalent


900
Carbon stock (MgC/ha)

800
700
600
500
400
300
200
100
0

Fig. 8.4 Ecosystem C pools in mangrove ecosystems representing ABC, BGC, and their CO2
equivalents across various mangrove ecosystems (1. Bhomia et al. 2016; 2. Banerjee et al. 2020;
3. Rahman 2015; 4. Vinod et al. 2018; 5. Vinod et al. 2019; 6. Banerjee et al. 2020; 7. Sahu et al.
2016; 8. Sahu et al. 2016)

8.4.1 Carbon Stored in Sundarban Mangrove (Bangladesh and


India)

Sundarbans mangrove is the largest mangrove cover globally (occupying around


4,000 sq km in India and 6,017 sq km in Bangladesh). Sundarbans forest, nationally
or internationally, is of great importance for its biodiversity and environmental
services (Seidensticker and Hai 1983; Iftekhar and Saenger 2008). Hence, the
assessment of carbon sequestration in Sundarbans mangrove (both Bangladesh and
Indian part) is of immense importance. Degradation of above-ground biomass in
Indian Sundarbans mangrove leads to the potential carbon emission of about
1567.98 Gg during 1975 to 2013, which is equivalent to US$64.29 million (Akhand
et al. 2017). The carbon stored in Sundarbans mangrove forest is about 160–360 t/ha
based on vegetation types and salinity (Rahman 2015), which for Bangladesh adds
approximately 70–158 million tons of carbon stock (Table 8.2). Humanitywatch
et al. 2011 estimated the cost of US$1.87 billion in the international market for about
56 million tons of carbon captured in the Bangladesh Sundarbans.

8.5 Threats to Mangrove Ecosystems

According to Saenger 2002 and Alongi 2002, the coast of Andaman has decreased
by 79% between 1961 and 1989 due to anthropogenic activities like agriculture,
aquaculture, and various other land use and land cover changes. A similar significant
loss of mangroves has been observed in Southern Thailand due to extensive shrimp
farming between 1975 and 1993 (CORIN 1995). The destruction of mangroves
around the globe disturbs the large stock of previously sequestered carbon (956 Mg
C ha-1) at the present rate of 1% annually. It causes an additional emission of around
184 K. Rao et al.

Table 8.2 Carbon sequestration and stock in Bangladesh mangrove ecosystems


Mangroves Carbon sequestration (annually) References
Blue C sequestration 1.15-1.39 t/ha Siikamäki et al. 2012;
rate Nellemann et al. 2009
Bangladesh 0.56 million tons Chowdhury et al. 2015
Bangladesh, 56 million tons Humanitywatch et al. 2011
Sundarban
Bangladesh, 91 million tons (36 million tons AG, Chanda et al. 2016
Sundarban 55 million tons BG)
Bangladesh 70-158 million tons Ahmed et al. 2017
Sundarban C storage
Indian Sundarban C 160-360 t/ha Rahman 2015
stock
Indian Sundarban C 212.5-312.5 t/ha Donato et al. 2011
stock

4500000
4000000
3500000
3000000
Area (ha)

2500000
2000000
1500000
1000000
500000
0

Countries

1980 1990 2000 2005

Fig. 8.5 South Asian Mangroves area changes in the year 1980, 1990, 2000, and 2005 obtained
from FAO-forest resources assessments

133 Tg C yr-1 to the atmosphere. On the global scale, mangroves’ destruction


contributes to 10% of global carbon dioxide release from the deforestation (Alongi
and Mukhopadhyay 2015). The severity and extent of impacts on mangrove carbon
emission may range from small scale (harvesting of trees for fuelwood) (Malik et al.
2015) to industrial scale (timber harvesting) (Sillanpää et al. 2017). Mangrove forest
degradation for the conversion to aquacultural ponds is the largest threat to Southeast
Asian mangroves (Richards and Friess 2016). The changes in the extent of the South
East Asian mangroves area by FAO, Forest Resources Assessments in the years
1980, 1990, 2000, and 2005 have been shown in Fig. 8.5 and reveal that there is a
decrease in mangrove areas as results of various land use land cover changes.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 185

8.5.1 Degradation and Deforestation

A recent estimate by Sanderman et al. 2018 reveals that mangrove degradation leads
to the annual emission of soil carbon of around 2.0-8.1 TgCyr-1. If all the mangroves
across the world are destroyed, and it is assumed that 95% of all mangrove carbon
gets oxidized, then the loss of CO2 would be approximately 30.2 Pg CO2
equivalents, which would be equivalent to around 6.5 years of carbon emissions
from the loss of the global forests (Alongi 2018; Kennedy et al. 2014). The adverse
effect of deforestation and degradation of the mangrove ecosystem leads to the
emission of carbon and other greenhouse gas (GHGs) like methane and nitrous
oxide, which may lead to anthropogenic climate change (Ahmed et al. 2013) in the
long run, if not mitigated with time. Fig. 8.6 represents the mangrove area that
underwent deforestation in Sundarban, Bangladesh. The 8500 ha of mangroves of
Chakaria, Sundarban has been deforested for shrimp farming (Hossain et al. 2001;
Shahid and Islam 2002). In the same way, 1800, 670, 290, 133, and 104 ha of
mangroves along Naf and offshore Island, Naf river (Shahid and Islam 2002;
Hossain 2001), Maiskhali Island, Jaliardwip Island, and Matabar Island has been
destroyed for the shrimp cultivation.

8.5.2 Land Use and Land Cover Changes (LULC)

Land use and land cover changes comprise the conversion of mangrove forest lands
to aquaculture, agriculture, upstream dams, forest cutting, exploitation, industrial
use, dredging urban development, etc. (Short and Wyllie-Echeverria 1996; Valiela
et al. 2001; Duke et al. 2007; Giri et al. 2008; Waycott et al. 2009; McLeod et al.
2011). LULC result in a loss of around 25-50% of the total global area since the last
50-100 years (McLeod et al. 2011). Loss of forested land continues and even

9000
Deforested are (ha)

8000
7000
6000
5000
4000
3000
2000
1000
0
Chakaria Naf river Naf River Maiskhali Jaliardwip Matabar
etuary and Island Island
offshore
islands
Mangrove deforested due to shrimp cultivation

Fig. 8.6 Area of mangrove deforested as a result of shrimp cultivation in Sundarbans, Bangladesh
mangroves (Ahmed et al. 2017)
186 K. Rao et al.

2000 2010 2017 % Change


80

60

40

20
% Changes

-20

-40

-60
% Decrease due to land conversion, increase
-80 in population, urbanization and settlement

Fig. 8.7 Land use and land cover changes from 2000 to 2017 in the mangrove forest of Sundarban
(Source: Thakur et al. 2020)

increasing in recent days depending upon the type of forest and ecosystems,
resulting in the loss of approximately 8000 km2 land each year (Valiela et al.
2001; Alongi 2002; Duarte et al. 2004; Bridgham et al. 2006; FAO 2016; Duarte
et al. 2008; Spalding 2010; Mcleod et al. 2011). If land use and land cover change
persist at the same rate, then the estimated loss of tidal marsh and seagrass ecosystem
would be around 30-40% (IPCC 2007) and 100% of the mangroves in next
100 years.
Thakur et al. 2020, in his study in Sundarban mangroves, identified eleven LULC
classes and detected various negative and positive changes as they have been
exposed to multiple developmental activities. Different classes include Mangrove
forest, mangrove swamp, Mudflats, river, sand beach, open scrubs, settlements,
plantations, aquaculture, waterlogged areas, and agricultural land. Among all these
classes, the percentage of land cover increases in mudflats (21.3%), sand beaches
(72.5%), and the waterlogged regions (52.7%) during these 17 years (Fig. 8.7). On
the other hand, decline in the land cover of mangrove forests (-21.3%), plantations
(-58.2%), and mangrove swamps (-49.3%) has been observed due to the unrestricted
destruction of a vast cover of open mangrove forests to meet the demand of land to
accommodate the continuously rising population (Dutta et al. 2014). Some parts of
the mangrove forests were also converted to mudflats, waterlogged areas, aquacul-
ture, and agricultural ponds, and human settlements. This highlights that mangrove
swamps, mangrove forests, and plantation decreased due to forest degradation,
growing urbanization, land conversion, increase in settlements, etc. This indicates
that human activities affected the mangroves on a large scale in a very short period
while natural factors (like a cyclone, coastal flooding, rainfall, and temperature)
affect them on a small level and with a slow pace (Banerjee et al. 2012; Ghosh et al.
2015). Thus, these mangroves act as carbon storehouses under natural conditions,
but rapid human activities may shift it from sink to source of carbon.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 187

8.5.3 Aquaculture

Aquaculture is a near-shore and on-shore saltwater culture and brackish water fish,
including shellfish. It has increased over the last three decades, with a mean rate of
8.6% annually. In 2012, the production of fish touched 66.6 million tons globally,
and the contribution of Asia in this production was 88.4%. There is a rapid
expansion in aquaculture industries between the 1980s and1990s on a global scale.
Conversion of mangroves to aquaculture leads to the removal of 60% of the soil
organic carbon and 85% of the living biomass carbon stock. A similar observation is
seen in the Dominican Republic (Kauffman et al. 2014). Since the last 50 years,
approximately one-third of mangroves in the whole world have been lost due to the
conversion of the mangrove land to the aquacultural and agricultural fields (Alongi
2002). Conversion of mangrove areas at such high rates may lead to the global
carbon emission of about 0.12 Pg C/yr, which is around 10% of the total carbon
emission from deforestation (Donato et al. 2011). In Asia, the rank of countries in
aquaculture production follows an order of China, Indonesia, India, Vietnam, the
Philippines, and Bangladesh (FAO 2016). Bangladesh is one of the leading aquacul-
ture industries in the world, with around 2.06 million tons of production in 2014-
2015 (FRSS 2016). Table 8.3 represents the loss of global mangroves due to the
increase in coastal aquaculture activities, which accounts for the economic loss of
3.78-17.01 billion.
In India, Andhra Pradesh is the most affected coastline in terms of aquaculture.
Significant changes in the distribution of land use patterns have been observed
between 1980 and 81 and 2000-01. The area occupied by aquaculture is around
40.11%, 38.16%, 29.82%, 28.81%, 21.56%, 5.86%, and 29.58% in Bhimavaram
Mandal, Kalla Mandal, Akiveedu Mandal, Mogalthuru Mandal, Narasapuram
Mandal, and Palakol Mandal, respectively, in West Godavari district, Andhra
Pradesh. The land under paddy shows decreases in percentage from 78.21% to
38.71% in Bhimavaram Mandal, 77.52% to 51.89% in Akiveedu Mandal, 73.32%
to 43.27% in Kalla Mandal, 60.05% to 37.25% in Mogalthuru Mandal, 70.28% to
64.62% in Palakol Mandal, 62.28% to 40.62% in Narasapuram Mandal (Dorababu
2013). The decrease in paddy land is because of the conversion of these paddy lands
to aquaculture during 1980-81 and 2000-01 (Fig. 8.8).

Table 8.3 Loss of global mangrove forests through coastal aquaculture


Features Information Reference
Mangroves area lost to aquaculture (ha) (other than 0.49 million Valiela et al. 2001
shrimp)
Total mangrove area lost to coastal aquaculture (ha) 1.89 million Valiela et al. 2001
Mangrove area lost to coastal aquaculture (%) 52 Valiela et al. 2001
The total economic value of mangrove loss to 3.78-17.01 Ahmed and Glaser
aquaculture (US$/yr) billion 2016
188 K. Rao et al.

Aquaculture Paddy Orchards Waste Land


90
Change of Land Use (%)

80
70
60
50
40
30
20
10
0
1980-1981

2000-2001

1980-1981

2000-2001

1980-1981

2000-2001

1980-1981

2000-2001

1980-1981

2000-2001

1980-1981

2000-2001
Bhimavaram Kalla Madalam Akiveedu Palakol Mogalthuru Narasapuram
Madalam Madalam Mandalam Mandalam

Fig. 8.8 Distribution of Land Use during 1980-81 and 2000-01 in West Godavari, Andhra Pradesh
(Source: Dorababu 2013)

8.5.4 Shrimp Farming

Currently, shrimp farming is among the crucial sectors of the country’s economy.
Global mangrove loss resulting from shrimp farming accounted for 1.4 million ha
(38%) (Valiela et al. 2001). In India, the duration between 1993-2000 was the golden
period for shrimp production, and during this time, India became the fifth-largest
country of global shrimp production (Briggs et al. 2004). As a result, many shrimp
farms took over a huge part of Indian coastlines. Initially, these industries were
supported by various government agencies (CIBA 2009), but later on, neither the
central government nor state government took an interest in developing uniform
practices needed to run the industries for a long run (Jong 1989; Jana and Jana 2003;
Puthucherril 2016). Later on, various new advanced technological developments
were adopted on the international level in Thailand, Taiwan, and China, etc.
(Kongkeo 1997) and India also benefitted from this advancement, and its business
became flourishes along many Indian coastlines. Among these coastlines, Andhra
Pradesh was at the forefront, followed by West Bengal, Kerala, Orissa, Karnataka,
Maharashtra, Tamil Nadu, Gujarat, and Goa. Fig. 8.9 represents an overview of the
massive increase in the area of shrimp farming from the 1990s until 2017 in all the
mangroves across Indian coastlines.
Bangladesh has also become a multimillion-dollar industry in shrimp farming in
global markets, particularly the United States of America (USA) and the European
Union (EU). In 2014-15, Bangladesh knew to export 44,278t of prawn, and shrimp
valued US$ 142million and US$364 million, respectively (FRSS 2016). Due to their
high export value, shrimp has been referred to as “white gold” in Bangladesh
(Ahmed et al. 2013). In Bangladesh, shrimp production has dramatically increased
from 56,569 t to 75,274 t from 2010-2011 to 2014-2015.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 189

1990-1995 1996-2000 2001-2005 2006-2010 2011-2015 2017#


90000
State-wise area under shrimp culture in India

80000
70000
60000
50000
40000
30000
(ha)

20000
10000
0

#Not applicable five years interval data.

Fig. 8.9 State-wise massive increase in the area of shrimp farming from the 1990s until 2017 in
India (Source: MPEDA 2018)

8.6 Emission of the Carbon Dioxide as Significant Greenhouse


Gas (GHGs) Concern

The concentration of CO2 is increasing continuously at a higher pace. Its value was
278 ppm in 1750, which rises to 390 ppm in 2011, 407.4 ppm in 2018 and would
likely to increase between 467 and 555 ppm by the year 2050 (Table 8.4) (IPCC
2007; Anderson et al. 2009; Stocker et al. 2013; Lindsey 2018) and can increase the
temperature which could melt the polar ice and sea level may rise to 5m (Detwiler
and Hall 1988; IUFRO 2009). Among GHGs, CO2 comprises 77% of its composi-
tion (Devi et al. 2012), and hence it is crucial to study its emission and sink.

8.6.1 Carbon Stock and Related Potential Carbon dioxide Emission


of Various Mangroves Across the Globe

Bhomia et al. 2016 studied the top 30 cm of mangrove soil and revealed that the total
amount of equivalent carbon dioxide (CO2e) releases when soil carbon is oxidized
would be 21.0  105 to 23.2  105 Mg CO2e for an area of 155–183 km2 in
Bhitarkanika mangroves. This is comparable to the amount of CO2 emitted by fossil

Table 8.4 Concentration Year CO2 concentration


of carbon dioxide from
1750 278 ppm
1750 to 2018
2011 390 ppm
2018 407.4
2050 467-555 ppm
190 K. Rao et al.

Carbon Stock (tC/ha) CO2 equivalent

3500
Carbon Stock (tC/ha) and CO 2

3000
2500
2000
equivalent

1500
1000
500
0

Fig. 8.10 The carbon stock and equivalent CO2 emission of many mangroves across the world. (1.
Pandey and Pandey 2013; 2. Ray et al. 2011; 3. Suresh et al. 2013; 4. ISFR 2019; 5. Kathiresan et al.
2013; 6. Mall 1991; 7. Khan et al. 2007; 8. Bosire et al. 2012; 9. Murdiyarso et al. 2010; 10. Adame
et al. 2013; 11. Jones et al. 2014; 12. Kauffman et al. 2011; 13. Kauffman and Donato 2012; 14.
Kusumaningtyas et al. 2019; 15. Kusumaningtyas et al. 2019; 16. Kusumaningtyas et al. 2019; 17.
Kusumaningtyas et al. 2019)

fuel (oil) combustion in Nepal in 2005 (IEA 2013). Similarly, various carbon stocks
have been compared across the world. CO2 equivalent has been calculated by
multiplying the carbon stock by 3.67 as one ton of carbon is equal to 3.67 tons of
CO2 and represents an equivalent amount of carbon lost from long-lived pool such as
sediments. Fig. 8.10 depicts the comparison of carbon stock and equivalent CO2
across many mangroves around the world. The maximum carbon stored and maxi-
mum equivalent CO2 emission is observed in Yap (Micronesia), Sian Ka’an bio-
sphere reserve (Mexico), Berau (Indonesia), Zambezi Delta (Central Mozambique),
West Bengal (India), Kalimantan (Indonesia), and Eastern Segara Anakan Lagoon
(SAL) (Indonesia). This suggests that since mangrove has a high capacity to
sequester carbon, it has the same ability to emit carbon in the atmosphere, if it gets
disturbed.

8.6.2 Estimated Carbon dioxide Emission Based on Ecosystem Loss


Since European Settlement

A similar observation has been reflected in Victorian coastlines (Southeast Australia)


where Lewis et al. 2018 represent that its total mean carbon stock is 4970 Mg C in
the top 30 cm of soils and estimated the potential losses of carbon stock since
European Settlements. The estimates of carbon loss are based on 50-90%
remineralization of organic carbon. Hence, they may underestimate the loss of actual
organic carbon deep down a meter or more (Pendleton et al. 2012). The total carbon
stock in the Inlets, French Island, Nooramunga Coast, Western Port, Shallow Inlet,
and Corner Inlet is 114, 328, 1094, 7256, 9054, and 11,972 MgC, respectively. The
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 191

Total stocks to 30 cm depth (MgC) 50% loss 90% loss


45000
crabon stock and emission

40000
35000
30000
25000
20000
15000
10000
5000
0
The Inlets French Island Nooramunga Western Port Shallow Inlet Corner Inlet
Coast

Fig. 8.11 Estimated Sediment carbon emissions based on ecosystem loss since European Settle-
ment along Victorian coastlines (Lewis et al. 2018)

carbon loss from 50% remineralization in the Inlets, French Island, Nooramunga
Coast, Western Port, Shallow Inlet, and Corner Inlet is 210, 602, 2008, 13314,
16614, and 21968 MgC, respectively, and from 90% mineralization is 378, 1083,
3614, 23996, 29905 and 39543 MgC, respectively (Fig. 8.11). The largest estimated
carbon stock losses could be observed in Western Port, Shallow Inlet, Corner Inlet,
totalling to approximately 95% of the estimated stock.

8.6.3 Carbon dioxide Emission from Mangrove Conversion


to Shrimp Farming

The rapid increase in aquaculture and shrimp farming industries has adversely
affected the mangroves of several countries, including China, India, Brazil, Mexico,
Bangladesh, Myanmar, Indonesia, Sri Lanka, Thailand, Vietnam, and the
Philippines (FAO 2007; UNEP 2014). Various adverse environmental impacts
have been overlooked over substantial economic benefits. In Puttalam Lagoon, Sri
Lanka, the mangroves conversion to shrimp ponds might cause a loss of 1,91,584
t of total carbon between 1990 and 2012, which is about 75.5% of the total carbon
loss (Bournazel et al. 2015). In Ecuador mangroves, about 80% of the living carbon
lost (7.01 million t) resulted from the replacement of the mangroves by shrimp
cultivation (Hamilton and Lovette 2015)
Kauffman et al. 2014 estimated that about 1,036,971 Mg C is lost due to the
mangrove conversion to cultivated shrimp ponds, salt ponds, and other uses. Donato
et al. 2011 estimated that mangrove stores about 20 PgC globally, which is about 2.5
times of annual carbon dioxide annually. Hence, land use and land cover type have a
significant effect on total carbon storage and sequestration rates in the coastal
ecosystem (Guo and Gifford 2002). Figure 8.12 represents the potential carbon
192 K. Rao et al.

4000
Potential CO2 emisson (Mg CO2 ha-1)

3500
3000
2500
2000
1500
1000
500
0
Tropical Tropical Global Indo-Pacific Indo-Pacific Thailand Dominican Sundarban
evergreen dry forest of mangrove region (4) region (5) Mangrove Republic (to Bangladesh
forest of Mexico (to cover (3) (to shrimp shrimp (estimate
Amazon (to pasture farm) (6) farm) (7) for next
pasture land) (2) century) (8)
land) (1)
Ecosystem

Fig. 8.12 Potential CO2 emission from mangrove conversion to different land use land cover
pattern. (1 Kauffman et al. 2003; 2. Kauffman et al. 2009; 3. Pendleton et al. 2012; 4. Donato et al.
2011; 5. Donato et al. 2011; 6. Yee 2010; 7. Kauffman et al. 2014; 8. Chanda et al. 2016)

dioxide loss due to the conversion of mangroves to shrimp farming across the globe,
where the Dominican Republic shows the maximum carbon dioxide per hectare.

8.7 Adaptation Strategies to Balance the Land Use Land Cover


and Ecosystem

It is difficult to control the spreading of aquaculture farming as a large population


depends on them for their livelihood, especially the poor inhabitants living across the
coastlines. The productivity of mangroves decreases following 5-10 years of con-
version of mangroves to aquaculture and when the ponds are abandoned (Bosma
et al. 2012; Cameron et al. 2019). Some of the adaptation strategies have been taken
into account to run the aquaculture industries without much affecting the ecology
and environment of coastal ecosystems. Some of the strategies are as follows:

8.7.1 IMTA (Integrated Multi-Trophic Aquaculture)

This is the procedure of producing diverse species of shellfish and finfish in


integrated farming with the seaweeds from various trophic levels. IMTA increases
profitability and productivity via recycling and reusing of nutrients. IMTA is based
on the principle of co-cultivation of different organic and inorganic extractive
species and fed fish (Troell et al. 2009; Chopin et al. 2010; Chopin et al. 2012).
This is also called as “greening of aquaculture.”
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 193

Benefits of IMTA-

1) Environmental friendly.
2) Hinder intensive fishing.
3) It could minimize the emission of blue carbon, thereby enhancing sequestration
and storage of blue carbon.
4) It increases the biodiversity (Clements and Chopin 2017) and resilience of the
coastal ecosystems (Worm et al. 2006; Levin and Lubchenco 2008).
5) IMTA may not be disturbed by natural calamities like sea-level rise, flood, and an
increase in water temperature.
6) Can acclimatize to a wide range of water salinity as molluscs, seaweeds, and
shrimp can tolerate variation in salinity.
7) Cultivation of seaweed in IMTA could play a crucial role in the sequestration of
blue carbon via photosynthesis (Chung et al. 2013).
8) Seaweeds in IMTA keep the water cool and clean due to the absorption of toxic
materials, pollutants, and sediments (Chung et al. 2013).

8.7.2 Restoration of Mangroves

Mangroves restoration would reimburse for mangrove loss by various aquaculture


and shrimp farming. Regeneration of mangrove includes plantations. There are two
main types of mangrove restoration: 1) mangrove restoration following natural
disturbances and 2) mangrove restoration following anthropogenic degradation
(Biswas et al. 2009).

8.7.3 REDD+ (Reducing Emissions from Deforestation and Forest


Degradation)

The REDD+ program facilitates the afforestation, reforestation, and mangroves


restoration and thus is wholly invested in the conservation of mangroves (UNEP
2014; Beymer-Farris and Bassett 2012; Olander et al. 2012). The REDD+ method is
appropriate for mitigating the emission of greenhouse gas. It plays a significant role
in minimizing anthropogenic emissions, thus helping in climate change mitigation.
REDD+ if applied worldwide, can avoid almost 2.5 billion tons of carbon dioxide
emission annually (Overmars et al. 2014). The combined approach of IMTA,
mangrove restoration, and REDD+ could bring a broad range of economic, social,
and environmental benefits.
194 K. Rao et al.

8.8 Conclusion

The concentration of carbon dioxide is increasing continuously as a result of various


anthropogenic activities. Hence, it is essential to recognize and acknowledge the
ecosystems which can store carbon. The ability of coastal ecosystems like
seagrasses, tidal marshes, and mangroves to store carbon is massive. This study
assessed the various carbon stocks in the mangroves across the globe and the effect
of anthropogenic disturbances on them. Indonesia has maximum stored carbon
(especially in its marine settings) among Asia. In India, Sundarbans shows a
significant amount of stored carbon. The disturbances include various land use
land cover changes which comprise forest degradation and deforestation, aquacul-
ture, and shrimp farming. This study reveals that mangrove’s forest land shows a
significant decrease in their area due to their conversion to aquaculture and shrimp
farming. Aquaculture and shrimp farming are the fastest growing food industry in
the world and show million-dollar profits in the global market. Bangladesh coast (In
Bangladesh) and the Andhra Pradesh coast (in India) are the most affecting the
coastal ecosystem in terms of shrimp farming and aquaculture.
On the one hand, these industries give a boost to the country’s economy; on the
other hand, it adversely affects the mangroves ecosystem by releasing previously
sequestered carbon in the form of carbon dioxide in the atmosphere. This reveals that
anthropogenic disturbances can turn the mangroves from carbon sink to source.
Since aquaculture farming helps in high revenue generation but negatively affects
the coastal ecosystem, it is essential to maintain a balance. Hence greening of
aquaculture has been introduced (like IMTA) along with mangrove forest restora-
tion, and REDD+.

References
Adame MF, Kauffman JB, Medina I, Gamboa JN, Torres O, Caamal JP, Reza M, Herrera-Silveira
JA (2013) Carbon stocks of tropical coastal wetlands within the karstic landscape of the
Mexican Caribbean. PloS one 8(2): p.e56569.
Ahmed N, Glaser M (2016) Coastal aquaculture, mangrove deforestation and blue carbon
emissions: is REDD+ a solution?. Mar Policy 66:58-66.
Ahmed N, Cheung WW, Thompson S, Glaser M (2017) Solutions to blue carbon emissions: Shrimp
cultivation, mangrove deforestation and climate change in coastal Bangladesh. Mar Policy
82:68-75.
Ahmed N, Occhipinti-Ambrogi A, Muir JF (2013) The impact of climate change on prawn
postlarvae fishing in coastal Bangladesh: socioeconomic and ecological perspectives. Mar
Policy 39:224-233.
Akhand A, Mukhopadhyay A, Chanda A, Mukherjee S, Das A, Das S, Hazra S, Mitra D,
Choudhury SB, Rao KH (2017) Potential CO2 emission due to loss of above ground biomass
from the Indian Sundarban mangroves during the last four decades. J Indian Soc Remote Sens
45(1):147-154.
Alongi DM (2002) Present state and future of the world’s mangrove forests. Environ Conserv 29
(3):331-349.
Alongi DM, Mukhopadhyay SK (2015) Contribution of mangroves to coastal carbon cycling in low
latitude seas. Agric For Meteorol 213:266-272.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 195

Alongi DM (2018) Mangrove forests. In Blue Carbon (pp. 23-36). Springer, Cham.
Anderson K, Starkey R, Bows A (2009) Defining dangerous climate change - a call for consistency.
Policy 14(2):103-110.
Banerjee K, Roy Chowdhury M, Sengupta K, Sett S, Mitra A (2012) Influence of anthropogenic
and natural factors on the mangrove soil of Indian Sundarbans wetland. Arch Environ Sci 6:80-
91.
Banerjee K, Sahoo CK, Bal G, Mallik K, Paul R, Mitra A (2020) High blue carbon stock in
mangrove forests of Eastern India. Trop Ecol 61:150-167.
Beymer-Farris BA, Bassett TJ (2012) The REDD menace: Resurgent protectionism in Tanzania’s
mangrove forests. Glob Environ Change 22(2):332-341.
Bhomia RK, MacKenzie RA, Murdiyarso D, Sasmito SD, Purbopuspito J (2016) Impacts of land
use on Indian mangrove forest carbon stocks: Implications for conservation and management.
Ecol Appl 26(5):1396-1408.
Biswas SR, Mallik AU, Choudhury JK, Nishat A (2009) A unified framework for the restoration of
Southeast Asian mangroves - bridging ecology, society and economics. Wet Ecol Manage17
(4):365-383.
Bosma R, Sidik AS, van Zwieten P, Aditya A, Visser L (2012) Challenges of a transition to a
sustainably managed shrimp culture agro-ecosystem in the Mahakam delta, East Kalimantan,
Indonesia. Wet Ecol Manag 20(2):89-99.
Bosire JO, Bandeira S, Rafael J (2012) Coastal climate change mitigation and adaptation through
REDD+ carbon programs in mangroves in Mozambique: Pilot in the Zambezi Delta. Determi-
nation of carbon stocks through localized allometric equations component, WWF.
Bournazel J, Kumara, MP, Jayatissa LP, Viergever K, Morel V Huxham M (2015) The impacts of
shrimp farming on land-use and carbon storage around Puttalam lagoon, Sri Lanka. Ocean Coast
Manage 113:18-28.
Bridgham SD, Megonigal JP, Keller JK, Bliss NB, Trettin C (2006) The carbon balance of North
American wetlands. Wetlands 26(4):889-916.
Briggs M, Funge-Smith S, Subasinghe R, Phillips M. Introductions and movement of Penaeus
vannamei and Penaeus stylirostris in Asia and the Pacific. RAP publication, 10(2004):92.
Cameron C, Hutley LB, Friess DA, Brown B (2019) High greenhouse gas emissions mitigation
benefits from mangrove rehabilitation in Sulawesi, Indonesia. Eco Ser 40:101035.
Chanda A, Mukhopadhyay A, Ghosh T, Akhand A, Mondal P, Ghosh S, Mukherjee S, Wolf J,
Lázár AN, Rahman MM, Salehin M (2016) Blue carbon stock of the Bangladesh Sundarban
mangroves: what could be the scenario after a century?. Wetlands 36(6):1033-1045.
Chopin T, Troell M, Reid GK, Knowler D, Robinson SMC, Neori A, Buschmann AH, Pang SJ,
Fang J (2010) Integrated multi-trophic aquaculture (IMTA)-A responsible practice providing
diversified seafood products while rendering biomitigating services through its extractive
components. Abstracts. Aquaculture Europe.
Chopin T, Cooper JA, Reid G, Cross S, Moore, C (2012) Open-water integrated multi-trophic
aquaculture: environmental biomitigation and economic diversification of fed aquaculture by
extractive aquaculture. Rev Aquac 4(4):209-220.
CIBA (2009) Training manual on Better management practices in shrimp farming. CIBA Special
Publications No.44. Madras, India, p. 133.
Chowdhury SR, Hossain MS, Sharifuzzaman SM, Sarker S (2015) Blue carbon in the coastal
ecosystems of Bangladesh. Project Document, Support to Bangladesh on Climate Change
Negotiation and Knowledge Management on Various Streams of UNFCCC Process Project,
funded by DFID and Danida, implemented by IUCN Bangladesh Country Office.
Chung IK, Oak JH, Lee JA, Shin JA, Kim JG, Park KS (2013) Installing kelp forests/seaweed beds
for mitigation and adaptation against global warming: Korean Project Overview. ICES J Mar Sci
70(5):1038-1044.
Clements JC, Chopin T (2017) Ocean acidification and marine aquaculture in North America:
potential impacts and mitigation strategies. Rev Aqu 9(4):326-341.
196 K. Rao et al.

Coastal Resource Institute (CORIN) (1995) The effect of aquaculture on agricultural land and
coastal environment. Mimeo, Prince of Songkla University, Songkhla, Thailand.
Detwiler RP, Hall CAS (1988) Tropical forests and the global carbon cycle. Science 239:42–47.
Devi B, Bhardwaj DR, Panwar P, Pal S, Gupta NK, Thakur CL (2012) Carbon allocation,
sequestration and carbon dioxide mitigation under plantation forests of north western Himalaya,
India. Ann Forest Res 56(1):123–135.
Donato DC, Kauffman JB, Murdiyarso D, Kurnianto S, Stidham M, Kanninen M (2011)
Mangroves among the most carbon-rich forests in the tropics. Nat Geosci 4(5):293-297.
Donato DC, Kauffman JB, Mackenzie RA, Ainsworth A, Pfleeger AZ (2012) Whole-island carbon
stocks in the tropical Pacific: Implications for mangrove conservation and upland restoration. J
Environ Manage 97:89–96.
Dorababu KK (2013) Impact of aquaculture on land use patterns, environment and economy: A
case study of west Godavari district, Andhra Pradesh, India. Int J Curr Res 5(7):1993–1996
Duarte CM, Middelburg JJ, Caraco N (2004) Major role of marine vegetation on the oceanic carbon
cycle. Biogeosciences 2:1–8
Duarte CM, Dennison WC, Orth RJ, Carruthers TJ (2008) The charisma of coastal ecosystems:
addressing the imbalance. Estuar Coast 31(2):233-238.
Duke NC, Meynecke JO, Dittmann S, Ellison AM, Anger K, Berger U, Cannicci S, Diele K, Ewel
KC, Field CD, Koedam N (2007) A world without mangroves?. Science 317(5834):41–42.
Dutta D, Das PK, Paul S, Sharma JR Dhadwal VK (2014) Spatio-temporal assessment of ecological
disturbance and its intensity in the mangrove forest using MODIS derived disturbance index.
International Archives of the Photogrammetry, Remote Sensing and Spatial Information
Sciences 8:555–559.
Eong OJ (1993) Mangroves-a carbon source and sink. Chemosphere 27(6):1097–1107.
Estrada GC, Soares ML (2017) Global patterns of aboveground carbon stock and sequestration in
mangroves. An Acad Bras Ciênc 89(2):973–989.
FAO, The State of World Fisheries and Aquaculture (2016) Contributing to Food Security and
Nutrition for All, Food and Agriculture Organization of the United Nations, Rome.
FAO, Food and Agriculture Organization (FAO) of the United Nations (2007) The World’s
Mangroves 1980–2005. FAO, Rome.
FRSS, Fisheries Statistical Report of Bangladesh, Fisheries Resources Survey System, Department
of Fisheries, Bangladesh, 2016.
Ghosh A, Schmidt S, Fickert T, Nüsser M (2015) The Indian Sundarban mangrove forests: history,
utilization, conservation strategies and local perception. Diversity 7(2):149-169.
Giri C, Zhu Z, Tieszen LL, Singh A, Gillette S, Kelmelis JA (2008) Mangrove forest distributions
and dynamics (1975–2005) of the tsunami-affected region of Asia. J Biogeograph 35
(3):519–528.
Granek E, Ruttenberg BI (2008) Changes in biotic and abiotic processes following mangrove
clearing. Estuar Coast Mar Sci 80(4):555-562.
Guo LB, Gifford RM (2002) Soil carbon stocks and land use change: a meta-analysis. Glob Change
Biol 8(4):345-360.
Hamilton SE, Friess DA (2018) Global carbon stocks and potential emissions due to mangrove
deforestation from 2000 to 2012. Nat Clim Change 8(3):240-244.
Hamilton SE, Lovette J (2015) Ecuador’s mangrove forest carbon stocks: A spatiotemporal analysis
of living carbon holdings and their depletion since the advent of commercial aquaculture. PloS
one 10(3):e0118880.
Hossain M, Lin CK, Hussain MZ (2001) Goodbye Chakaria Sundarban: the oldest mangrove forest.
Wet Sci Prac18(3):19-22.
Hossain MS (2001) Biological aspects of the coastal and marine environment of Bangladesh. Ocean
Coast Manage 44(3-4):261-282.
Humanity watch Carbon Trading, the Sundarbans and Climate Justice, Campaign Paper, Equity and
Justice Working Group-Bangladesh, Humanitywatch, 2011.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 197

IEA (International Energy Agency) (2013) Definition. Paris, France: Total Primary Energy Supply.
http://www.iea.org/stats/defs/Tpes.asp; accessed on 24 June 2013.
Iftekhar MS, Saenger P (2008) Vegetation dynamics in the Bangladesh Sundarbans mangroves: a
review of forest inventories. Wet Ecol Manage 16(4):291-312.
India State of Forest Report (ISFR) (2019).
Intergovernmental Panel on Climate Change (IPCC) (2007) Climate change 2007: Assessment
Report. IPCC, Valencia.
IUFRO (2009) Adaptation of forests and people to climate change. A global assessment report.
IUFRO World Series 22:224
Jana BB, Jana S (2003) The potential and sustainability of aquaculture in India. J Appl Aquacul 13
(3-4):283-316.
Jones TG, Ratsimba HR, Ravaoarinorotsihoarana L, Cripps G, Bey A (2014) Ecological variability
and carbon stock estimates of mangrove ecosystems in north-western Madagascar. Forests 5
(1):177-205.
Jong JD (1989) Aquaculture in India. Rijksdienst Voor Ondernemend. 2:1117–1152.
Levin SA, Lubchenco J (2008) Resilience, robustness, and marine ecosystem-based management.
Bioscience 58(1):27-32.
Lindsey R (2018) Climate change: atmospheric carbon dioxide. National Oceanographic and
Atmospheric Administration, News & Features. August.
Kathiresan K, Anburaj R, Gomathi V, Saravanakumar K (2013) Carbon sequestration potential of
Rhizophora mucronata and Avicennia marina as influenced by age, season, growth and sedi-
ment characteristics in southeast coast of India. J Coast Conserv 17(3):397-408.
Kauffman JB, Donato DC (2012) Protocols for the measurement, monitoring and reporting of
structure, biomass, and carbon stocks in mangrove forests (pp. 50-p). Bogor, Indonesia: CIFOR.
Kauffman JB, Heider C, Cole TG, Dwire KA, Donato DC (2011) Ecosystem carbon stocks of
Micronesian mangrove forests. Wetlands 31(2):343-352.
Kauffman JB, Heider C, Norfolk J, Payton F (2014) Carbon stocks of intact mangroves and carbon
emissions arising from their conversion in the Dominican Republic. Ecol Appl 24(3):518-527.
Kauffman JB, Hughes RF, Heider C (2009) Dynamics of C and nutrient pools associated with land
conversion and abandonment in Neotropical landscapes. Ecol Appl 19:1211-1222.
Kauffman JB, Steele MD, Cummings DL, Jaramillo VJ (2003) Biomass dynamics associated with
deforestation, fire, and, conversion to cattle pasture in a Mexican tropical dry forest. Forest Ecol
Manage 176(1-3):1-12.
Kennedy H. et al., (2014) In Supplement to the 2006 IPCC Guidelines for National Greenhouse Gas
Inventories: Wetlands (eds Hiraishi, T. et al.), Intergovernmental Panel on Climate Change,
Gland, Switzerland.
Khan MNI, Suwa R, Hagihara A (2007) Carbon and nitrogen pools in a mangrove stand of Kandelia
obovata (S., L.) Yong: vertical distribution in the soil–vegetation system. Wet Ecol Manage 15
(2):141-153.
Kongkeo H (1997) Comparison of intensive shrimp farming systems in Indonesia, Philippines,
Taiwan and Thailand. Aquacul Res 28(10):789-796.
Kristensen E, Bouillon S, Dittmar T, Marchand C (2008) Organic carbon dynamics in mangrove
ecosystems: a review. Aquat Bot 89(2):201-219.
Kusumaningtyas MA, Hutahaean AA, Fischer HW, Pérez-Mayo M, Ransby D, Jennerjahn TC
(2019) Variability in the organic carbon stocks, sources, and accumulation rates of Indonesian
mangrove ecosystems. Estuar Coast Mar Sci 218:310-323.
Lewis CJE, Carnell PE, Sanderman J, Baldock JA, Macreadie PI (2018) Variability and vulnerabil-
ity of coastal ‘blue carbon’stocks: a case study from southeast Australia. Ecosystems 21(2):263-
279.
Malik A, Fensholt R, Mertz O (2015) Mangrove exploitation effects on biodiversity and ecosystem
services. Biodivers Conserv 24(14):3543-3557.
Mall LP (1991) Study of biomass, litter fall, litter decomposition and soil respiration in monogenic
mangrove and mixed mangrove forests of Andaman Islands. Trop Ecol 32:144-152.
198 K. Rao et al.

Mcleod E, Chmura GL, Bouillon S, Salm R, Björk M, Duarte CM, Lovelock CE, Schlesinger WH,
Silliman BR (2011) A blueprint for blue carbon: toward an improved understanding of the role
of vegetated coastal habitats in sequestering CO2. Front Ecol Environ 9(10):552-560.
MPEDA (2018) Annual Report 2017–18. The Marine Products Export Development Authority.
Cochin, India, p. 59.
Murdiyarso D, Donato D, Kauffman JB, Kurnianto S, Stidham M, Kanninen M (2009) Carbon
storage in mangrove and peatland ecosystems: A preliminary account from plots in Indonesia.
Working paper 48. Bogor Banat, Indonesia: Center Internat Forest Res 35:1-35.
Murdiyarso D, Hergoualc’h K, Verchot LV (2010) Opportunities for reducing greenhouse gas
emissions in tropical peatlands. Proc Nat Acad Sci 107(46):19655-19660.
Murdiyarso D, Purbopuspito J, Kauffman JB, Warren MW, Sasmito SD, Donato DC, Manuri S,
Krisnawati H, Taberimadon S, Kurnianto S (2015) The potential of Indonesian mangrove
forests for global climate change mitigation. Nat Clim Change 5(12):1089-1092.
Nellemann C, Corcoran E, Duarte CM, Valdes L, DeYoung C, et al., (2009) Blue Carbon. A Rapid
Response Assessment. United Nations Environment Programme, GRID-Arendal website.
www.grida.no. Accessed 2011 Nov 11.
Olander LP, Galik CS, Kissinger GA (2012) Operationalizing REDD+: Scope of reduced emissions
from deforestation and forest degradation. Curr Opin Environ Sustain 4(6):661-669.
Overmars KP, Stehfest E, Tabeau A, van Meijl H, Beltrán AM, Kram T (2014) Estimating the
opportunity costs of reducing carbon dioxide emissions via avoided deforestation, using
integrated assessment modelling. Land Use Policy 41:45-60.
Pandey CN, Pandey R (2013) Carbon sequestration in mangroves of Gujarat, India. Internat J Bot
Res 3(2):57-70.
Pendleton L, Donato DC, Murray BC, Crooks S, Jenkins WA, Sifleet S, Craft C, Fourqurean JW,
Kauffman JB, Marbà N, Megonigal P (2012) Estimating global “blue carbon” emissions from
conversion and degradation of vegetated coastal ecosystems. PloS One 7(9):e43542.
Puthucherril TG (2016) Sustainable aquaculture in India: looking back to think ahead. In Aquacul-
ture Law and Policy. Edward Elgar Publishing.
Rahman MM (2015) Carbon and nitrogen dynamics and carbon sequestration in soils under
different residue management. The Agriculturists 12(2):48-55.
Ray R, Ganguly D, Chowdhury C, Dey M, Das S, Dutta MK, Mandal SK, Majumder N, De TK,
Mukhopadhyay SK, Jana TK (2011) Carbon sequestration and annual increase of carbon stock
in a mangrove forest. Atmos Environ 45(28):5016-5024.
Richards DR, Friess DA (2016) Rates and drivers of mangrove deforestation in Southeast Asia,
2000–2012. Proc Nat Acad Sci 113(2):344-349.
Saenger P (2002) Mangrove ecology, silviculture and conservation. Springer Science & Business
Media.
Sahu SC, Kumar M, Ravindranath NH (2016) Carbon stocks in natural and planted mangrove
forests of Mahanadi Mangrove Wetland, East Coast of India. Curr Sci 110:2253-2260.
Sahu SC, Suresh HS, Murthy IK, Ravindranath NH (2015) Mangrove area assessment in India:
implications of loss of mangroves. J Earth Sci Climat Change 6(5):1.
Sanderman J, Hengl T, Fiske G, Solvik K, Adame MF, Benson L, Bukoski JJ, Carnell P, Cifuentes-
Jara M, Donato D, Duncan C (2018) A global map of mangrove forest soil carbon at 30 m spatial
resolution. Environ Res Lett 13(5):055002.
Seidensticker J, Hai MA (1983) The Sundarbans wildlife management plan: conservation in the
Bangladesh coastal zone.
Shahid MA, Islam J (2002) March. Impact of denudation of mangrove forest due to shrimp farming
on the coastal environment in Bangladesh. In Technical proceedings of BAU-NURAD work-
shop on environment and socio-economic impacts of shrimp farming in Bangladesh 5:67-75.
Short FT, Wyllie-Echeverria S (1996) Natural and human-induced disturbance of seagrasses.
Environ Conserv 23:17-27.
Siikamäki J, Sanchirico JN, Jardine SL (2012) Global economic potential for reducing carbon
dioxide emissions from mangrove loss. Proc Nat Acad Sci 109(36):14369-14374.
8 Estimation of Blue Carbon Stock of Mangrove Ecosystem and Its Dynamics in. . . 199

Sillanpää M, Vantellingen J, Friess DA (2017) Vegetation regeneration in a sustainably harvested


mangrove forest in West Papua, Indonesia. Forest Ecol Manage 390:137-146.
Sjöling S, Mohammed SM, Lyimo TJ, Kyaruzi JJ (2005) Benthic bacterial diversity and nutrient
processes in mangroves: impact of deforestation. Estuar Coast Shelf Sci 63(3):397-406.
Spalding M (2010) World atlas of mangroves. Routledge.
Stocker TF, Qin D, Plattner GK, Tignor M, Allen SK, Boschung J, Nauels A, Xia Y, Bex V,
Midgley PM (2013) IPCC, 2013. Climate change.
Suresh B, Manjappa S, Puttaiah ET (2013) Dynamics of phytoplankton succession in Tungabhadra
river near Harihar, Karnataka (India). J Microbiol Antimicrob 5(7):65-71.
Sweetman AK, Middelburg JJ, Berle AM, Bernardino AF, Schander C, Demopoulos AWJ, Smith
CR (2010) Impacts of exotic mangrove forests and mangrove deforestation on carbon
remineralization and ecosystem functioning in marine sediments. Biogeosciences 7(7).
Strangmann A, Bashan Y, Giani L (2008) Methane in pristine and impaired mangrove soils and its
possible effect on establishment of mangrove seedlings. Biol Fert Soils 44(3):511.
Thakur S, Maity D, Mondal I, Basumatary G, Ghosh PB, Das P, De TK (2020) Assessment of
changes in land use, land cover, and land surface temperature in the mangrove forest of
Sundarbans, northeast coast of India. Environ Develop Sustain 1-27.
Troell M, Joyce A, Chopin T, Neori A, Buschmann AH, Fang JG (2009) Ecological engineering in
aquaculture-potential for integrated multi-trophic aquaculture (IMTA) in marine offshore
systems. Aquaculture 297(1-4):1-9.
UNEP (2014) The importance of mangroves to people: A call to action. Cambridge: UNEP World
Conservation Monitoring Centre.
Valiela I, Bowen JL, York JK (2001) Mangrove Forests: One of the World’s Threatened Major
Tropical Environments: At least 35% of the area of mangrove forests has been lost in the past
two decades, losses that exceed those for tropical rain forests and coral reefs, two other well-
known threatened environments. Bioscience 51(10):807-815.
Van der Werf GR, Morton DC, DeFries RS, Olivier JG, Kasibhatla PS, Jackson RB, Collatz GJ,
Randerson JT (2009) CO2 emissions from forest loss. Nat Geosci 2(11):737-738.
Vinod K, Anasu Koya A, Kunhikoya VA, Shilpa PG, Asokan PK, Zacharia PU, Joshi KK (2018)
Biomass and carbon stocks in mangrove stands of Kadalundi estuarine wetland, south-west
coast of India. Ind J Fisher 65(2):89-99.
Vinod K, Asokan PK, Zacharia PU, Ansar CP, Vijayan G, Anasukoya A, Kunhi Koya VA,
Nikhiljith M (2019) Assessment of Biomass and Carbon Stocks in Mangroves of Thalassery
Estuarine Wetland of Kerala, South West Coast of India. J Coast Res 86(SI):209-217.
Waycott M, Duarte CM, Carruthers TJ, Orth RJ, Dennison WC, Olyarnik S, Calladine A,
Fourqurean JW, Heck KL, Hughes AR, Kendrick GA (2009) Accelerating loss of seagrasses
across the globe threatens coastal ecosystems. Proc Nat Acad Sci 106(30):12377-12381.
Worm B, Barbier EB, Beaumont N, Duffy JE, Folke C, Halpern BS, Jackson JB, Lotze, H.K.,
Micheli, F., Palumbi, S.R. and Sala, E., (2006) Impacts of biodiversity loss on ocean ecosystem
services. Science 314(5800):787-790.
Yee S (2010) REDD and BLUE carbon: carbon payments for mangrove conservation.
Responses of Mangrove Ecosystems
to Climate Change in the Anthropocene 9
Daniel M. Alongi

Abstract

Mangrove ecosystems are well-adapted to living in a dynamic and harsh environ-


ment and have survived catastrophic climate events since their first appearance
66 Mya along Tethys Sea shores. There have been past episodes of localized
extinction due mostly to abrupt, rapid rises in sea-level. Living at the edge
between land and sea, mangrove ecosystems are inclined to be resilient to
environmental disturbance, such as highly variable changes in salinity, solar
insolation, temperature, tidal inundation, freshwater inputs, and sea-level.
Based on current knowledge of mangrove responses to environmental change,
mangroves in subsiding river deltas, on low Pacific and Caribbean islands, and in
the arid tropics are likely to decline in area, biodiversity, forest structure, and
function into the Anthropocene. Mangrove forests will continue to expand
latitudinally along the Gulf of Mexico, the east coast of Florida, the coasts of
South Africa, eastern Australia, northern New Zealand, south eastern Brazil, and
China. Regions that experience tropical storms or other extreme events, as in
northern Australia, the Gulf of Mexico, the east India coast, Florida, the
Philippines, south China coast, and the northern Caribbean, mangroves will likely
suffer increased damage, destruction, and loss. The east coast of Sumatra, north
coast of Java, Sulawesi, and southern Vietnam are likely to lose mangroves
because of their low tidal range. Mangroves in other parts of Southeast Asia
may response positively to increased rainfall and warmer temperatures, as on the
west coast of Peninsular Malaysia and the southwest coast of Thailand.
Mangroves will likely respond positively or exhibit no significant change along
the tropical west and north coasts of South America and the Caribbean and Pacific

D. M. Alongi (*)
Tropical Coastal & Mangrove Consultants, Pakenham, VIC, Australia

# The Author(s), under exclusive license to Springer Nature Singapore Pte 201
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_9
202 D. M. Alongi

coasts of Central America. Globally, mangroves are likely to survive, perhaps in


altered states, into the next century.

Keywords

Anthropocene · Climate change · Ecological impacts · Mangroves · Sea-level


rise · Tropics

9.1 Introduction

Mangrove forests and their associated waterways, being at the interface between
land and sea, may portend ecological changes in the coastal zone from climate
change. The intertidal environment in low latitudes where mangroves live is physi-
cally and geologically dynamic. Wide variations in temperature, wave action, tides,
salinity, anoxia, and rainfall are just some of the factors driving and shaping
mangrove ecosystems over time and space (Duke et al. 1998; Alongi 2016).
Mangrove ecosystems are fairly robust and adaptable in the face of environmental
uncertainty, surviving in normal circumstances by key adaptive features such as
large below-ground storage and transformation rates of carbon, nitrogen, and other
nutrients; simple architecture and self-design; highly efficient but complex biotic
controls; species redundancy; and multiple feedbacks which serve to facilitate and
augment recovery from, or resilience to, natural and human disturbances (Alongi
2008, 2015). Evidence for their high adaptability to disturbance comes from patterns
of recovering stands being reminiscent of pioneer-phase forest characteristics, as
forest composition and structure arise from a complex interplay of physiological
tolerances and competitive interactions leading to a mosaic of arrested or interrupted
succession sequences, in response to physiochemical gradients and changes in
shoreline evolution (Lugo 1980; Fromard et al. 1998; Alongi 2008). Not all
mangroves survive disturbances virtually unscathed, as the extent of impact depends
on the extent, severity, and time scale of the disturbance event.
Climate change consists of a series of types of disturbance depending on whether
the effects are direct or indirect. The global rise in sea-level, increases in atmospheric
greenhouse gases and temperatures, and changing atmospheric moisture and precip-
itation patterns are direct effects of climate change. Indirect effects are changes in
seasonal patterns; changes in extreme weather events such as extreme high water
events, droughts, storms and cyclones; acidification; changes in coastal circulation
affecting tidal cycles and tidal exchange of nutrients and dispersal of propagules;
changing salinity gradients affecting tidal exchange and species distribution and
composition; changes in freshwater inflow and allochthonous sediment input; and
degradation of adjacent ecosystems closely linked to mangroves (Alongi 2015;
Jennerjahn et al. 2017). These impacts must be considered against a plethora of
other human impacts that affect many mangrove forests: deforestation; subsidence
due to extraction of water and other natural resources; damming and diversion of
freshwater inflows; changes in riverine sediment inflows by increased catchment
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 203

erosion and damming; overharvesting of mangrove wood, fish, and shellfish; pollu-
tion; changes in hydrology and tidal flushing due to construction of roads, levees,
and dredging of navigation channels.
This chapter assesses the responses of mangroves to climate change and the likely
outcomes over long temporal scales against the background of other human-induced
disturbances. Also, it is important to consider the evidence for how mangroves have
responded to earlier environmental perturbations, such as historical changes in
sea-level, in order to gain some insight into how they may respond to global change
in the Anthropocene.

9.2 Past, Present, and Future Responses to Sea-Level Change

Since their first appearance along the shores of the Tethys Sea during the Late
Cretaceous-Early Tertiary period, mangroves have had to adapt and endure numer-
ous variations in sea-level and other climatic events up to the present. Mangroves in
the Quaternary period experienced a sea-level that was 120–125 m lower than
present at the Last Glacial Maximum, with two periods of very rapid rise (>20 m)
at both 14 and 11 ky BP due to abrupt climatic shifts during the transition from the
last glacial into the present interglacial (Endfield and Marks 2012).
The present distribution of mangroves is a legacy of previous responses to
changes in sediment accretion or erosion, surface elevation, hydrology, weather,
climate, and various disturbances, and interactions among and between populations
and communities (Woodroffe et al. 2016). The existence of relic pollen and peat
deposits, however, provides abundant evidence of dramatic changes in mangroves
over geologic time, especially in relation to late Quaternary changes in sea-level
(Kim et al. 2005; Yulianto et al. 2005; Li et al. 2012; Urrego et al. 2013; Berger et al.
2013), with an overall pattern of paleo-ecological mangrove succession. The geo-
logical record and present analyses of mangrove sediment accretion rates versus
rates of sea-level rise (see Fig. 2 in Alongi 2015) indicate that mangroves have kept
pace with rising sea-level (Woodroffe et al. 2016). Regionally, however, some
mangroves have gone locally extinct in the face of very rapid sea-level rise;
reconstructed historical patterns have shown significant change depending on rate
of sea-level rise as well as rates of subsidence or uplift (McCloskey and Liu 2013;
Punwong et al. 2013; Limaye et al. 2014).
Evidence from deep mangrove peat deposits indicates that over thousands of
years mangroves have kept pace with sea-level rise, as found in the Caribbean
(McKee et al. 2007) and the northern Indian Ocean (Rashid 2014). The fact that
these peat deposits are buried beneath existing mangroves and Holocene sediments,
however, indicates that eventually the rate of sea-level rise increased beyond a
critical threshold at which mangroves were not able to keep pace (McIvor et al.
2013). Records of relic pollen reveal fluctuating responses of mangrove succession
in synchrony with changing sea-level, as found on Borneo (Yulianto et al. 2005), the
Caribbean coast of Colombia (Urrego et al. 2013), the Sunda Shelf of Southeast Asia
204 D. M. Alongi

(Hanebuth et al. 2011), East Africa (Punwong et al. 2013), West Africa (Kim et al.
2005), and the Galápagos (Seddon et al. 2011).
Modern evidence from time series analysis of photos, remote sensing images,
digital terrain models, and rates of sediment accretion implies that mangrove
responses to sea-level rise correspond roughly to patterns of surface elevation
changes. Along the Pacific coast of Mexico, for example, fringing mangroves
drowned because of rising sea-level accompanied by warm waters of El Niño events
but mangroves located in higher intertidal positions were drive inland for a net
increase in mangrove area (Lopez-Medellin et al. 2011). Mangrove responses to
rising sea-level are likely to be complex with local variations playing a key role in
predicting whether mangroves of a given locality or region will survive. For
instance, in Micronesia, sedimentation is sufficient to offset elevation losses on
some islands but not on others as mangroves set low on the shoreline are more
susceptible than forests at higher elevations (Krauss et al. 2010).
The future is in doubt for mangroves occupying tropical river deltas and low
islands, the former currently at risk of changing drastically due to damming, changes
in coastal ocean circulation, and sedimentation and the latter at risk due to their low
profile relative to sea-level. On low islands in the Pacific, mangroves are migrating
landwards with rising sea-level but are constrained by coastal developments such as
roads and levees (Lovelock et al. 2015). In Micronesia and Melanesia local extinc-
tion of mangroves is occurring as sea-level rises and mangroves cannot migrate
landwards (Lovelock et al. 2015). Similarly, many deltaic islands of the Sundarbans
along the Indian and Bangladesh coast are undergoing subsidence and disappearance
as a decline in sediment input from the Ganges and other rivers has caused subsi-
dence and a concomitant decline in mangroves on the central and eastern islands of
the Sundarbans (Rahu et al. 2012; Rashid 2014). In some river deltas sea-level rise,
storms and cyclones coupled with diminishing sediment and freshwater supply have
enhanced subsidence, causing a shift of mangroves to higher elevations. In deltas
across Africa and Asia, about 2% of mangroves disappeared between 2000–2016,
mainly through erosion and conversion to agriculture (Lagomasino et al. 2019).
However, in some regions, there was rapid expansion of mangroves over this 16-yr
period, resulting in new, taller, and more carbon-dense forests. In the Amazon
Macrotidal Mangrove Coast, large, migrating mud banks cyclically accrete and
erode in synchrony with oscillations in tides and sediment erosion and accretion,
resulting in mangroves migrating frequently to adapt to changes in the position of the
shoreline (Schettini et al. 2020).
Are mangroves keeping pace with current rates of sea-level rise? An updated
analysis (Fig. 9.1) of mangrove accretion rates versus local mean sea-level rise
suggests that mangroves (i.e. below the solid line in Fig. 9.1), particularly in
Australia, New Zealand, the Caribbean, Central America and on low Pacific islands
and in subsiding river deltas such as the Sundarbans and in Southeast Asia are not
keeping pace currently. However, mangroves (i.e. data points located above the solid
line in Fig. 9.1) located in other areas of Southeast Asia and the Pacific (e.g. Papua
New Guinea), South America, Africa, the Middle East, South Asia, and South Asia
(primarily China) are keeping pace with current sea-level rise. Many of these latter
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 205

25
North America
Africa
Middle East
20
Mean accretion rate (mm a )

Caribbean and Central America


-1

South Asia
Pacific Islands
East Asia
Australia and New Zealand
15 Southeast Asia
South America

10

0
0 2 4 6 8 10
-1
Current mean sea-level rise (mm a )

Fig. 9.1 The relationship between measured rates of mangrove soil accretion (mm a 1) and current
rates of mean sea-level rise (mm a 1) worldwide. The sea-level rise data are from satellite altimetry
or tide gauge data available on the website: http://www.nodc.noaaa.gov/General/sealevel.html.
Mangrove sedimentation data are from references in Alongi (2009, 2012, 2015, 2018) and Sasmito
et al. (2016) and updated from data in Hayden and Granek (2015), Hoque et al. (2015), Carnero-
Bravo et al. (2016, 2018), MacKenzie et al. (2016), Marchio et al. (2016), Almahasheer et al.
(2017), Phillips et al. (2017), Ruiz-Fernández et al. (2018), Chappel (2018), Cusack et al. (2018),
Murdiyarso et al. (2018), Pérez et al. (2018), Saderne et al. (2018), Fu et al. (2019), Hale et al.
(2019), Kusumaningtyas et al. (2019), Soper et al. (2019) Swales et al. (2019), Bomer et al. (2020),
Hapsari et al. (2020), Matos et al. (2020), Hatje et al. (2021), Passos et al. (2021). The solid line
delimits a 1:1 relationship between accretion rate and rate of sea-level rise

forests occur in areas of rapid accretion due to highly impacted and populated
catchments, especially in China, Brazil, and India. The wide scatter of data points
reflects how mangroves in disparate coastal settings in different parts of world
respond so differently to the same rate of sea-level rise. Most of the mangrove
accretion rates were measured using radionuclides such as 137Cs and 210Pb and are
not highly reliable indicators of accretion compared with the use of surface elevation
tables; there is also considerable uncertainty in the rates of sea-level rise as there are
large variations seasonally and with changes in atmospheric pressure and weather.
The accretion rates also do not reflect the importance of changes in surface elevation
gain; a mangrove forest may be rapidly accumulating soil on the surface but the local
area may be subsiding, resulting in a net decrease relative to sea-level, as is occurring
in the Sundarbans (Hanebuth et al. 2013). A detailed analysis of recent trends in
mangrove surface elevation changes across the Indo-Pacific region using data from
206 D. M. Alongi

surface elevation table instruments shows that for 69% of mangroves the current rate
of sea-level rise exceeded the soil surface elevation gain (Lovelock et al. 2015).
Further, model analysis suggests that mangrove forests located at low tidal range and
low sediment delivery could be submerged by 2070. As shown for macrotidal
mangroves in northern Australia, mangroves have migrated consistently over geo-
logical time in synchrony with post-glacial sea-level rise; how mangroves in this
region adjust in future depends not only on sediment availability but also on local
topography (Woodroffe 2018). In contrast, created mangroves in Tampa Bay,
Florida, store below-ground carbon and rates of surface elevation change enable
them to adjust to sea-level rise (Krauss et al. 2017). A more recent analysis indicates
that while mangroves expanded between 9800 and 7500 years ago at a rate driven
mainly by the rate of relative sea-level rise, it was very likely (90% probability) that
mangroves were unable to sustain accretion when relative sea-level rise exceeded
6.1 mm a 1 (Saintilan et al. 2020), in agreement with the data in Fig. 9.1 at rates of
sea-level rise greater than 6 mm a 1. Mangrove forests are likely losers with respect
to rises in sea-level, especially in regions of substantial subsidence (e.g. the
Sundarbans, the Solomon Islands; Albert et al. 2017), a low tidal range, changes
in precipitation, and poor ecological conditions (Cinco-Castro and Herrera-Silveira
2020). The reality is that mangroves may respond in complex ways to sea-level rise.
Some mangroves will probably survive if the rate of sea-level rise is slow enough in
some locations, but there will likely be significant changes in community composi-
tion, forest structure, morphology, and anatomy, including changes in vascular
vessel densities, fibre wall thickness, bark anatomy, formation of hypertrophied
lenticels, adventitious roots, and increased aerenchyma development (Ellison and
Farnsworth 1997; Wang et al. 2007; Yáñez-Espinosa and Flores 2011; Reef and
Lovelock 2014). Experimental studies indicate species-specific responses to
sea-level rise and waterlogging (Ye et al. 2003, 2004, 2010; Cardona-Olarte et al.
2006; Chen and Wang 2017). High tolerance is exhibited by the cosmopolitan
species, Avicennia marina, to waterlogging, but responses are highly variable in
relation to immersion depth and length of time, salinity, temperature, and other
factors (Lu et al. 2013; Mangora et al. 2014). Another cosmopolitan species,
Rhizophora stylosa, similarly has a high tolerance to waterlogging as experiments
simulating growth of R. stylosa seedlings to sea-level rise found that the species was
flood-tolerant with high stem growth rate and leaf assimilation rate as well as
efficient utilization of carbohydrate reserves stored in hypocotyls of seedlings;
both growth and physiology were affected by salinity and by an increase in flooding
time (Chen and Wang 2017). R. stylosa exhibited competitive dominance at high
salinity, a good adaptation of seedlings to future sea-level rise.
Future survival of mangrove forests ultimately depends on the rapidity of future
increases in sea-level rise. The current prediction (Jevrejeva et al. 2019) is that
sea-level will continue to rise at a median rate of 6.2 mm a 1 until 2100, which is
faster than most current rates of sea-level rise (Fig. 9.1); most mangroves at or above
the rate of 6.2 mm a 1 are at a high risk of drowning.
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 207

9.3 Responses to Rising Atmospheric CO2

Fifty years ago, the mean atmospheric CO2 concentration was 326.17 ppm, 21%
lower than the current CO2 concentration of 416.39 ppm and will continue to rise to
600–800 ppm by 2100 (see https://www.ipcc.ch/data/). Experimental evidence has
repeatedly shown that elevated CO2 concentrations enhance photosynthesis, growth,
and leaf chlorophyll a concentration in mangroves, with responses being species-
specific and variable, depending on salinity, temperature, nutrient availability, and
water-use efficiency (Farnsworth et al. 1996; Ball et al. 1997; Snedaker and Araújo
1998; McKee and Rooth 2008; Cherry et al. 2009; McKee et al. 2012; Reef et al.
2016; Jacotot et al. 2018; Yin et al. 2018; Tamimia et al. 2019; Manea et al. 2020;
Maurer et al. 2020). The early studies in the 1990s showed that growth of
Rhizophora stylosa, R. apiculata, and R. mangle was enhanced by increasing CO2
concentrations, but only a low salinity. Snedaker and Araújo (1998) showed that
R. mangle, Avicennia germinans, Conocarpus erectus, and Laguncularia racemosa
exhibited increases in transpiration efficiency but a decline in stomatal conductance
and transpiration with increasing CO2 concentrations. The response of most man-
grove species to increasing CO2 levels will be complex, with many species thriving,
but some species declining or exhibiting no or little change. Interactive effects
among CO2 concentrations, salinity, temperature, and nutrient availability imply
that intertidal position and regional location may be an important determinant in the
response of most species. Species patterns within an estuary may change based on
the ability of each species to respond to spatial and temporal variations in the above-
mentioned drivers.
More recent experiments have supported the findings of the earlier studies. In a
glasshouse study, Reef et al. (2016) examined the effects of elevated CO2 and
nutrient availability on seedlings of Avicennia germinans and found large gains in
growth and photosynthesis when seedlings grown under elevated CO2 were supplied
with elevated nutrient concentrations compared to their responses under ambient
conditions. Growth was greatly enhanced only under high nutrient conditions and
elevated CO2; root volume doubled under low nutrient and elevated CO2 conditions
relative to ambient nutrient and CO2 levels. Biochemical pathways play a role in
mangrove responses to climate change as A. germinans produces the osmolyte,
glycine betaine, which increases tolerance to environmental stressors. Under expo-
sure to increasing salinity, ambient and high CO2 concentrations, and a dose of
glycine betaine, A. germinans seedlings exhibited increased salt tolerance and higher
photosynthetic rates (Maurer et al. 2020); under elevated CO2, the temperature
optimum for photosynthesis increased by 4  C. In a similar experiment, Reef et al.
(2015) found confounding effects of salinity and elevated CO2 concentrations on the
survivorship, growth, photosynthesis, root architecture and leaf nutrient and ion
concentrations in seedlings of A. germinans. The optimum salinity for growth
shifted higher with greater CO2 concentrations, with carbon assimilation rates
significantly higher under elevated CO2. However, at higher salinity, growth
declined even at elevated CO2 levels, although there was great water-use efficiency.
This outcome was likely due to non-stomatal limitations to growth at high salinities.
208 D. M. Alongi

Under conditions of elevated CO2 concentration (800 ppm) and increased tidal
flooding to simulate sea-level rise, Jacotot et al. (2018) found that net photosynthetic
rates and water-use efficiency of A. marina and R. stylosa were enhanced and more
pronounced in the warm season, suggesting that predicted temperature increases
would further enhance photosynthesis. However, these gains were minimized with
longer flooding duration but only by 5%.
Elevated CO2 also alters the rhizosphere microbiome, as evidenced by growth
experiments with Kandelia obovata (Yin et al. 2018). Over a period of 20 weeks,
elevated CO2 increased leaf chlorophyll a levels and root microbial biomass, with
some alteration of ammonia-oxidizing archaea. Further, there was a shift in carbon
utilization from the preferred carbon sources of sugars, amino acids, and carboxylic
acids under ambient conditions to the use in the order of amino
acids > carbohydrates > polymers > carboxylic acids > amines > phenolic acids,
indicating a shift in metabolic pathways under elevated CO2. When subjected to
ambient CO2 and a temperature of 38  C, R. apiculata seedlings responded posi-
tively to the higher temperature but elevated CO2 (650 ppm) enhanced growth only
at a lower temperature (21  C). Under conditions of high CO2 and temperature, the
seedlings nearly perished (Tamimia et al. 2019), indicating complex outcomes to
elevated CO2 concentrations when mangroves are subjected to other drivers such as
increasing temperatures and flooding.
These complex responses may, nevertheless, offer a competitive advantage to
mangroves as the increasing encroach upon saltmarshes (see next section). Manea
et al. (2020) grew the mangroves, Aegiceras corniculatum and A. marina individu-
ally and in a model saltmarsh community under increasing CO2 and reduced salinity.
The mangroves experienced stronger competition from saltmarsh species under
elevated CO2, A. marina seedlings produced 48% more biomass under elevated
CO2 when grown in competition with saltmarsh species. A. corniculatum was not
affected by elevated CO2 but had 36% greater growth under seawater salinity
compared to hypersaline conditions. Rising atmospheric CO2 concentrations and
lower salinity associated with sea-level rise may thus enhance the establishment of
mangrove seedlings in saltmarshes, facilitating mangrove encroachment.

9.4 Responses to Increasing Temperature

Physiologically, mangrove responses to increasing temperature, as with most plant


and animal species, follow a sigmoid curve with an initial linear rise in respiration,
growth and photosynthesis slowing, plateauing, and then declining as a critical lethal
threshold is reached. The critical temperature threshold is likely to be species-
specific, although rates of leaf photosynthesis for most mangrove species peak at
temperatures at or below 30  C (Alongi 2009). Leaf CO2 assimilation rates of many
species decline as temperature increases from 33 to 35  C.
Temperature increases alone are likely to result in increased growth, reproduc-
tion, photosynthesis and respiration, changes in forest community structure, diver-
sity and are already resulting in an expansion of latitudinal limits in the southern
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 209

United States and the Caribbean (Kennedy et al. 2016; Osland et al. 2013, 2020;
Cavanaugh et al. 2019; Macy et al. 2019), New Zealand (Stokes et al. 2010; Suyadi
et al. 2019), Australia (Rogers et al. 2005; Williamson et al. 2011), southern Africa
(de Boer 2002; Peer et al. 2018), southeastern Brazil (Ximenes et al. 2018), and
southern China (Durango-Cordero et al. 2013). Limits to the expansion of
mangroves are due to the frequency and intensity of extreme cold events (Cavanaugh
et al. 2019). For example, an examination of 38 sites spread across the mangrove
range in the Gulf of Mexico and Atlantic coasts of North America and found that for
A. germinans near their northern range limit, the temperature threshold for leaf
damage is close to 4  C with mortality thresholds closer to 7  C (Osland et al.
2020).
Mangrove expansion may be coupled to changes in precipitation (Wang et al.
2014) as temperature alone does not always delimit the latitudinal range of
Rhizophora and Avicennia because of large regional differences in monthly temper-
ature change, as warmest monthly temperatures are higher at the latitudinal limits in
the northern than in the southern hemisphere (Quisthoudt et al. 2012). Mangrove
expansion and subsequent saltmarsh contraction are consistent with the poleward
increase in temperatures and the frequency and intensity of extreme cold events, but
other factors such as extreme weather events, genetic plasticity, impacts of changes
in ocean currents on dispersal abilities, and changes in precipitation are likely
co-factors (Saintilan et al. 2014; Kennedy et al. 2016; Ximenes et al. 2018).
Rising temperatures and expansion of mangroves further into higher latitudes
may impact other mangrove flora and fauna. Rising temperatures also affect animal
physiology in different ways. For example, the generalist fiddler crab, Minuca
rapax, loses less water than the fiddler crab, Leptuca thayeri (Principe et al. 2018)
due to differences in carapace permeability; survivability is higher for M. rapax
under desiccated conditions but is more affected by temperature increase on its
physiology. The fiddler crab, Leptuca uruguayensis is more sensitive to warming
than Leptuca leptodactyla showing physiological and behavioural differences
(da Silva Vianna et al. 2020). The latter species was able to adjust its metabolic
rate to temperature increase and reducing ammonia excretion, whereas the former
species showed adaptation limits. Fiddler crabs inhabiting vegetated areas are more
vulnerable to higher temperatures and may change its geographic range, while
fiddler crabs on mud and sandflats are more tolerant to temperature rise and may
have a competitive advantage as global temperatures rise. A large-scale mesocosm
experiment found that an increase of 1.2  C leads to the homogenization and
flattening of mangrove root epibiont communities, leading to 24% increase in the
overall cover of algal epibionts on roots but 33% decline in diversity of epibiont
species and a decrease in structural complexity (Walden et al. 2019).
Along the South African coast, fiddler crabs have spread farther south while other
fauna such as the gastropod, Terebralia palustris, has disappeared, although there
does not appear to be a decrease in diversity with an increase in latitude (Peer et al.
2018). The transition to higher latitudes and the expansion of mangroves at the
expense of saltmarshes indicates the complex interactions are occurring during the
transition, most notably competition. Several studies have indicated that mangrove
210 D. M. Alongi

species that invade saltmarsh areas are superior competitors (Macy et al. 2019;
Manea et al. 2020). Migration may be mediated by biotic interactions and may be
facilitated by increasing propagule abundance from greater reproductive rates and
greater genetic variation due to outcrossing (Proffitt and Travis 2014). Surveying the
Atlantic and Gulf coasts of Florida, Proffitt and Travis (2014) found that reproduc-
tive frequencies varied significantly but increased with latitude and strongly along
the Gulf coast with a concomitant increase in outcrossing. Adaptation to a new
environment is perpetuated and promoted by the self-enforcing nature of migration
as more colonizers lead to more propagules and outcrossing leads to greater genetic
variation.

9.5 Responses to Changes in Precipitation and Extreme


Weather Events

Mangrove responses to increasing, decreasing, or more variable precipitation may be


less complex, but any responses must be considered with co-occurring drivers such
as increasing temperatures, atmospheric CO2 concentrations, and rises in sea-level.
Generally, mangrove forests attain their peak biomass and productivity in the wet
tropics due to continually warm temperatures and high rainfall (Alongi 2009). Thus,
increases in precipitation will likely result in greater biomass and productivity.
Predictions of shifts in tropical rainfall patterns indicate two views: one predicts
that wet regions will get wetter and the other view suggests increased rainfall where
the rise in sea surface temperature exceeds the mean surface warming in the tropics,
i.e. “warmer-get-wetter”. Computer simulations indicate that the pattern of ocean
warming induces ascending atmospheric flow at the Equator and subsidence with
distance from the Equator, anchoring a band of annual rainfall increase near the
Equator that reflects the “warmer-get-wetter” view (Huang et al. 2013). However,
the ascending motion oscillates back and forth across the Equator with the Sun,
pumping moisture upwards and causing seasonal rainfall anomalies following a
“wet-get-wetter” pattern. Thus, seasonal mean rainfall in the tropics combines the
“wet-get-wetter” and “warmer-get-wetter” patterns. Recently, rainfall-based
thresholds have been identified for mangrove range limits in western North America,
western Gulf of Mexico, western South America, western Australia, the Middle East,
northwest Africa, east central Africa, and west central Africa (Osland et al. 2017).
Not only will there be changes in precipitation rates and in regional patterns, but
the frequency and intensity of extreme weather events, such as cyclonic storms and
prolonged droughts, is predicted to increase. Temperature anomalies, defined as
extreme temperature events more than three standard deviations from the long-term
mean from 1951–1980, have shifted more than one standard deviation towards
higher values, leading to more extreme warming events (Hansen et al. 2012). The
increased occurrence of such events is having a dramatic impact on mangroves, such
as mass mortality events (Lovelock et al. 2017). This shift towards extreme temper-
ature events has also resulted in a concomitant decrease in extreme freeze events
(Hayes et al. 2020) resulting in a shift favouring mangrove expansion within the
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 211

mangrove-marsh ecotone and expanding their range poleward, as studied most


extensively along the Gulf of Mexico and the Florida Atlantic coastline (Cavanaugh
et al. 2014; Saintilan et al. 2014; Osland et al. 2020). The clearest example of
mangrove mortality due to an extreme weather event is the massive dieback of
mangroves in the Gulf of Carpentaria in northern Australia (Duke et al. 2017; Sippo
et al. 2018, 2020). Mangrove forests in the gulf suffered severe dieback (6% of
vegetation) during the summer of 2015–2016 along 1000 km of shoreline. The onset
of the dieback was coincident with unprecedented high temperatures, low precipita-
tion, and lack of the normal summer wet season (Duke et al. 2017). An unusually
lengthy severe drought coupled with a temporary drop in sea-level contributed to
mass mortality. The dieback had severe consequences for mangrove functioning as
evidenced by a shift from a dominance of oceanic carbon outwelling to increased
atmospheric CO2 emissions and decreased alkalinity exports (Sippo et al. 2020); this
shift was likely driven by reduced mangrove productivity and increased oxygen soil
permeation. Dieback of mangroves in the Gulf of Carpentaria also resulted in a
trophic shift in crabs with a loss of litter-feeding crabs but an increase in crabs that
feed on microphytobenthos; infauna was unaffected (Harada et al. 2019). An
examination of data since the 1960s showed about 36,000 ha of mangrove mortality
worldwide, about 70% of which was attributed to low frequency, high intensity
weather events (typhoons, cyclones, hurricanes) and climate extremes (Sippo et al.
2018). Tropical cyclonic storms account for 45% of the reported mangrove mortality
area since the 1960s, being the largest cause of mortality. Recent large-scale
mortality, however, associated with extreme climatic events in Australia accounts
for 22% of all reported forest loss over the past six decades, suggesting the
increasing importance of extreme climatic events (Sippo et al. 2018). In Mangrove
Bay in north Western Australia, there have been two dieback events over a 16-year
period, with the most recent one coincident with the dieback in the Gulf of
Carpentaria (Lovelock et al. 2017). The diebacks in Mangrove Bay were coincident
with periods of very low sea-level due to intensification of El Niño-Southern
Oscillation (ENSO) leading to increased soil salinities and subsequent canopy loss
and reduced recruitment. In June 2016, a hailstorm with winds over 100 km h 1
caused mangrove dieback in eastern Brazil (Servino et al. 2018). The losses were
extensive, corresponding to 29% of total forest area; 15-mo later some areas of
dieback showed initial recovery while others continued to degrade. Recovery was
not helped by the El Niño creating mild drought conditions. The impact of any one
cyclonic disturbance is often significant, with some stands taking decades to fully
recover, as evidenced from mangrove forests in the Caribbean and Florida (Imbert
2018; Rivera-Monroy et al. 2019); such climatic disturbances prevent these
ecosystems from ever reaching a steady-state. However, not all mangroves are
significantly affected by extreme climate events as found for mangroves on a
Colombian Caribbean Island that appear to be resilient to short drought events
related to ENSO (Galeano et al. 2017). At the other extreme, mangroves such as
A. germinans may be resilient to extreme freeze events due to genetically based
freeze tolerances (Hayes et al. 2020). Nevertheless, increasing mortality events are
likely in future in light of forecasts of increasing frequency, intensity, and
212 D. M. Alongi

destructiveness of cyclonic storms and climatic extremes, such as heat waves and
low and high sea-level episodes.

9.6 Responses to Coastal Ocean Acidification

Ocean acidification occurs because the global ocean takes up about one third of the
atmospheric carbon released from fossil fuel combustion, cement production, and
land-use change, with the subsequent hydrolysis of increasing amounts of CO2 in
seawater increasing the hydrogen ion concentration thereby reducing the pH of
ocean water and causing wholesale shifts in seawater carbonate chemistry (Doney
et al. 2009). Surface ocean Ph has declined since preindustrial times by about 0.1
unit and is predicted to decline a further 0.3–0.4 units by later this century. Acidifi-
cation in the coastal ocean is a more complex process as carbonate chemistry in
estuarine and coastal waters is strongly regulated by changes in biological activity
related to increases in anthropogenic delivery of nutrients by rivers, groundwater,
and eutrophication (Borges and Gypens 2010). Eutrophication strongly amplifies
acidification due to the accumulation of algal biomass and its subsequent decompo-
sition, decreasing dissolved oxygen (DO) levels and contributing to hypoxia
(Wallace et al. 2014). Hypoxia increases pCO2 values and upwelling processes, if
any, can bring CO2 enriched water to coastal waters, amplifying the effects of
acidification.
Acidification has a significant impact on tropical marine life, with several marine
organisms of some taxonomic groups showing decreased growth and physiological
tolerance to lower pH. These groups include tropical hermatypic corals which show
a significant decline in calcification, the process involving the formation of calcium
carbonate skeletons. Other calcifying organisms such as photosynthetic calcareous
algae, photosynthetic symbiont-bearing foraminifera, and some species of molluscs,
jellyfishes, fishes, echinoderms, and pteropods similarly show a decline in calcifica-
tion as well as shell dissolution; reduction in shell mass; growth reductions; reduced
metabolism, fertility, and embryo development; increased mortality; reduced ther-
mal tolerance; reduced food intake and increased ventilation (Hofmann et al. 2010).
Not all estuarine, coastal, and shelf organisms exhibit a negative response to
acidification, as many organisms show either a positive or mixed response to
acidification. Seagrasses, brown macroalgae, kelps, sea anemones, fishes, and most
non-calcifying organisms exhibit positive responses, while some Caribbean corals,
plankton, fleshy macroalgae, upright calcareous algae, and crustose coralline algae
show contrasting responses (negative, positive, no effect), depending on species-
specific tolerances and the degree of acidification (Alongi 2020).
Seagrasses and mangroves will be the most resilient ecosystems to the effects of
coastal acidification. Several seagrass species usually respond positively, or not at
all, to lower pH by their capacity to modify pH within canopies and within their
habitat. Mangrove ecosystems may prove to be the most resistant and resilient in the
face of coastal acidification. As detailed by Alongi (2020), the pH of mangrove soils
is ordinarily low, within the range of 4–7, especially in the mangroves of South and
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 213

Southeast Asia and Africa. Mangrove soils have low pH due to high rates of bacterial
respiration, high concentrations of polyphenolic acids, and the net effects of meta-
bolic processes associated with the trees and their root systems.
Nearly all estuarine and nearshore waters in the tropics, including mangrove tidal
creek waters and other waterways, naturally exhibit very wide variations in salinity
(range: 0.1–48), pH (range: 4–9), pCO2 (range: 4–32,763μatm), and [CO32 ] and are
a strong source of CO2 emissions to the atmosphere due to pCO2 and [CO32 ]
oversaturation (see Table 4 in Alongi 2020). Oversaturation and highly variable pH
are the net result of high rates of bacterial respiration, eutrophication, and the
influence of fluvial discharge, including export of alkalinity, organic matter and
dissolved inorganic carbon (DIC), deposition of anthropogenic acids and bases,
intense weathering, land-use change, acid sulphate soil discharge, and acidic ground-
water. These chemical factors predispose mangroves and other coastal ecosystems to
resilience to coastal acidification.
Mangroves are apparent buffers of acidification in the tropical coastal zone (Sippo
et al. 2016). Carbon (DIC, dissolved CO2) and alkalinity dynamics measured in six
Australian mangrove tidal creeks showed a mean export of DIC, but alkalinity fluxes
ranged from a small import and a large export; a net import of free CO2 was
measured, equivalent to one third of the estimated air–water CO2 flux. Sippo et al.
(2016) upscaled these results globally and showed that mangrove alkalinity exports
are equivalent to about 14% of global river or continental shelf alkalinity fluxes. The
effect of DIC and alkalinity exports creates a measurable increase in pH indicating
that mangroves partly counteract acidification in adjacent tropical coastal waters.
Mangroves may thus be one of the largest sources of alkalinity to the tropical coastal
ocean, providing buffering against acidification.

9.7 Predictions and Conclusions

Considerable uncertainty exists in predicting mangrove responses to climate change,


but reasonable prognostications can be made based on our considerable knowledge
of current mangrove responses to temperature, current sea-level rise, salinity, rain-
fall, and past and present impacts of tropical storm systems and extreme drought.
Likely regional responses to projected changes in climate are listed in Table 9.1.
Acidification is not included as the predicted decline in coastal ocean pH is within
the boundaries of current pH variability in mangrove waters and soils. The likely
scenarios assume that mangrove responses will include complex changes in floral
and faunal species composition, morphology, biodiversity, biomass, physiology,
growth and productivity, and that functionality (growth, respiration, productivity,
etc.) will increase with increases in temperature, precipitation, and atmospheric CO2
concentrations up to the critical physiological thresholds of individual species, but
only when other environmental conditions are favourable.
Regionally, mangrove forests will continue to expand latitudinally along the Gulf
of Mexico, the east coast of Florida, the coasts of South Africa, eastern Australia,
northern New Zealand, southeastern Brazil, and China. On low islands in the Pacific
Table 9.1 Predicted responses of mangrove ecosystems to expected climate change impacts by 2100
214

Tropical
Sea-level cyclone/
rise (SLR) Temperature extreme events
Region (mm a 1) increase Precipitation change Salinity change increase Responses
North 3.0–5.1 Very likely Increase very likely No change Likely Continued latitudinal expansion along
America Gulf of Mexico coast as  C and [CO2]
increase. Increase in damage/
destruction from increased frequency
of hurricanes.
Africa 1.1–3.8 Very likely Increase in North and Decrease (0–2.0) in No change Mangroves continue poleward
South, decrease in Central West and expansion in South Africa. Most
West East Africa mangroves highly fragmented so at
higher risk of losses due to continued
deforestation and degradation.
Middle East 2.2–3.3 Very likely Likely decrease Increase (0–2.0) No Change Landward expansion/migration
(up to 7  C) unlikely. Likely losses in Red Sea and
the Levant due to high  C, less rain,
higher salinity.
Caribbean 1.0–2.5 Very likely Increase very likely Increase (0.5–1.0) Likely Most low island mangroves not
and Central keeping pace with current SLR
America (Fig. 9.1) but predicted increase in
rainfall during ENSO likely to increase
sediment delivery along S Caribbean
coast of Central America. Arid-zone
mangroves likely to decline as
salinities and  C increase, especially
both Mexico coasts.
South Asia 1.0–1.4 Very likely Increase very likely Increase (0–0.5) Likely Arid-zone mangroves and those in
in India and Sri Indus delta of Pakistan likely to decline
Lanka due to low rainfall, subsidence, and
D. M. Alongi
9

low sediment delivery. E India


vulnerable due to low tidal range,
subsidence, and increased strength and
frequency of cyclones, especially the
Sundarbans.
Pacific 1.4–2.0 Very likely Increase very likely No change Likely Low islands of Oceania highly
Islands vulnerable to SLR due to lack of
upland space for landward migration.
East Asia 1.4–3.4 Very likely Increase very likely Increase (0–0.5) Likely S China mangroves vulnerable to
predicted SLR, low sediment input,
and increased strength and frequency
of typhoons.
Australia and 2.0–3.8 Very likely Increase very likely No change Likely Most mangroves in SE Queensland and
New Zealand SE Australian coast not keeping pace
with current SLR (Fig. 9.1) but
continued encroachment into salt
marshes. Mangroves in NW Australia
likely to decline due to increasing
aridity and SLR. New Zealand
mangroves likely to continue to expand
on north island as  C warm.
Southeast 2.3–4.5 Very likely Increase very likely Decrease (0–1.0) Likely E Philippine mangroves vulnerable to
Asia increased strength and frequency of
typhoons. E Sumatra, Sulawesi, and S
Vietnam vulnerable to predicted SLR
because of low tidal range. River delta
Responses of Mangrove Ecosystems to Climate Change in the Anthropocene

mangroves (e.g. Mekong,


Ayeyarwaddy deltas) likely to decline
with SLR, subsidence, and decreased
sediment supply from damming and
river modification.
215

(continued)
Table 9.1 (continued)
216

Tropical
Sea-level cyclone/
rise (SLR) Temperature extreme events
Region (mm a 1) increase Precipitation change Salinity change increase Responses
South 2.0–3.5 Very likely Very likely increase Increase (0–1.0) Likely South American mangroves along
America in west and north, north coast unlikely to be impacted as
decrease in SE increased rainfall will result in
increased sediment supply
(e.g. Amazon, Orinoco Rivers). SE
Brazil mangroves likely to expand
south due to increasing  C. Mangroves
on Pacific coast of South America
likely to continue increasing stand size,
but no latitudinal extension due to cold
Peruvian current and arid conditions.
Sources: IPCC (2013, 2014), Alongi (2015), Gilman et al. (2016), He et al. (2016), Ward et al. (2016), Wilson (2017), Amuzu et al. (2018), Mukhopadhyay et al.
(2018), Schuerch et al. (2018), Mafi-Gholami et al. (2020)
D. M. Alongi
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 217

and in the Caribbean, mangrove will likely decline as sea-level rise traps mangroves
that have no or little upland space to colonize. In the arid tropical regions of
northwest Australia, Pakistan, the Arabian Peninsula, and the coasts of Mexico,
mangrove forest will likely decline due to increased aridity as salinities increase,
freshwater becomes scarcer, and critical temperature thresholds are reached more
frequently. Mangroves will likely decline in tropical river deltas subject to subsi-
dence such as the Sundarbans, Mekong, Zaire, Ayeyarwaddy, and Fly Rivers, as
sediment and freshwater delivery declines due to damming and alterations to river
flow and as sea-level rises. In areas that experience tropical storms or other extreme
events, such as in northern Australia, the Gulf of Mexico, the east India coast,
Florida, the Philippines, south China coast, and the northern Caribbean, mangroves
will likely suffer increased damage, destruction, and loss. Some areas of Southeast
Asia, such as the east coast of the island of Sumatra, north coast of Java, Sulawesi,
and southern Vietnam are likely to lose mangroves because of their low tidal range.
In contrast, mangroves in other parts of Southeast Asia may response positively to
increased rainfall and warmer temperatures (e.g. west coast of Peninsular Malaysia,
southwest coast of Thailand). Mangroves will likely respond positively or exhibit no
significant change along the tropical west and north coasts of South America and the
Caribbean and Pacific coasts of Central America (e.g. Costa Rica).
Mangroves are likely to survive further into the Anthropocene, but most impacts
will be negative rather than positive. However, the greatest threat to mangroves
continues to be, and will likely remain, deforestation and degradation which must be
considered with climate change impacts.

References
Albert S, Saunders MI, Roelfsema, Leon JX, Johnstone E, Mackenzie JR, Hoegh-Guldberg O,
Grinham AR, Phinn SR, Duke NC, Mumby PJ, Kovacs E, Woodroffe CD (2017) Winners and
losers as mangrove, coral and seagrass ecosystems respond to sea-level rise in Solomon Islands.
Environ Res Lett 12: 094009, https://doi.org/10.1088/1748-9326/aa7e68.
Almahasheer H, Serrano O, Duarte CM, Arias-Ortiz A, Masque P, Irigoien X (2017) Low carbon
sink capacity of Red Sea mangroves. Sci Rept 7:9700, https://doi.org/10.1038/s41598-017-
10424-9.
Alongi DM (2008) Mangrove forests: resilience, protection from tsunamis, and responses to global
climate change. Estuar Coast Shelf Sci 76:1–13.
Alongi DM (2009) The energetics of mangrove forests. Springer, Netherlands.
Alongi DM (2012) Carbon sequestration in mangrove forests. Carbon Manage 3:313–322.
Alongi DM (2015) The impact of climate change on mangrove forests. Curr Clim Change Rep 1:
30–39.
Alongi DM (2016) Mangroves. In: Kennish M (ed) Encyclopedia of estuaries, Springer-Verlag,
Berlin, p 393–404.
Alongi DM (2018) Blue carbon: coastal sequestration for climate change mitigation. Springer
Briefs in Climate Studies, Springer Nature, Cham Switzerland.
Alongi DM (2020) Vulnerability and resilience of tropical coastal ecosystems to ocean acidifica-
tion. Examines Mar Biol Oceanogr 3, https://doi.org/10.313031/EIMBO.2020.03.000562.
218 D. M. Alongi

Amuzu J, Jallow BP, Kabo-Bah AT, Yaffa S (2018) The climate change vulnerability and risk
management matrix for the coastal zone of the Gambia. Hydrology 5: 14, 10.10.3390/
hydrology5010014.
Ball MC, Cochrane MJ, Rawson HM (1997) Growth and water use of the mangroves Rhizophora
apiculata and R. stylosa in response to salinity and humidity under ambient and elevated
concentrations of atmospheric CO2. Plant Cell Environ 20: 1158–1166.
Berger JF, Charpentier V, Crassard R, Martin C, Davtian G, Lopez-Sáez JA (2013) The dynamics
of mangrove ecosystems, changes in sea level and the strategies of Neolithic settlements along
the coast of Oman (6000-3000 cal. B.C.). J Archaeol Sci 40:3087–3104.
Bomer EJ, Wilson CA, Hale RP, Hossain AN, Rahman FA (2020) Surface elevation and sedimen-
tation dynamics in the Ganges-Brahmaputra tidal delta plain, Bangladesh: evidence for man-
grove adaptation to human-induced tidal amplification. Catena 187:104312, https://doi.org/10.
1016/j.catena.2019.104312.
Borges AV, Gypens N (2010) Carbonate chemistry in the coastal zone responds more strongly to
eutrophication than to ocean acidification. Limnol Oceanogr 55: 346–353.
Cardona-Olarte P, Twilley RR, Krauss KW, Rivera-Monroy V (2006) Responses of neotropical
mangrove seedlings grown in monoculture and mixed culture under treatments of hydroperiod
and salinity. Hydrobiol 569; 325–391,
Carnero-Bravo V, Sanchez-Cabeza JA, Ruiz-Fernández AC, Merino-Ibarra M, Hillaire-Marcel C,
Corcho-Alvarado JA, Röllin S, Diaz-Asencio M, Cardoso-Mohedano JG, Zavala-Hidalgo J
(2016) Sedimentary records of recent sea-level rise and acceleration in the Yucatan Peninsula.
Sci Tot Environ 573:1063–1069.
Carnero-Bravo V, Sanchez-Cabeza JA, Ruiz-Fernández AC, Merino-Ibarra M, Corcho-Alvarado
JA, Sahli H, Hélie JF, Preda M, Zavala-Hidalgo J, Díaz-Asencio M, Hillaire-Marcel C (2018)
Sea-level rise sedimentary record and organic carbon fluxes in a low-lying tropical coastal
ecosystem. Catena 162:421–430.
Cavanaugh KC, Kellner JR, Forde AJ, Gruner DS, Parker JD, Rodriguez W, Feller IC (2014)
Poleward expansion of mangroves is a threshold response to decreased frequency of extreme
cold events. Proc Nat Acad Sci 111:723–727.
Cavanaugh KC, Dangremond EM, Doughty CL, Williams AP, Parker JD, Hayes MA,
Rodriguez W, Feller IC (2019) Climate-driven regime shifts in a mangrove-salt marsh ecotone
over the past 250 years. Proc Nat Acad Sci 116:21602–21608.
Chappel AR (2018) Soil accumulation, accretion, and organic carbon burial rates in mangrove soils
of the Lower Florida Keys: a temporal and spatial analysis. MS Thesis, Univ S Florida, St
Petersburg, USA.
Cherry JA, McKee KL, Grace JB (2009) Elevated CO2 enhances biological contributions to
elevation change in coastal wetlands by offsetting stressors associated with sea-level rise. J
Ecol 7:67–77.
Cinco-Castro S, Herrera-Silveira J (2020) Vulnerability of mangrove ecosystem to climate change
effects: the case of the Yucatan Peninsula. Ocean Coast Manage 192: 105196, https://doi.org/10.
1016/j.ocecoaman.2020.105196.
Chen L, Wang W (2017) Ecophysiological responses of viviparous mangrove Rhizophora stylosa
seedlings to simulated sea-level rise. J Coast Res 33: 1333–1340.
Cusack M, Saderne V, Arias-Ortiz A, Masque P, Krishnakumar PK, Rabaoui L, Qurban MA,
Qasem AM, Prihartato P, Loughland RA, Elyas AA (2018) Organic carbon sequestration and
storage in vegetated coastal habitats along the western coast of the Arabian Gulf. Environ Res
Lett 13:074007, https://doi.org/10.1088/1748-9326/aac899.
de Boer WF (2002) The rise and fall of the mangrove forests in Maputo Bay, Mozambique.
Wetlands Ecol Manage 10:313–322.
de Silva Vianna B, Miyai CA, Augusto A, Costa TM (2020) Effects of temperature increase on the
physiology and behavior of fiddler crabs. Physiol Beh 215; 112765, https://doi.org/10.1016/j.
physbeh.2019.112765.
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 219

Doney SC, Fabry VJ, Feely RA, Kleypas JA (2009) Ocean acidification: the other CO2 problem.
Annu Rev Mar Sci 1: 169–192.
Duke NC, Ball MC, Ellison JC (1998) Factors influencing biodiversity and distributional gradients
in mangroves. Glob Ecol Biogeogr Lett 7: 27–47.
Duke NC, Kovacs JM, Griffiths AD, Preece L, Hill DJE, van Oosterzee P, Mackenzie J, Morning
HS, Burrows D (2017) Large-scale dieback of mangroves in Australia’s Gulf of Carpentaria: a
severe ecosystem response, coincidental with an unusually extreme weather event. Mar Freshw
Res 68: 1816–1829.
Durango-Cordero JS, Satyanarayana B, Zhang J, Wang J, Chen M, Fanghong X, Chan JC,
Kangying L, Bogaert J, Koedam N, Dahdouh-Guebas F (2013) Vegetation structure at
Zhangiang Mangrove National Nature Reserve (ZMMNR), PR China: a comparison between
original and non-original trees using ground-truthing, remote sensing and GIS techniques. http://
www.vliz.be/imisdocs/publications/232700.pdf.
Ellison AM, Farnsworth EJ (1997) Simulated sea level change alters anatomy, physiology, growth,
and reproduction of red mangrove (Rhizophora mangle L.). Oecologia 112: 435–446.
Endfield GH, Marks RB (2012) Historical environmental change in the tropics. In: Metcalf SE,
Nash DJ (eds) Quaternary environmental changes in the tropics, Wiley, Chichester, p 360–391.
Farnsworth EJ, Ellison AM, Gong WK (1996) Elevated CO2 alters anatomy, physiology, growth,
and reproduction of red mangrove (Rhizophora mangle L.). Oecologia 108: 599–609.
Fromard F, Puig H, Mougin E, Marty G, Betoulle A, Cadamuro L (1998) Structure, above-ground
biomass, and dynamics of mangrove ecosystems: new data from French Guiana. Oecologia 115:
39–53.
Fu H, Zhang Y, Ao X, Wang W, Wang M (2019) High surface elevation gains and prediction of
mangrove responses to sea-level rise based on dynamic surface elevation changes at
Dongzhaigang Bay, China. Geomorphol 334:194–202.
Galeano, A., Urrego, L.E., Botero, V. and Bernal, G. (2017). Mangrove resilience to climate
extreme events in a Colombian Caribbean Island. Wetlands Ecology and Management
25:743–760
Gilman E, Van Lavieren H, Ellison J, Jungblut V, Wilson L, Areki F, Brighouse G, Bungitak J,
Dus E, Henry M, Sauni Jr I, Kilman M, Matthews E, Teariki-Ruatu N, Tukia S, Yuknavage K
(2016) Pacific Island mangroves in a changing climate and rising sea. UNEP Regional Seas
Reports and Studies No. 179. UNEP, Regional Seas Programme, Nairobi, Kenya.
Hale RP, Wilson CA, Bomer EJ (2019) Seasonal variability of forces controlling sedimentation in
the Sundarbans National Forest, Bangladesh. Front Earth Sci 7:211, https://doi.org/10.3389/
feart.2019.00211.
Hanebuth TJJ, Voris HK, Yokoyama Y, Saito Y, Okuno J (2011) Formation and fate of sedimentary
depocentres on Southeast Asia’s Sunda Shelf over the past sea-level cycle and biogeographic
implications. Earth Sci Rev 104: 92–110.
Hanebuth TJ, Kudrass HR, Lindstader J, Islam B, Zander AM (2013) Rapid coastal subsidence in
the central Ganges-Brahmaputra Delta (Bangadesh) since the 17th century deduced from
submerged salt-producing kilns. Geol 41:987–990.
Hansen J, Sato M, Ruedy R (2012) Perception of climate change. Proc Nat Acad Sci 109: E2415–
E2423.
Hapsari KA, Jennerjahn TC, Lukas MC, Karius V, Behling H (2020) Intertwined effects of climate
and land use change on environmental dynamics and carbon accumulation in a mangrove-
fringed coastal lagoon in Java, Indonesia. Glob Change Biol 26: 1414–1431.
Harada Y, Fry B, Lee SY, Maher DT, Sippo JZ Connolly RM (2019) Stable isotopes indicate
ecosystem restructuring following climate-driven mangrove dieback. Limnol Oceanogr 65:
1251–1263.
Hatje V, Masque P, Patire VF, Dorea A, Barros F (2021) Blue carbon stocks, accumulation rates,
and associated spatial variability in Brazilian mangroves. Limnol Oceanogr 66:321–334.
Hayden HL, Granek EF (2015) Coastal sediment elevation change following anthropogenic
mangrove clearing. Estuar Coast Shelf Sci 165:70–74.
220 D. M. Alongi

Hayes MA, Shor AC, Jesse A, Miller C, Kennedy JP, Feller I (2020) The role of glycine betaine in
range expansions; protecting mangroves against extreme freeze events. J Ecol 108: 61–69.
He YH, Mok HY, Lai EST (2016) Projection of sea-level change in the vicinity of Hong Kong in
the 21st century. Intl J Climat 36: 3237–3244.
Hofmann GE, Barry JP, Edmunds PJ, Gates RD, Hutchins DA, Klinger T, Sewell MA (2010) The
effect of ocean acidification on calcifying organisms in marine ecosystems: an organism to
ecosystem perspective. Annu Rev Mar Sci 41: 127–147
Hoque MM, Abu Hena MK, Ahmed OH, Idris MH, Rafiqul HA, Billah MM (2015) Can mangroves
help combat sea-level rise through sediment accretion and accumulation? Malay J Sci 34:78–86.
Huang P, Xie S-P, Hu K, Huang G, Huang R (2013) Patterns of the seasonal response of tropical
rainfall to global warming. Nature Geosci 6: 357–361.
Imbert D (2018) Hurricane disturbance and forest dynamics in east Caribbean mangroves. Eco-
sphere 9: e02231, https://doi.org/10.1002/ecs2.2231.
IPCC (2013) Climate change 2013: the physical science basis. Cambridge University Press,
New York, USA.
IPCC (2014) Climate change 2014: impacts, adaptation, and vulnerability. Cambridge University
Press, New York, USA.
Jacotot A, Gensous S, Marchand C, Allenbach M (2018) Effects of elevated atmospheric CO2 and
increased tidal flooding on leaf gas-exchange parameters of two common mangrove species:
Avicennia marina and Rhizophora stylosa. Photosynth Res 138: 249–260.
Jennerjahn, TC, Gilman E, Krauss KW, Lacerda LD, Nordhaus I, Wolanski E (2017) Mangrove
ecosystems under climate change. In: Rivera-Monroy VH, Lee SY, Kristensen E, Twilley RR
(eds), Mangrove ecosystems: a global biogeographic perspective, Springer, Cham, Switzerland,
p 211–244.
Jevrejeva S, Frederikse T, Kopp RE, Cozannet GL, Jackson LP, van de Wal RSW(2019) Probabi-
listic sea level projections at the coast by 2100. Survey Geophys 40: 1673–1696.
Kennedy JP, Pil MW, Proffitt CE, Boeger WA, Stanford AM, Devlin DJ (2016) Postglacial
expansion pathways of red mangrove, Rhizophora mangle, in the Caribbean Basin and Florida.
Am J Bot 103: 260–276.
Kim J-H, Dupont L, Behling H, Versteegh GJM (2005) Impacts of rapid sea-level rise on mangrove
deposit erosion: application of teraxeroland Rhizophora records. J. Quatern Sci 20: 221–225.
Krauss KW, Cahoon DR, Allen JA, Ewel KC, Lynch JC, Cormier N (2010) Surface elevation
change and susceptibility of different mangrove zones to sea-level rise on Pacific high islands of
Micronesia. Ecosystems 13: 129–143.
Krauss KW, Cormier N, Osland MJ, Kirwan ML, Stagg CL, Nestlerode JA, Russell MJ, From AS,
Spivak AC, Dantin DD, Harvey JE, Almario AE (2017) Created mangrove wetlands store
belowground carbon and surface elevation change enables them to adjust to sea-level rise. Sci
Rept 7: 1030, https://doi.org/10.1038/s41598-017-01224-2.
Kusumaningtyas MA, Hutahaean AA, Fischer HW, Pérez-Mayo M, Ransby D, Jennerjahn TC
(2019) Variability in the organic carbon stocks, sources, and accumulation rates of Indonesian
mangrove ecosystems. Estuar Coast Shelf Sci 218:310–323.
Lagomasino D, Fatoyinbo T, Lee S, Feliciano E, Trettin, Shapiro A, Mangora MM (2019)
Measuring mangrove carbon loss and gain in deltas. Environ Res Lett 14:025002, https:doi.
org/10.1088/1748-9326/aaf0de.
Li Z, Saito Y, Mao L, Tamura T, Li Z, Song B, Zhang Y, Lu A, Sieng S, Li J (2012) Mid-Holocene
mangrove succession and its response to sea-level change in the upper Mekong River delta,
Cambodia. Quatern Res 78:386–399.
Limaye RB, Kumaran KPN, Padmalal D (2014) Mangrove habitat dynamics in response to
Holocene sea level and climate changes along southwest coast of India. Quatern Intl 325:
116–125.
Lopez-Medellin X, Ezcurra E, González-Abraham C, Hak J, Santiago LS, Sickman JO (2011)
Oceanographic anomalies and sea level rise drive mangroves inland in the Pacific coast of
Mexico. J Veg Sci 22: 143–151.
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 221

Lovelock CE, Cahoon DR, Friess DA, Guntenspergen GR, Krauss KW, Reef R, Rogers K,
Saunders ML, Sidik F, Swales A, Saintilan N, Thuyen LX, Triet T (2015) The vulnerability
of Indo-Pacific mangrove forests to sea-level rise. Nature 526:559–563.
Lovelock, CE, Feller IC, Reef R, Hickey S, Ball MC (2017) Mangrove dieback during fluctuating
sea levels. Sci Rept 7: 1680, https://doi.org/10.1038/s41598-017-01927-6.
Lu WZ, Chen LZ, Wang WQ, Tam NFY, Lin GH (2013) Effects of sea level rise on mangrove
Avicennia population growth, colonization and establishment: evidence from a field survey and
greenhouse manipulation experiment. Acta Oecol 49: 83–91.
Lugo AE (1980) Mangrove ecosystems: successional or steady state? Biotropica 12: 65–72.
Matos CR, Berredo JF, Machado W, Sanders CJ, Metzger E, Cohen MC (2020) Carbon and nutrient
accumulation in tropical mangrove creeks, Amazon region. Mar Geol 429:106317, https://doi.
org/10.1016/j.margeo.2020.106317.
MacKenzie RA, Foulk PB, Klump JV, Weckerly K, Purbospito J, Murdiyarso D, Donato DC, Nam
VN (2016) Sedimentation and belowground carbon accumulation rates in mangrove forests that
differ in diversity and land use: a tale of two mangroves. Wetlands Ecol Manage 24:245–261.
Macy A, Sharma S, Sparks E, Goff J, Heck KL, Johnson MW, Harper P, Cebrian J (2019)
Tropicalization of the barrier islands of the northern Gulf of Mexico: a comparison of herbivory
and decomposition rates between smooth cordgrass (Spartina alterniflora) and black mangrove
(Avicennia germinans). PLoS ONE 14: e0210144, https://doi.org/10.1371/journal.pone.
0210144.
Mafi-Gholami D, Zenner, EK, Jaafari A (2020) Mangrove regional feedback to sea level rise and
drought intensity at the end of the 21st century. Ecol Ind 110: 105972, https://doi.org/10.1016/j.
ecolind.2019.105972.
Manea A, Geedicke I, Leishman MR (2020) Elevated carbon dioxide and reduced salinity enhance
mangrove seedling establishment in an artificial saltmarsh community. Oecologia 192:
273–280.
Mangora MM, Mtolera MSP, Björk M (2014) Photosynthetic responses to submergence in man-
grove seedlings. Mar Freshw Res 65: 497–504.
Marchio DA, Savarese M, Bovard B, Mitsch WJ (2016) Carbon sequestration and sedimentation in
mangrove swamps influenced by hydrogeomorphic conditions and urbanization in Southwest
Florida. Forests 7:116, 10.30116.390/f706.
Maurer R, Tapis ME, Shor AC (2020) Exogenous root uptake of glycine betaine mitigates improved
tolerance to salinity stress in Avicennia germinans under ambient and elevated CO2 conditions.
FASEB J 34, https://doi.org/10.1096/fasebj.2020.34.s1.07221.
McCloskey TA, Liu K-B (2013) Sedimentary history of mangrove cays in Turneffe Islands, Belize:
evidence for sudden environmental reversals. J Coast Res 29: 971–983.
McIvor A, Spencer T, Möller I, Spalding M (2013) The response of mangrove soil surface elevation
to sea level rise. Natural Coastal Protection Series: Report 3. Cambridge Coastal Research Unit
Working Paper 42. The Nature Conservancy and Wetlands International.
McKee KL, Cahoon DR, Feller IC (2007) Caribbean mangroves adjust to rising sea level through
biotic controls on change in soil elevation. Glob Ecol Biogeogr 16: 545–556.
McKee KL, Rooth JE (2008) Where temperate meets tropical: multi-factorial effects of elevated
CO2, nitrogen enrichment, and competition on a mangrove-salt marsh community. Glob
Change Biol 14:971–984.
McKee K, Rogers K, Saintilan N (2012) Response of salt marsh and mangrove wetlands to changes
in atmospheric CO 2, climate, and sea level. In Middleton BA (ed), Global change and the
function and distribution of wetlands. Springer, Dordrecht, p 63–96.
Mukhopadhyay A, Payo A, Chanda A, Ghosh T, Chowdhury SM, Hazra S (2018) Dynamics of the
Sundarbans mangroves in Bangladesh under climate change. In Nicholls RJ, Hutton CW, Adgar
WN, Hanson SE, Munsur Rahman Md, Salekin M (eds), Ecosystem services for well-being in
deltas. Palgrave Macmillan, Cham, Switzerland, p 489–503.
222 D. M. Alongi

Murdiyarso D, Hanggara BB, Lubis AA (2018) Sedimentation and soil carbon accumulation in
degraded mangrove forests of North Sumatra, Indonesia. BioRxiv:325191, https://doi.org/10.
1101/325191.
Osland MJ, Enwright N, Day RH, Doyle TW (2013) Winter climate change and coastal wetland
foundation species: salt marshes vs. mangrove forests in the southeastern United States. Glob
Change Biol 19:1482–1494.
Osland M, Feher L, Griffith K, Cavanaugh K, Enwright N (2017) Climatic controls on the global
distribution, abundance, and species richness of mangrove forests. Ecol Monogr 87: 341–359.
Osland M, Day RH, Hall CT, Feher LC, Armitage AR, Cebrian J, Dunton KH, Hughes AR, Kaplan
DA, Langston AK, Macy A, Weaver CA, Anderson GH, Cummins K, Feller IC, Snyder CM
(2020) Temperature thresholds for black mangrove (Avicennia germinans) freeze damage,
mortality and recovery in North America: refining tipping points for range expansion in a
warming climate. J Ecol 108: 654–665.
Passos T, Penny D, Sanders C, De Franca E, Oliveira T, Santos L, Barcellos R (2021) Mangrove
carbon and nutrient accumulation shifts driven by rapid development in a tropical estuarine
system, northeast Brazil. Mar Pollut Bull 166:112219, https://doi.org/https://doi.org/10.1016/j.
marpolbul.2021.112219.
Peer N, Rajkaren A, Miranda NAF, Taylor RH, Newman B, Porri (2018) Latitudinal gradients and
poleward expansion of mangrove ecosystems in South Africa; 50 years after Macnae’s first
assessment. African J Mar Sci 40: 101–120.
Pérez A, Libardoni BG, Sanders CJ (2018) Factors influencing organic carbon accumulation in
mangrove ecosystems. Biol Lett 14:20180237, https://doi.org/10.1098/rsbl.2018.0237.
Phillips DH, Kumara MP, Jayatissa LP, Krauss KW, Huxham M (2017) Impacts of mangrove
density on surface sediment accretion, belowground biomass and biogeochemistry in Puttalam
Lagoon, Sri Lanka. Wetlands 37:471–483.
Principe SC, Augusto A, Costa TM (2018) Differential effects of water loss and temperature
increase on the physiology of fiddler crabs from distinct habitats. J Therm Biol 73: 14–23.
Proffitt CE, Travis S (2014) Red mangrove life history variables along latitudinal and anthropogenic
stress gradients. Ecol Evol 4:2352–2359.
Punwong P, Marchant R, Selby K (2013) Holocene mangrove dynamics in Makoba Bay, Zanzibar.
Paleogeogr Paleoclim Paleoecol 379–380: 54–67.
Quisthoudt K, Schmitz N, Randin CF, Dahdouh-Guebas F, Robert EM, Koedam N (2012)
Temperature variation among mangrove latitudinal range limits worldwide. Trees
26:1919–1931.
Rahu A, Das S, Banerjee K, Mitra A (2012) Climate change impacts on Indian Sundarbans: a time
series analysis (1924–2008). Biodivers Conserv 21: 1289–1307.
Rashid T (2014) Holocene sea-level scenarios in Bangladesh. Springer Briefs in Oceanography.
https://doi.org/10.1007/978-981-4560-99-3.
Reef R, Lovelock CE (2014) Historical analysis on mangrove leaf traits throughout the 19th and 20th
centuries reveal differential responses to increases in atmospheric CO2. Glob Ecol Biogeogr 23:
1209–1214.
Reef R, Winter K, Morales J, Adame MF, Reef DL, Lovelock CE (2015) The effect of atmospheric
carbon dioxide concentrations on the performance of the mangrove Avicennia germinans over a
range of salinities. Physiol Plant 154: 358–368.
Reef R, Slot M, Motro U, Motro M, Motro Y, Adame MF, Garcia M, Aranda J, Lovelock CE,
Winter K (2016) The effects of CO2 and nutrient fertilization on the growth and temperature
response of the mangrove Avicennia germinans. Photosynth Res 129: 159–170.
Rivera-Monroy VH, Danielson TM, Castañeda-Moya E, Marx BD, Travieso R, Zhao X, Gaiser EE,
Farfan LM (2019) Long-term demography and stem productivity of Everglades mangrove
forests (Florida, USA): resistance to hurricane disturbance. Forest Ecol Manage 440: 79–91.
Rogers K, Saintilan N, Heijnis H (2005) Mangrove encroachment of salt marsh in Western Port
Bay, Victoria: the role of sedimentation, subsidence, and sea level rise. Estuar 28:551–559.
9 Responses of Mangrove Ecosystems to Climate Change in the Anthropocene 223

Ruiz-Fernández AC, Agraz-Hernández CM, Sanchez-Cabeza JA, Díaz-Asencio M, Pérez-Bernal


LH, Keb CC, López-Mendoza PG, y Correa JB, Ontiveros-Cuadras JF, Saenz JO, Castellanos
JR (2018) Sediment geochemistry, accumulation rates and forest structure in a large tropical
mangrove ecosystem. Wetlands 38:307–325.
Saderne V, Cusack M, Almahasheer H, Serrano O, Masqué P, Arias-Ortiz A, Krishnakumar PK,
Rabaoui L, Qurban MA, Duarte CM (2018) Accumulation of carbonates contributes to coastal
vegetated ecosystems keeping pace with sea-level rise in an arid region (Arabian Peninsula). J
Geophys Res: Biogeosciences 123:1498–1510.
Saintilan N, Wilson N, Rogers K, Rajkaren A, Krauss KW (2014) Mangrove expansion and salt
marsh decline at mangrove poleward limits. Glob Change Biol 20: 147–157.
Saintilan N, Khan NS, Ashe E, Kelleway JJ, Rogers K, Woodroffe CD, Horton BP (2020)
Thresholds of mangrove survival under rapid sea level rise. Science 368:1118–1121.
Sasmito SD, Murdiyarso D, Friess DA, Kurnianto S (2016) Can mangroves keep pace with
contemporary sea-level rise? A global data review. Wetlands Ecol Manage 24:263–278.
Schettini CA, Asp NE, Ogston AS, Gomes VJ, McLachlan RL, Fernandes ME, Nittrouer CA,
Truccolo EC, Gardunho DC (2020) Circulation and fine-sediment dynamics in the Amazon
Macrotidal Mangrove Coast. Earth Surf Process Landform 45:574–589.
Schuerch M, Spencer T, Temmerman S, Kirwan ML, Wolff C, Lincke D, McOwen CJ, Pickering
MD, Reef R, Vafeidis AT, Hinkel J, Nicholls RJ, Brown S (2018) Future response of global
coastal wetlands to sea level rise. Nature 561: 231–234.
Seddon AWR, Froyd CA, Leng MJ, Milne G, Willis KJ (2011) Ecosystem resilience and threshold
response in the Galápagos coastal zone. PLoS ONE 6: e22376, https://doi.org/10.1371/journal.
pone.0022376.
Servino RN, de Oliveira Gomes LE, Bernardino AF (2018) Extreme weather impacts on tropical
mangrove forests in the Eastern Brazil Marine Ecoregion. Sci Tot Environ 628–629: 233–240.
Sippo JZ, Maher DT, Tait DR, Holloway C, Santos IR (2016) Are mangroves drivers or buffers of
coastal acidification? Insights from alkalinity and dissolved inorganic carbon export estimates
across a latitudinal transect. Glob Biogeochem Cycles 30: 753–766.
Sippo JZ, Lovelock CE, Santos IR, Sanders CJ, Maher DT (2018) Mangrove mortality in a
changing climate: an overview. Estuar Coast Shelf Sci 215: 241–249.
Sippo JZ, Sanders CJ, Santos IR, Jeffrey LC, Call M, Harada Y, Maguire K, Brown D, Conrad SR,
Maher DT (2020) Coastal carbon cycle changes following mangrove loss. Limnol Oceangr.
https://doi.org/10.1002/lno.11476.
Snedaker SC, Araújo RJ (1998) Stomatal conductance and gas exchange in four species of
Caribbean mangroves exposed to ambient and increased CO2. Mar Freshw Res 9:325–327.
Soper FM, MacKenzie RA, Sharma S, Cole TG, Litton CM, Sparks JP (2019) Non-native
mangroves support carbon storage, sediment carbon burial, and accretion of coastal ecosystems.
Glob Change Biol 25:4315–4326.
Stokes DJ, Healy TR, Cooke PJ (2010) Expansion dynamics of monospecific, temperate mangroves
and sedimentation in two embayments of a barrier-enclosed lagoon, Tauranga Harbour,
New Zealand. J Coast Res 26:113–122.
Suyadi S, Gao J, Lundquist CJ, Schwendenmann L (2019) Land-based and climatic stressors of
mangrove cover change in the Auckland Region, New Zealand. Aqua Conserv Mar Freshw
Ecosyst 29, https://doi.org/10.1002/aqc.3146.
Swales A, Reeve G, Cahoon DR, Lovelock CE (2019) Landscape evolution of a fluvial sediment-
rich Avicennia marina mangrove forest: insights from seasonal and inter-annual surface-eleva-
tion dynamics. Ecosystems 22:1232–1255.
Tamimia B, Wan Juliana WA, Nizam MS, Zain CRCM (2019) Elevated CO2 concentration and air
temperature impacts on mangrove plants (Rhizophora apiculata) under controlled environment.
Iraqi J Sci 60: 1658–1666.
Urrego LE, Correa-Metrio A, González C, Castaño AR, Yokoyama Y (2013) Contrasting responses
of two Caribbean mangroves to sea-level rise in the Guajira Peninsula (Colombian Caribbean).
Paleogeogr Paleoclim Paleoecol 370: 92–102.
224 D. M. Alongi

Walden G, Noirot C, Nagelkerken I (2019) A future 1.2  C increase in ocean temperature alters the
quality of mangrove habitats for marine plants and animals. Sci Tot Environ 690: 596–603.
Wallace RB, Baumann H, Grear JS, Aller RC, Gobler CJ (2014) Coastal ocean acidification: the
other eutrophication problem. Estuar Coast Shelf Sci 148: 1–13.
Wang W, Xiao Y, Chen L, Lin P (2007) Leaf anatomical responses to periodical waterlogging in
simulated semidiurnal tides in mangrove Bruguiera gymnorrhiza seedlings. Aquat Bot 86:
223–228.
Wang X, Piao S, Ciais P, Friedlingstein P, Myneni RB, Cox P, Heimann M, Miller J, Peng S,
Wang T, Yang H (2014) A two-fold increase of carbon cycle sensitivity to tropical temperature
variations. Nature 506:212–215.
Ward RD, Friess DA, Day RH, Mackenzie RA (2016) Impacts of climate change on mangrove
ecosystems: a region by region overview. Ecosys Health Sustain 2: e01211, https://doi.org/10.
1002/ehs2.1211.
Williamson GJ, Boggs GS, Bowman DM (2011) Late 20th century mangrove encroachment in the
coastal Australian monsoon tropics parallels the regional increase in woody biomass. Reg
Environ Change 11:19–27.
Wilson R (2017) Impacts of climate change on mangrove ecosystems in the coastal and marine
environments of Caribbean Small Island Developing States (SIDS). Caribbean Mar Clim
Change Rept Card: Sci Rev 61–82.
Woodroffe CD (2018) Mangrove response to sea level rise: palaeoecological insights from
macrotidal systems in northern Australia. Mar Freshw Res 69: 917–932.
Woodroffe CD, Rogers K, McKee KL, Lovelock CE, Mendelssohn IA, Saintilan N (2016)
Mangrove sedimentation and response to relative sea-level rise. Annu Rev Mar Sci 8:243–266.
Ximenes AC, Ponsoni L, Lira CF, Koedam N, Dahdouh-Guebas F (2018) Does sea surface
temperature contribute to determining range limits and expansion of mangrove in Eastern
South America (Brazil)? Remote Sens 10: 1787, https://doi.org/10.3390/rs10111787.
Yáñez-Espinosa L, Flores J (2011) A review of sea-level rise effect on mangrove forest species:
anatomical and morphological modifications. In Casalengo E (ed) Global warming
impacts- case studies on the economy, human health, and on urban and natural environments.
InTech, New York, p 253–276
Ye Y, Tam NFY, Wong YSLCY (2003) Growth and physiological responses of two mangrove
species (Bruguiera gymnorrhiza and Kandelia candel) to waterlogging. Environ Exp Bot 49:
209–221.
Ye Y, Tam NFY, Wong YSLCY (2004) Does sea level rise influence propagule establishment,
early growth and physiology of Kandelia candel and Bruguiera gymnorrhiza? J Exp Mar Biol
Ecol 306: 197–215.
Ye Y, Gu YT, Gao HY, Lu CY (2010) Combined effects of simulated tidal sea-level rise and
salinity on seedlings of a mangrove species. Hydrobiol 641: 287–300.
Yin P, Yin M, Cai Z, Wu G, Lin G, Zhou J (2018) Structural instability of the rhizosphere
microbiome in mangrove plant Kandelia candel under elevated CO2. Mar Environ Res 140:
422–432.
Yulianto E, Rahardjo AT, Noeradi D, Siregar DA, Hirakawa K (2005) A Holocene pollen record of
vegetation and coastal environmental changes in the coastal swamp forest at Batulicin, South
Kalimantan, Indonesia. J Asian Earth Sci 25: 1–8.
Roles of Mangroves in Combating
the Climate Change 10
Anupam Kumari and Mangal S. Rathore

Abstract

Mangrove ecosystem possesses great potential to mitigate the adverse impacts of


climate change and shows a higher range of ecological stability. Mangroves grow
luxuriantly in the harsh conditions such as high salinity, temperature, extreme
tides, strong winds and muddy and anaerobic soil. They evolved with well-
developed morphological, ecological as well as physiological adaptations that
make them more resistant and resilient to overcome the effect of adverse condi-
tion. Mangroves have unique properties in terms of structure and function such as
viviparous germination, well-developed aerial roots, lack of growth rings, adapt-
able to environmental changes and more efficient nutrient retention capabilities.
Various important ecosystem services such as nutrient cycling, carbon storage,
soil formation and ecotourism were provided by mangroves to support the
livelihoods of coastal societies in the tropical and subtropical area. Mangrove
ecosystem is the most productive ecosystem and stores more amount of carbon in
their above- and below-ground parts than that of terrestrial forest. This ecosystem
acts as a functional unit which involved plants, animals, microorganism and their
interaction with a surrounding environment. Mangroves perform several impor-
tant ecological functions such as coastal protection, carbon sequestration,
enriching biodiversity of coastal areas; promote land accretion and support
fisheries. They also have more economical significant. This chapter aims to
provide deep insight of mangroves in combating climate change.

A. Kumari · M. S. Rathore (*)


Academy of Scientific and Innovative Research, (AcSIR), Ghaziabad, India
Division of Applied Phycology and Biotechnology, CSIR-Central Salt and Marine Chemicals
Research Institute (CSIR-CSMCRI), Bhavnagar, Gujarat, India
e‐mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 225
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_10
226 A. Kumari and M. S. Rathore

Keyword

Mangroves · Carbon sequestration · Coastal ecosystem · Biodiversity · Climate


change

10.1 Introduction

Climate change is described as long-term identifiable variations in the state of


climate like atmospheric CO2, elevated temperature, precipitations, etc. (Korres
et al. 2016). Apart from this, several natural events that happened in Earth’s history
were also responsible for climate change. Hence, the concept of ‘climate change’
implies the changes in weather patterns happening across the last century and
triggered by anthropogenic activities that release greenhouse gasses such as CO2,
CH3, NO2, etc. (De Ollas et al. 2019). Virtually, global agriculture is frequently
influenced by environmental/climate change and these changes have become serious
global issues with their various, extensive and persistent effects. The problems of
food security imposed by climate/environmental changes/abiotic stresses have
become a major challenge for the researchers of plant science (Nikalje et al. 2018).
The word ‘abiotic stress’ involves various stresses occurring due to multifaceted
environmental conditions like salinity, drought, heat, cold, freezing, heavy metals,
UV and hypoxia (Hirayama and Shinozaki 2010). The responses of plants to abiotic
stress are essential and imperative for their existence as well as to combat with the
changing environment (Fancy et al. 2017). These responses occur at various levels,
such as morphological, physiological, biochemical, molecular and cellular
(Nakashima et al. 2009). For the survival, plant deploys various adaptive
mechanisms, including synthesis and accumulation of osmo-protective molecules,
up- and down-regulation of multiple genes, producing stress-responsive proteins and
hormones, etc. (Nakashima et al. 2009; Fancy et al. 2017).
Among the various abiotic stresses, salinity, water (drought and flood) and heat
stresses are the most severe threat to crop production in different parts of the world
(Qadir et al. 2008). Elevated salt concentration in the soil causes destruction of the
plant as well as soil and finally leads to desertification. The adverse impacts of
salinity on plants are associated with osmotic, ionic and oxidative stresses. In the
soil, higher concentrations of salt decrease soil water potential that imposes water
scarcity or osmotic strain. Salt stress causes more accumulation of Na+ and Cl and
reduction in the concentration of K+ and Ca2+ which in turn causes ionic stress and
nutrient imbalance.
Water stress arises due to either scarcity of water (drought) or excess of water
(flood). The most common occurring water stress is water deficit condition,
i.e. drought stress. Drought is a common word used to explain atmospheric or
weather phenomena and describe as a period without rainfall. Most commonly,
according to agricultural and physiological perspectives, drought stress happens
when the accessible water for plants in the soil declines due to decrease in soil
moisture content at a particular time and low rainfall (Keyvan 2010; Dai 2011). The
10 Roles of Mangroves in Combating the Climate Change 227

impacts of drought stress vary based on several aspects, like strength and extent of
the stress, genotype of the plant or growing stage, and moreover the imprint of the
earlier stress periods present on the plant (Fleta-Soriano and Munné-Bosch 2016;
Joshi et al. 2016). The most common symptoms of drought are loss of leaf turgor,
wilting, etiolation, yellowing and premature downfall of leaves (Jaleel et al. 2009;
Bhargava and Sawant 2013). Some uncommon symptoms are also observed like
crack of bark and twig, branch dieback, necrosis, impeded growth and ultimately
death to plants (Sapeta et al. 2013).
Flood stress reduces the supply of oxygen to the roots. Due to the lower supply of
oxygen critical functions of roots became out of order such as limited uptake of
nutrients and respiration. Moreover, in the mangrove environment higher
concentrations of methane are also linked to anoxia that affects the growth and
development of plants. Decreased survival, photosynthesis and productivity are the
major consequences of flooding (Ellison 2000). Stress arises because of flooding
causing stomata closure; consequently it causes negative regulation of the photosyn-
thetic apparatus leading to ROS generation in the chloroplast (García-Sánchez et al.
2007). During the waterlogged situation, mangroves recorded to be influenced by
limitation of gases diffusion such as CO2, O2, reduced light intensity under water and
accumulation of ethylene, etc. (Tomlinson 1986). Decreased light intensity in the
water during flood conditions is known to cause oxidative stress in plants.
With fluctuating climate scenario and raising the global temperature, heat stress
became a major issue for nowadays agriculture. Among the forever changing
components of the climate, the aberrant rise in the global mean temperature in
current years is the major factor for vulnerability of current agriculture (Chakraborty
et al. 2014). The deleterious effect of elevated temperature impacts growth, devel-
opment, metabolism and crop yield resulting in severe financial losses. Elevated
temperature curtails the lifespan of plants, enhances senescence and severely
influences the productivity. Higher temperature also tends to cause a modification
in membrane fluidity resulting in changes in membrane function (Barnabás et al.
2008).
Biotic stress occurs in the plants due to the interaction of living organisms like
especially bacteria, viruses, insects, nematodes, fungi and weeds. These biotic stress-
causing factors directly deprive the nutrients of their host, cause damage to plants,
resulting in less productivity or even plant death (Das et al. 2016). Moreover, water
deficiency, cell wall damage, mechanical wounding, osmotic stress and impairment
of plant metabolism also occur during biotic stress (Abood and Lösel 2003). They
affect photosynthesis processes like the rate of photosynthesis per leaf area is
reduced due to virus infection and reduction in leaf area occur because of infection
of chewing insects (Kathiresan and Bingham 2001). Due to the pre as well as
postharvest losses of the crop, biotic stress can become a major challenge for
nowadays.
228 A. Kumari and M. S. Rathore

10.2 Mangroves

Mangroves belong to a special plant group that grows luxuriantly at the interface
between land and sea of the tropical as well as subtropical areas (Kathiresan and
Bingham 2001). The word ‘mangrove’ refers to both specific types of plants as well
as their special types of habitat. These special types of habitats are known as tidal
forest/mangrove forest/mangal/swamp/wetland (Saenger 2002; Spaulding et al.
2010). Mangroves act as true ecotones due to possessing some characteristics of
both marine and terrestrial biomes (Alongi 2009). They have developed a lot of
distinct structural, functional and physiological adaptation like viviparous germina-
tion, aerial roots, lack of growth rings, etc. These adaptations provide the capability
to tolerate and thrive in very harsh conditions like high salinity, temperature, extreme
tides and strong winds as well as muddy and anaerobic soil (Alongi 2009). Man-
grove is one of the most productive ecosystems across the world and provides unique
ecological environments that accommodate a rich collection of species. They have
great importance in several fields like ecology, economy as well as social and offer a
significant array of ecosystem goods and services (Gopal and Chauhan 2006).
Mangroves evolved with well-developed morphological, ecological as well as
physiological adaptations that make them more resistant and resilient to overcome
the impact of adverse circumstance.

10.2.1 Spatial Distribution

Mangroves cover <1% of tropical and 0.4% of global forest areas (FAO 2007; Van
Lavieren et al. 2012). They mostly occur on all the continents found between the
Tropic of Cancer and Capricorn. Globally, mangroves occupy around 75% of
tropical coastal line and mainly confined between 30  C north and 30  C south
latitudes (Saenger 2002; Singh et al. 2012). The latitudinal distribution is restricted
by essential climate factors like the occurrence of extreme cold weather and aridity
(Osland et al. 2013; Saintilan et al. 2014). Moreover, the distribution and structural
development are further restricted by the availability of freshwater or rainfall even
within the areas with appropriate temperature (Osland et al. 2014; Alongi 2015).
Across the world, mangroves are recorded in 123 countries and presently total areas
occupied by mangroves are around 152,000 km2 (Spaulding et al. 2010). Almost
75% of mangroves are recorded in 15 countries, among them just 6.9% are protected
(Thomas et al. 2017). Mangroves belong to the South as well as Southeast Asia
constitutes the world’s utmost diverse and extensive groups of mangroves and
account for 41.1% of global mangroves. Indonesia serves as the major source of
mangroves and contributes more than 20% of the total mangroves (Hamilton and
Casey 2016). These plants establish a unique and rare type of ecosystem. However,
they are threatened as devastated almost five times faster than forests. For instance,
the most threatened regions of mangrove are known as in America (North as well as
Central) due to several activities like aquaculture, coastal development, hurricanes,
etc. (Van Lavieren et al. 2012). India contributes 3.1% mangroves of entire world
10 Roles of Mangroves in Combating the Climate Change 229

cover. Gujarat covered the second largest area of mangroves after Sundarbans in
India. At present, the term ‘mangrove’ involves eighty four species from twenty four
genera and sixteen families. In spite of this, just seventy species are considered as
real mangroves whilst others are mangrove associates (Thatoi et al. 2016). All
mangroves not only differ in the species composition, but also exhibit different
proportions of shrubs, trees, and so on. Different species of mangroves are not
distributed uniformly. An unique zonation of mangrove species is recorded. The
zonation is determined by several factors like height of the land, salinity and tide at a
certain level. Moreover, the capability to compete with other species is also a major
factor considered in zonation. Rhizophora and Xylocarpus are highly salt as well as
flood-tolerant mangrove, hence reported near to the shoreline as harsh conditions are
found there. Bruguiera and Sonneratia are reported less salt tolerant and observed
mainly in the inland area.

10.2.2 History

The word ‘mangrove’ is originated from the Guarani (official language of Paraguay).
During the initial phase of the 1610S the term was written like ‘mangrow’ deriving
from Spanish ‘mangle’ or Portuguese ‘mangue’, however, later phase of 1690S the
word ‘mangrow’ moved into an English word ‘mangrove’ by etymology. Several
terms are associated with mangrove, viz. mangrove ecosystem/mangrove forest
community/mangal. Many other terms are also used for mangrove forest like oceanic
rain forest, tidal forest, tidal swamp forest, mangrove swamp and coastland wood
land (Kathiresan and Bingham 2001; Van Lavieren et al. 2012; Hamilton and Casey
2016; Thatoi et al. 2016). For example, mangal or mangrove forest community and
mangrove ecosystem are formed by association of microbes and animals with plants,
respectively (MacNae 1969). It is advised that the term ‘mangrove’ must be signified
to exclusive mangrove species whilst the term ‘mangal’ to the forest community in
its place (Nabeelah Bibi et al. 2019). Mangroves have been discovered and studied
since the ancient period. First of all, mangrove was discovered by Nearchus in the
Persian Gulf during 325 B.C. In the late 325 B.C., Theophrastus (Greek philosopher)
also found the existence of mangrove plants and documented in the book ‘Historia
Plantarum’ (Kathiresan and Bingham 2001; Spaulding et al. 2010; Schneider 2011).
Both of them explained about Rhizophora tree that held up by roots as a polyp,
leaves, flowers, etc. (MacNae 1969; Spaulding et al. 2010). Subsequently, the
ancient Greeks know about three areas of mangroves, viz. Arabian Sea, Persian
Gulf and Red Sea during 323 B.C. (Schneider 2011). In the realm of information,
mangroves originated from Indo-Malayan areas and most of the mangrove species
are documented there across the world (Hamilton and Casey 2016).
Mangroves are extremely old, probably arising immediately after the first
angiosperms about 114 million years back (Duke 1992). The first genera of
mangroves evolves were most possibly Avicennia and Rhizophora that emerging
almost at the end of the Cretaceous period (Chapman 1976). The propagules and
seeds of mangroves possess a unique characteristic that supports them to float on the
230 A. Kumari and M. S. Rathore

water surface. Owing to this property, it was very simple/effortless for spreading of
mangroves species by water dispersion from India and East Africa to Central and
South America around 23–66 million years back.

10.2.3 Features of Mangroves

Mangroves have developed with unique characteristics as they occurred in the


transition zone, where low soil oxygen contents and daily rises and falls of the tide
and salinity are recorded. Consequently, to cope with this changing environment,
they have evolved with both halophytic and xeromorphic features. The most distinct
morphological feature of mangroves is associated with their root system, for
instance, the buttress roots of Heritiera and Xylocarpus, the pneumatophores of
Avicennia, Lumnitzera and Sonneratia, the root knees of Xylocarpus, Ceriops and
Bruguiera and the slit roots of Rhizophora (Kathiresan and Bingham 2001). These
specialized roots work like respiratory system and assist gas exchange in mangroves
growing in anaerobic as well as in saline soil. Numerous mangrove roots do not
penetrate deeply in the anaerobic rock layer. As an alternative, they produce plentiful
lateral roots to support them.
Another distinct morphological appearance was observed in the leaves of
mangroves. They have thick, leathery, wax-like, dark green (except Nypa fruticans),
elliptical shaped and moderate size leaves. Wax-like leaves help them to retain more
water content and prevent water loss, or stomata positioned in such a manner that
causes less evaporation (Kathiresan and Bingham 2001). Most commonly, the leaves
show dorsiventral symmetry; however, isolateral leaves are also observed in the
Phoenix paludosa and Sonneratia apetala (Das et al. 1995). The leaves are arranged
in adjusted decussate (bijugate) form in which less than 180  C angle kept between
each pair and proceeding pair. This type of arrangement prevents self-shading of the
leaves and makes branch system (Tomlinson 1986). Mangrove leaves contain six
kinds of stomata that only differ in the arrangement of guard and subsidiary cells. In
the majority of species, either external or external and internal both the surface of the
stomatal pore is covered with horn or beak-like cuticular outgrowth that lower the
stomatal transpiration (Das and Ghose 1993). Generally, bundle sheath fibres and
bundle sheath extensions are absent in the leaves. On the other hand, they have
extended tracheids lasting in the ends of the vein. Branched sclereids are well
developed and most abundant in Aegiceras, Aegialitis, Rhizophora and Sonneratia.
Perhaps, mechanical support of leaves is provided by sclereids. It is assumed that
sclereids and tracheids also participate in the storage of water (Tomlinson 1986).
Mangrove woods also possess the unique anatomical characteristics that are
explained by Tomlinson (1986). They lack growth rings or conspicuously anoma-
lous; for instance, in Avicennia. Thus, getting old is difficult. The wood has
exclusive features that make it possible to overcome the elevated osmotic potential
of seawater as well as the transpiration triggered by high temperatures (Kathiresan
and Bingham 2001). Several narrow vessels are going from end to end of the wood.
The density of these vessels ranges from 33 to 270 nm such as in Excoecaria and
10 Roles of Mangroves in Combating the Climate Change 231

Aegiceras, respectively. These vessels facilitate to generate high tensions in the


xylem because a little reduction in the diameter of vessels creates improperly more
rise in flow resistance (Scholander et al. 1964). Simple perforation plates are found
in the vessel elements that form the vessels. In spite of this, scalariform perforation
plates are seen in the member of Rhizophoraceae (expect Kandelia). Size as well as
the allocation of the vessels is the key factor that vigorously influencing the water
conduction via wood (Kathiresan and Bingham 2001). Movement of water occurs
most quickly by ring-porous woods in which the largest vessels are present in the
outermost growth layer. However, slowest water conduction occurs in the diffuse-
porous woods in which vessels are present more uniformly in size and allocations.
Most of the mangrove woods are diffuse-porous but only Aegialitis rotundifolia is a
ring-porous wood (Kathiresan and Bingham 2001).
Some special characteristics are also observed in the seeds and during the
seedling stage. Mostly, mangrove fruits are cigar-shaped, greenish in colour and
length ranges from 2 to 25 cm. Endosperm haustoria are made by Avicennia marina
at the time of early embryonic histodifferentiation. As soon as the growth phase is
started extra-ovular type of embryonic development is noted subsequently. Thus, the
mature seed is enclosed via a pericarp initiating completely from the wall of the
ovary. Starting to the last of histodifferentiation up to the maturation of seeds,
cotyledon cells turn into enormously vacuolated and have more quantity of soluble
sugars and serve as major nutrient reserves in the mature seeds. Embryos require
abundant oxygen to grow since the waterlogged soil is oxygen deficient. Fertilized
seeds develop directly into seedlings even though attached to the parent tree. The
phenomenon is known as vivipary. Just after germination, the seedling grows to
form a propagule. This propagule falls over the water surface and floats till finding a
suitable place to grow well (Miththapala 2008).

10.3 Factor Affecting the Growth of Mangroves

Climate change imposed severe threats to mangrove ecosystems. The climate change
factors involve changes in sea level, temperature, atmospheric CO2 concentration,
precipitation, storms, high water events, circulation patterns of ocean as well as the
health of functionally associated adjoining ecosystems (Gilman et al. 2008). These
factors most commonly influence the growth of mangroves. However, mangroves
show resistance and resilience to mitigate the potential impact of climate change
(Kandasamy 2017).

10.3.1 Temperature

Global warming is an important challenge for threatening the integrity of the


mangrove ecosystem. Mean global temperatures are expected to increase very
quickly (1.1–6.4  C) in the twenty-first century (Solomon et al. 2007). The warmer
environment will create several injuries and possibilities for mangroves as well as
232 A. Kumari and M. S. Rathore

salt marshes plants. Due to temperature variations such as average temperature and
extremely low-temperature like frosts in North America, mangrove exhibits complex
dynamics at their latitudinal distribution (Cavanaugh et al. 2014). It is very tough to
predict the exact impact of elevated temperature. However, the poleward movement
of a few species is known in response to high temperatures (Saintilan et al. 2014). It
is expected that elevated surface temperature influences the mangroves in various
ways like changes in the composition of species, phenological patterns (for example,
flowering and fruiting timing), photosynthesis, respiration, microbial decomposition
and in the case when the temperature does not go beyond upper threshold enhances
the productivity of mangroves. Moreover, higher temperature causes expanding of
the ranges of mangroves in higher latitudes where other factors such as appropriate
physiographic conditions and supply of propagules are abundant and temperature
acts only by limiting factor (Field 1995; Ellison 2000). The majority of the mangrove
show maximum density of the shoot when the average air temperature reaches 25  C
whilst the production of leaves stops when the temperature falls below 15  C
(Hutchings and Saenger 1987). The optimal leaf temperature for photosynthesis is
achieved when leaf temperature is assumed to be in the range from 28 to 32  C. The
photosynthesis ceases when the temperature of the leaf is between 38 and 40  C
(Andrews et al. 1984).

10.3.2 Atmospheric CO2 Concentration

The level of atmospheric CO2 has enhanced by 35% from the pre-industrial period.
The CO2 concentration was 280 ppm (parts per million by volume) in 1880 and it has
reached 379 ppmv in 2005 as well as the acidity of the ocean has also increased by
25% (Solomon et al. 2007). In spite of large difference, each model forecast a further
enhancement in the level of CO2 by the century’s end. Someone has anticipated that
the CO2 level will be either double or even triple than of the present level (Van
Lavieren et al. 2012). It will be tough to envisage the responses; however, the rate of
photosynthesis and respiration, nutrient availability, water use efficiency and salinity
will probably change (Ball et al. 1997). Inside the estuaries, the patterns of species
are more probable to vary depending on the interactive impacts of change in CO2
concentration, sea level, temperature as well as weather patterns to the species-
specific responses. However, there will possibly be few or even no changes in the
production of the canopy (Alongi 2002, 2008). Higher CO2 conditions are likely to
improve the mangrove growth when the gain of carbon is restricted by demand for
evaporation of the leaves however, not at the time it is restricted by root salinity.
There is no proof that higher CO2 concentration will enhance the growth of man-
grove species in a higher range of salt concentration. Although all species may not
react in the same way and other environmental factors like salinity, temperature,
nutrient quantities as well as the hydrological regime can affect the response of
mangrove to the elevated level of atmospheric CO2 (Field 1995). The impact of
increased CO2 on mangrove is not well known and scarcity of research is found in
this area. Enhanced concentration of CO2 could modify the ability of competition,
10 Roles of Mangroves in Combating the Climate Change 233

hence changing the composition of the community along gradients of salinity—


humidity (Ball et al. 1997).

10.3.3 Sea Level

Sea level rise is one of the major consequences of global warming. During the
twentieth century, 12–22 cm rise of sea level has documented (Holgate and
Woodworth 2004; Thomas et al. 2004). Many climate models have anticipated a
hastened rate of rising sea levels in the upcoming decades (Church and White 2006;
Solomon et al. 2007). The range of rising in global sea level projection is
0.18–0.59 m as of 1980–1999 to the last of the twenty-first century. Latest findings
on the acceleration of changes in global sea levels signify that the upper level of
projections is expected to happen (Church and White 2006). The mangroves will
face serious threats in the future due to sea level rise (Field 1995; Gilman et al. 2008).
Expectations of loss of mangroves range from 30% to even extinction (Duke et al.
2007; IPCC 2007). Many physical, as well as biological factors, determine the extent
of mangrove loss. According to the rates of sea level rise (SLR) in addition to vertical
accretion, mangroves can move towards landward. SLR and vertical accretion both
are interconnected to slope and space available at the edge of landward. Adaptation
to the changes in sea level, range of tide, sediment supply, etc. mainly depends on
specific species (Alongi 2002, 2008). For establishing mangroves in the new regions
several factors are needed like appropriate hydrology, the composition of sediment,
available waterborne seedling as well as potential to compete with other species of
plants (Krauss et al. 2003). If sea level rise happens slowly enough and sufficient
space exists for expansion, then mangrove can adapt to changes in the sea level.
Mangroves would be likely to migrate to landward as the sea level rises. Further-
more, with the rise in sea level, the width of the mangrove system can reduce. The
capability of landward migration of mangrove is dictated by local environments like
infrastructure (dikes, roads, sea walls, urbanization) as well as topography (steep
slope). It has been assumed that the mangroves found in riverine regions with the
closely packed forest of mangroves are less prone to sea level rise. Mangroves
located in small islands, scarcity of rivers, coastal development, ground water
extraction, underground mining, tectonic movement and steep topography are
assumed to be most vulnerable to sea level rise. (Ellison and Stoddart 1991).
The four main factors describing below are responsible for the resistance and
resilience of mangroves in response to changes in sea level: (1) the vulnerability of
mangroves is determined by the rate of sea level rise in proportion to the sediment
surface of mangroves (Cahoon et al. 2006; Cahoon and Hensel 2006; Gilman et al.
2008). Change in sediment elevation occurs at different rates in different zones of
mangrove vegetation, thus the composition of mangrove species affects the response
of mangroves (Krauss et al. 2003; Rogers et al. 2005; McKee et al. 2007). In
response to changing sea levels, certain zone shows more resistance and resilience.
Moreover, different species of mangroves also needed different time periods for
colonizing in the new habitat, which happens with change in relative sea level. The
234 A. Kumari and M. S. Rathore

fast colonizing species may compete rapidly with slower colonizing species and
become more dominant (Lovelock and Ellison 2007). (2) The physiographic loca-
tion involving differences in the slope of land previously and presently occupies by
mangroves as well as obstacles found during their landward migration influencing
the resistance and resilience of mangroves (Gilman et al. 2008). Lastly (3) the
collective impacts of all stress factors influence the resistance and resilience ability
of mangroves. It is not anticipated that mangrove responds according to the
assumptions of Bruun rule (a model used for prediction of erosion of beaches) as
sediment budget processes of mangrove are differ from beaches. In addition, on a
small scale and site-specific estimations inaccurate outcomes are generated by
predictive models of coastal erosion (Bruun 1988; List et al. 1997; Pilkey and
Cooper 2004). The accelerated rise of sea level likely to stimulate the replacement
of salt marsh plants by mangroves when mangroves and salt marsh plants co-exist. In
the many estuaries of SE Australia, the replacement of salt marsh plants by
mangroves was observed by Saintilan and Williams (1999).

10.3.4 Precipitation

Global rainfall is expected to enhance almost 25% by 2050 due to climate change.
Though extent and direction of change in precipitation will not be uniform, they
would differ according to season and spatial distribution (Houghton et al. 2001;
Solomon et al. 2007). Precipitation is more likely to increase in higher latitudes,
while in most of the subtropical areas (mainly at the poleward margins of subtropics)
precipitation is likely to decrease (Solomon et al. 2007). According to the recent
prediction of the IPCC, substantial enhancement in precipitation is likely to happen
in northern and central Asia, northern Europe as well as North Eastern and South
eastern America. They have also anticipated drought conditions in southern Asia,
southern Africa, the Mediterranean and in the Sahel (Solomon et al. 2007). The
trends for long-term precipitation were not observed in the other parts. Mangrove
growth and spatial distribution are expected to affect by changes in precipitation
patterns (Field 1995; Ellison 2000). Excess of precipitations especially overflow of
freshwater can change patterns of salinity which in turn affects vegetation of the
coastal ecosystem. Primarily based on the links found in the rainfall trends and
habitat condition of mangrove, low rainfall and more evaporation will cause higher
salinity, lower seedling growth and survival, less net primary productivity, decreas-
ing biodiversity. Moreover, due to the formation of hypersaline flats from upper tidal
zones, a significant reduction in mangrove area will also be observed (Field 1995;
Duke et al. 1998). The area having less precipitation will contribute less freshwater
to mangroves resulting in more salinity. For instance, Heritiera fomes (‘Sundari’)
and Nypa fruticans species of mangroves are continuously disappearing from the
Sundarbans regions of India. It happens because the supply of freshwater is
completely shut off due to the siltation of Bidyadhari (Mitra et al. 2009). Further-
more, highly salt-tolerant species like Ceriops (Rhizophoraceae family) occupied the
place of Heritiera fomes and Nypa fruticans (VYAS 2012). Recently, less
10 Roles of Mangroves in Combating the Climate Change 235

salt-tolerant species of mangroves are continuously vanishing from the entire Indian
mangrove ecosystem (except Andaman and Nicobar islands) and more salt-tolerant
species such as Avicennia marina are becoming dominant. Species of
Rhizophoraceae (True mangrove) were dominant in Muthupet around 150 years
back. However, they are extinct spatially these days (Kandasamy 2017). In the
1950s, mangrove wetlands of Godavari had around 90% populations of tall and
closely packed tress of Avicennia officinalis, Lumnitzera racemosa and Excoecaria
agallocha. However, their population have reduced to 37% and are replaced by salt
marsh bushes of Suaeda nudiflora and S. maritima. The reductions in the consis-
tency and quantity of freshwater supply to the mangroves are the main cause for
variations in the species composition of mangroves (Kandasamy 2017).
The salt concentration in the mangrove tissue increases with increasing soil
salinity while the water content decreases concomitantly, which leads to low pro-
ductivity (Field 1995). The sulphate content of seawater will also increase with
elevated levels of salinity which would further stimulate peat decomposition and
cause the vulnerability of mangroves in response to changes in sea level (Snedaker
1993, 1995). At the same time, more rainfall causes enhanced growth and biodiver-
sity of mangroves and increases in the area of mangrove (Duke et al. 1998). More
diverse, taller and more productive mangroves are located on the shorelines having
more rainfall than the shorelines having less rainfall, recorded in the majority of the
locations worldwide along with Australia (Duke et al. 1998; Ellison 2000).
Mangroves inhabiting the area of more rainfall have higher productivity and diver-
sity most likely because of less exposure to salinity and sulphate content as well as
more supply of nutrients and fluvial sediment (McKee 1993; Ellison 2000). More
intense rainfall is expected to affect many physical processes and erosion in the tidal
wetlands and catchments. In the arid regions, less rainfall combined with more
evaporation which possibly leads to reduction in the area of mangroves, less
diversity and conversion of the landward area to the barren hypersaline surface
(Snedaker 1995).
Due to the variations in local weather, floods and droughts occur which may
possess a significant impact on vegetation. The vegetation of salt-tolerant species is
promoted during the duration of low to moderate flooding whilst catastrophic
flooding causes mortality of plant and subsequently, the establishment of the annual
plants occur. It is observed in the Nueces Estuary of Texas (USA) in the study of
semi-arid, subtropical salt marsh (Forbes and Dunton 2006). Variations in the pattern
of rainfall may promote the shift of vegetation at the distribution boundaries. A
positive relationship was observed between variables of rainfall and landward
migration of mangroves during the study of more than 32 years on patterns of
rainfall and geographical distribution in the Moreton Bay mangrove forest (South-
east Queensland, Australia) (Eslami-Andargoli et al. 2009). Many other factors like
geomorphology, local weather and disturbance can alter the patterns and rates of
mangrove expansion. There are some more explanations regarding landward migra-
tion of mangroves towards salt marsh such as modified patterns of tidal, enhanced
sedimentation and nutrient contents and anthropogenic disturbance (Saintilan and
Williams 1999).
236 A. Kumari and M. S. Rathore

10.3.5 Storms

According to IPCC (Intergovernmental Panel on Climate Change) projections on


climate change, peak intensities of wind and precipitation of the tropical cyclone
along with tropical cyclone mean are expected to increase in the twenty-first century
due to climate change (Houghton et al. 2001; Solomon et al. 2007). Moreover, the
warmer climate is anticipated to produce stronger tropical cyclones having highly
intense precipitation and strong speeds of winds (Solomon et al. 2007). The intensity
of the effect will be depending upon strength, extent, size and frequency of storms.
There are several effects of storms, viz. defoliation, up-rooting, the mortality of trees
as well as enhance stress, which arises due to a changing sediment elevation of
mangrove. Sediment elevation occurs because of erosion, compression and deposi-
tion of soil as well as peat collapse (Cahoon and Hensel 2006; Piou et al. 2006). The
recovery rate of damage caused by the storm may be very slow. The heights of storm
surge are expected to rise if the occurrence of low pressure and strong wind
enhances. This happens due to the occurrence of more frequent and severe storms
because of the changing climate (Church et al. 2001; Solomon et al. 2007). The more
frequent and intense storms have great potential to mangroves damaged by means of
defoliation and mortality. The area having a bulk mortality of trees along with fewer
surviving plantlets and trees would suffer from the permanent conversion of the
ecosystem. It occurs since recovery via recruitment of seedling possibly will not
happen as a result of changes in the sediment elevation and associated changes in
hydrology (Cahoon et al. 2003). Other natural incidents like a tsunami (Indian Ocean
tsunami on 26th December 2004) may also cause serious impairment to mangroves
(Danielsen et al. 2005; Kathiresan and Rajendran 2005; Dahdouh-Guebas et al.
2006). Several models have projected poleward storm shifts and highly intense
(but less) tropical storm paths through a little latitudinal degree in both hemispheres
(Geng and Sugi 2003; Yin 2005; Bengtsson et al. 2006). Variations in the average
precipitation are expected to have little impacts on vegetation as compared to
variations in the occurrence of extreme events. The elevation of wetland may also
change due to episodic incidents like cyclones, hurricanes, lightning strikes, storms,
storm surges and flooding associated with it, strong wind as well as excess flow of
fresh water (Whelan et al. 2005; Cahoon et al. 2006). As the frequencies and
intensity of these episodic incidents are predicted to rise in linked with climate
change and their effect can increase on wetland resilience in the future (Christensen
et al. 2007; Kundzewicz et al. 2007). Besides the rise in sea level, the intensity of
storm surges is projected to enhance 5–10% by 2050 which may also serve as a flood
for mangroves. The species composition and health of mangrove may be affected by
storm because of changes in sedimentation, inundation, recruitment and salinity.
Avicennia and Sonneratia are highly susceptible as compared to Rhizophora species
in response to storm surge. It happens because the species of Rhizophora have slit
roots that remain standing in the above rise in sea level whilst pneumatophores roots
are mainly submerged during the rise in sea level and found in the Avicennia and
Sonneratia species. Additionally, in the area of mangroves trapping of sediment and
accumulation of peat are facilitated by slit roots. Storms and tropical cyclones occur
10 Roles of Mangroves in Combating the Climate Change 237

most frequently in the Bay of Bengal, and east coast of India is severely affected than
the west coast. Though, Rhizophora spp. works as protecting force against the
natural calamity in India. In general, species of Avicennia regenerate after cyclones
in these regions. Thus, Indian mangroves exhibit resistance to cyclones.

10.4 Response of Mangroves to Environmental Stresses/


Climate Change

Mangrove ecosystems are the rarest and toughest ecosystems on the Earth. Their
natural environment is comprised of abiotic stress (salt, flood, heavy metal, low
oxygen) and biotic hindrance of different nature and extent (Das et al. 2016).
Mangroves have developed the distinctive capability to combat with this extraordi-
nary magnitude and type of stress due to their continuous exposure. However, the
susceptibility and responses of mangrove against climate change will be extremely
impacted by human interference (Das et al. 2016).

10.4.1 Soil Characteristics

The soil characteristics exhibit a significant effect on the growth and nutrition of
mangroves.
Electrical conductivity, cation exchange capacity, pH and saltiness are some
important characteristics of the soil (Kathiresan and Bingham 2001). However, the
concentrations of nutrients seem to be the most important factor. Mangrove
ecosystems are well-balanced and most efficient sink of nutrients along with the
actual import of dissolved nutrients like nitrogen, silicon and phosphorus
(Kathiresan and Bingham 2001). The availability of nutrients can limit growth,
development and ultimately the productivity of several mangrove forests. Moreover,
different concentration of nutrients may also affect the distribution of species and
their competitive abilities (Chen and Twilley 1998). The limitation of nutrients can
differ according to individual habitats of mangrove. For instance, seedlings of
Rhizophora apiculata perform significantly well in the cultivated area enriched
with potassium (Kathiresan and Bingham 2001). Most commonly, mangroves
found in the soil with low nutrient carbonate having the limitation of phosphorous.
Available phosphorous is likely to be bound with calcium and efficiently retaining it
inside the sediments. Almost seven and three-fold enhancements in the rate of stem
elongation and in the leaf area were observed, respectively, in the seedling of
Rhizophora mangle enriched with phosphorous during mesocosm and field
experiments. No such types of responses were detected in the case of nitrogen
addition (Koch and Snedaker 1997). Similarly, the growth of dwarf R. mangle in
the Belizean mangal is limited by low availability of phosphorous (Feller 1995).
They also encourage the growth of tough and long-lasting leaves known as
sclerophyll. The development of sclerophylls possibly will be an adaption for
conserving nutrients in the habitats of oligotrophic. Other mechanisms may also
238 A. Kumari and M. S. Rathore

be found in the mangroves to retain useful nutrients like most abundant concentra-
tion of nitrogen and potassium have recorded in the mature and photosynthetically
active leaves than senescent leaves. It appears as an outcome of the translocation of
nutrients out of aged leaves into other parts of plants prior to the leaves fall. Due to
the damage of the mangrove ecosystem, its capability of nutrient retention may
compromise. For instance, substantial loss of nitrogen and phosphorous have
detected in the severely damaged area of mangroves in North Queensland (Kaly
et al. 1997). This may happen because of a dramatic decline in the burrows density
and a decrease in the population of crab. The disturbance impact will vary according
to habitat and will be determined by the characteristics of the sediments and patterns
of flow on each site. As no differences was recorded in the soil nutrient concentration
of healthy and degraded areas of mangroves in Pedada Strait (Triwilaida Intari
1990). The distinctive features of mangrove sediments are sulphides that affect the
distribution of mangroves. The concentration and distribution of the sulphides are
controlled by the mixing of tidal, bioturbation and also via mangrove themselves.
Most often less reduction of soils is observed close to the aerial roots of some species
that directs low sulphide content. Moderately reducing soil with fewer sulphide
contents has recorded in the area dominated by R. mangle in neotropical Florida
mangrove forest. On the other hand, highly reducing soil with more sulphide
contents area has dominated by Avicennia germinans (McKee 1993). Unexpectedly,
the same pattern of soil characteristics has not been detected in a similar mangrove
ecosystem residing in Brazil. As strongly reducing soil with high sulphide contents
are found in the area dominated by R. mangle. At the same time, soil sulphide
content of A. germinans is extremely variable due to the change of rhizosphere from
oxygenated to anoxic conditions (Lacerda et al. 1995). The more exchangeable trace
metals are also found in the soil of Avicennia (Lacerda et al. 1995). In young forests,
sulphate reduces very slowly into sulphide that leads to high nutrient contents and
low toxicity of sulphide (Alongi et al. 1998). Higher contents of sulphide may
damage seedlings of mangrove, cause closure of stomata, decrease gas exchange,
decline growth and ultimately mortality increases. The rate of sulphate reduction can
increase by disturbance and heavy organic input. Clearance of mangrove ecosystem
or gaps in the formation of the canopy may alter the physical as well as chemical
properties of the underlying soil. This led to anaerobiosis condition and also
increases sediment sulphide contents. Most often, soils of mangroves are deficient
in oxygen or without oxygen. So mangroves have distinct features to combat with
this condition. The presences of pneumatophores (breathing roots) are the most
prominent feature. This is above-ground roots which are occupied with spongy
tissue. Several little holes are found in the bark that allows oxygen to be transported
to the root parts.

10.4.2 Salinity

Salinity of coastal areas is controlled by many factors like climate, topography,


hydrology and tidal flooding. These factors strongly influence the growth and
10 Roles of Mangroves in Combating the Climate Change 239

productivity of the mangrove ecosystem. They may also affect competitive abilities
of the species. In general, mangroves are facultative halophytes that tolerate both
fluctuating and higher salt concentrations. Some researchers have also classified
mangrove as obligate halophytes. Many species of mangrove attain optimal growth
at 5–25% salinities of seawater. The growth of mangroves occurs in more or less
waterlogged soil and water with varied concentration of salinity and may be equal to
salt concentration of open sea. Soil salinization is the most severe environmental
stresses mainly in the areas of arid, semi-arid and mangrove ecosystems (Das et al.
2016). Higher salt accretions lower down the soil water potential that makes more
difficult for plant to obtain nutrients and water from soil. Hence, water deficit
condition arises in plants as a result of salinity stress that lead to physiological
drought. The most deleterious impacts of salinity on plants include ionic stress,
osmotic stress, oxidative stress, enhanced ROS production, reduced growth, photo-
synthesis and ultimately low productivity (Doganlar et al. 2010; Kosová et al. 2013).
Salt-tolerant mechanisms of mangroves are alike to those in glycophytes. However,
most likely mangrove may sequester or exclude salt ions more effectively (Jithesh
et al. 2006).
Mangroves developed various adaptive mechanisms like ion sequestration, salt
accumulation and salt excreting to cope up with these adverse conditions. The leaves
of mangroves tend to be thicker and smaller in response to salt stress. Small leaf
releases more heat by way of convection thus enhances cooling of the leaf.
Sonneratia alba and Sonneratia lanceolata (closely related species) showed inter-
specific differences in salinity tolerance in association with differences in their
distribution besides natural seasonal salinization (Ball and Pidsley 1995). Sonneratia
alba may grow in 100% salinities of sea water while Sonneratia lanceolata grow in
50% salinities of sea water. Moreover, vivipary characteristics of mangroves also
play a critical role in imparting enhanced salinity stress tolerance (Zheng et al. 1999).
Four genera of mangrove (Rhizophoraceae family) are known to have vivipary
characteristics, viz. Bruguiera, Ceriops, Rhizophora and Kandelia (Tomlinson
1986). It has believed that vivipary feature of mangroves helps in germination during
saline conditions. Typical salt glands structures are present in the species of Acan-
thus, Avicennia, Aegialitis and Aegiceras that remove excess salt concentration and
imparting more salt tolerant. Furthermore, the analogous structure of salt gland is
found in the species of Conocarpus and Laguncularia (Tomlinson 1986). Structures
similar to ‘Pimples’ or ‘tiny bumps’ resembling a certain extent with salt glands are
present in the petioles of the leaves L. racemosa and C. erectus (Tomlinson 1986).
Salt leaks from salt glands of Avicennia and Aegiceras have deposited as a crystal of
salts in their leaves. The species of Avicennia, Sonneratia, Rhizophora and
Xylocarpus also accumulates salt in the bark of roots and stem (Scholander 1968).
One more distinct property of mangroves is the development of the succulent
structure by leaves thickening to store more water content (Suárez and Sobrado
2000). For instance, the thickness of the leaves as well as water content may increase
in Laguncularia racemosa during salinity stress. The enhanced water content diluted
the absorbed salt, thus reduced the damage caused by salinity stress to some extent
(Sobrado 2005). Due to the development of stress-induced succulence, the rate of
240 A. Kumari and M. S. Rathore

photosynthesis may also enhance by an increment in the surface of the internal leaf
for gas exchange. The wax present in some mangroves leaves also acts as protective
features and contributes less transpiration in them than other species lacking this
trait. Mangroves possess numerous anatomical and morphological features that are
much alike to higher water use efficiency features of terrestrial xerophytes. These
features allow mangroves to grow in high saline or physiological dry condition
devoid of any visible adverse impacts of severe water stress (Parida and Jha 2010).
Many genera of mangrove retain sunken stomata and thickly layered of epidermis
having waxy cuticle. The species of Rhizophora, Avicennia, Ceriops and Bruguiera
have more prominent sunken stomata beneath the epidermis (Miller et al. 1975). The
waxy cuticles are covered with various shaped hairs such as stellate scales in the
species of Heritiera and tricellular peltale stellate hairs in Avicennia (Miller et al.
1975). To combat with the adverse impacts of ROS production mangroves have
developed strongly regulated mechanisms involving ROS scavenging (enzymatic
and non-enzymatic) pathways (Das et al. 2016). During salt stress, the steep incre-
ment in the SOD level was documented in the Bruguiera parviflora and
B. gymnorrhiza (Takemura et al. 2000; Parida et al. 2004). Around 8.1 times
increment in the SOD activity was observed in B. gymnorrhiza when subjected to
salinity stress (500 mM for 9 days) (Takemura et al. 2000). In A. marina higher SOD
as well as POX activity was detected under saline conditions (Cherian et al. 1999).
The activity of catalase enzyme was varied in the species of Bruguiera under salt
stress. Reduce catalase activity was recorded in the Bruguiera parviflora when
exposed to salt stress (400 mM NaCl) (Jithesh et al. 2006). During salt stress,
increased activity of antioxidant enzymes (SOD and POD) was also detected in
the Aegialitis rotundifolia, Bruguiera gymnorrhiza, Xylocarpus mekongensis,
Xylocarpus granatum, Heritiera fomes, Phoenix paludosa (Dasgupta et al. 2012).
Higher activities of antioxidant enzymes signifying their protecting role in response
to oxidative stress (Atreya et al. 2009; Foyer and Shigeoka 2011). Apart from
enzymatic antioxidants, non-enzymatic antioxidants (ascorbic acid, carotenoids,
tocopherol and glutathione) also perform an essential function in ROS scavenging
under stress conditions.

10.4.3 Temperature

Climatic factors like temperature strongly impact diversity and the global distribu-
tion of mangrove forests (Duke et al. 2017). Generally, in the tropical wet climate,
mangrove ecosystems are most diverse, abundant and highly productive. Chilling
and freezing temperatures may cause reduced biomass, less productivity and man-
grove mortality. Increased temperature is also likely to change the expansion of
latitudinal limit, diversity and composition of the community (Duke et al. 2017). The
poleward expansion of mangroves occurs most probably in response to the tempera-
ture and sea level rise (Reid and Beaugrand 2012). Elevated temperatures are
expected to give an advantage to the Pacific Islands. As mangrove diversity is
anticipated to increase at higher latitudes (at present, occupied by the only species
10 Roles of Mangroves in Combating the Climate Change 241

of Avicennia) due to warming (Burns 2001). Warming is estimated to facilitate the


expansion of mangroves into salt marshes of the Pacific Islands (Burns 2001).
Thermal tolerance is varied in the mangroves species of China. Based on the
temperature tolerance, Li and Lee (1997) classified mangrove species of China
into three groups: (1) thermophilic eurytopic species (e.g. Rhizophora stylosa,
Bruguiera gymnorrhiza, B. sexangula, Acrostichum aureum and Excoecaria
agallocha), (2) thermophilic stenotopic species (e.g. R. apiculata, R. mucronata,
Pemphis acidula, Nypa fruticans and Lumnitzera littorea) and (3) cold resistant
eurytopic species (e.g. Aegiceras corniculatum, Avicennia marina and Kandelia
candel). Changes in the soil temperature are likely to increase at the same rate and
magnitude as that of sea surface temperature. However, changes in temperature of
soil are mostly very less than that of atmospheric temperature, because saturated soil
has a higher ability to retain heat. So, mangroves are not expected to be much
adversely affected even the temperature of soil increases. The rate of bacterial
growth and multiplication also increases due to the elevated temperature of sediment
and ultimately enhances the rate of regeneration and nutrient cycling. The leaf
inclination is most essential for the regulation of temperature. For instance, 10  C
differences were found in the leaf air temperature on a clear day when the leaf held
horizontally. Moreover, photosynthetic gas exchange is most sensitive to vapour
pressure deficits and temperature of leaf air. Mangroves apparently respond to
elevated temperature; however, the exact temperature is not known at which their
functionality plateaus or they die. Most of the species show maximum photosynthe-
sis at temperature 30  C or below. In many species rate of CO2 assimilation
decreases either gradually or sharply when the temperature rises from 33 to 35  C
(Ball and Sobrado 2002). Photosynthesis is often decline in exposed leaves because
of photo-inhibition. Only elevated temperature is expected to enhanced photosyn-
thesis, respiration and growth.

10.4.4 Metal and Organic Pollutant

Heavy metals are major environmental pollutants, their toxic effects exert severe
problem with the growth and development of plants (Zornoza et al. 2002; Yadav
2010). Some important metals like iron (Fe), nickel (Ni), manganese (Mn), zinc
(Zn) and copper (Cu) are needed for proper growth, development and functioning of
plants. However, the higher concentration of these metals exhibits the toxic impact
on plants (Rai et al. 2004). Even though highly selective transporters are found in the
plants, but some other metals like metalloid arsenic (As), aluminium (Al), lead
(Pb) and cadmium (Cd) are also taken by plants that may cause toxicity even at
low concentration (Sebastiani et al. 2004). The most adverse effect of heavy metals
on plants is lipid peroxidation that causes deterioration of membrane. Moreover, due
to the heavy metal stress, reduction in the net photosynthesis occurs through the
damage in the photosynthetic electron transport system and photosynthetic metabo-
lism (Vinit-Dunand et al. 2002). ROS production is the major response of plant
during heavy metal stress. However, mangroves have developed great ability to
242 A. Kumari and M. S. Rathore

mitigate the deleterious effects of toxicity of heavy metal stress. They also have well-
developed anti-oxidative defence system to combat with the adverse impacts of
heavy metals caused oxidative damage. Higher activities of both enzymatic (APX,
GPX, GR, SOD) and non-enzymatic antioxidants in heavy metal stressed plants
have suggested their role in adaptive mechanisms (Verma and Dubey 2003; Yadav
2010). Due to the closeness of mangrove habitats with industrialized areas and
population centres, they have continuous exposure/inputs of heavy metals. The
sediments of mangroves also exhibit significant contamination of heavy metals.
Several studies showed that mangrove has great potential to keep heavy metals
and tolerate their higher concentration (Zhang et al. 2007; Huang and Wang 2010;
Huang et al. 2010). During heavy metal stress, antioxidant activities of mangroves
have been evaluated by many researchers (Macfarlane and Burchett 2001; Zhang
et al. 2007; Caregnato et al. 2008). For instance, significantly higher activity of
peroxidase was recorded in A. marina when exposed to Zn, Pb and Cu metal stress
(concentration less than toxic effect) (Macfarlane and Burchett 2001). Enhanced
activity of catalase enzyme was scored in the leaves and roots of K. candel and
B. gymnorrhiza during stress (Hg, Cd and Pb). Moreover, the leaves of
B. gymnorrhiza showed higher lipid peroxidation than K. candel under stress.
However, higher catalase activity and lower lipid peroxidantion in K. candel
validated its better tolerance of heavy metal stress than B. gymnorrhiza (Zhang
et al. 2007). A significant increment in the quantity of glutathione, proline and
phytochelatins has been recorded in the leaves of B. gymnorrhiza and K. candel
under various heavy metal stresses such as Hg, Pb and Cd (Huang and Wang 2010).
Leaf of A. marina exhibited a noteworthy constructive relationship between the
activities of guaiacol peroxidase and zinc concentration during the studies of antiox-
idant (glutathione) and lipid peroxidation. Likewise, the lipid hydroperoxides con-
centration was also increased in proportion to increasing zinc concentration when
subjected to 2 and 8 weeks of metal stress (Caregnato et al. 2008). Higher mortality
(around 47% in the course of the first 4 years) was observed in the seedlings of
Rhizophora apiculata when planted in the area earlier used for tin mining. However,
mortality was attributed to the varied distribution of soil particle and
microtopography instead of metal contamination.
Due to three characteristics of mangrove habitats have made it preferred sites
from long ago for the dumping of sewage: (1) flow by waste disperses of habitat
from an area to other parts, (2) filtration of nutrients itself by vegetation from the
water and (3) physical processes as well as the soil of mangroves, microbes and algae
associated with mangroves absorb a huge quantity of the pollutants (Wong et al.
1997). Nitrogen and phosphorous (nutrients) are found majorly in the pollutions.
The soil of mangroves retained both phosphorus and nitrogen when treated with
synthetic wastewater (Tam and Wong 1996). This observation suggests that
mangroves tolerate organic pollutants, results must be seen very carefully as they
cannot hold in other habitation. The sewage dumping impacts will be determined by
sewage quantity, dumping duration and the distinct properties of each mangrove
ecosystem. The patterns of water flow through habitat are especially important as it
will determine the rate of flushing and pollutant residence times. Elevated levels of
10 Roles of Mangroves in Combating the Climate Change 243

organic pollution may cause infection, demise and variation in the composition of
species within mangrove forest. Sewage discharge into the Red sea killed
pneumatophores of A. mariana (Mandura 1997). The pneumatophores deficiency
reduced uptake of nutrient, the surface area for respiration and retarded the mangrove
growth. Away from nutrients, organic pollution of the mangrove ecosystem can
contain other debris and anthropogenic chemicals. The sediments of mangroves in
Chengue Bay and Cienaga Grand de Santa Marta have noteworthy residues of
organochlorine pesticide. The levels of some of these pollutants change according
to season. The mangrove habitat of Jamaica contains a large quantity of
non-mangrove wood and plastic. The amount of these solid wastes shows a strong
relationship with rainfall in an adjacent metropolitan region.

10.4.5 Carbon Dioxide

The concentrations of atmospheric CO2 have been rising continuously and their
levels are likely to be more elevated in the upcoming century. It is assumed that
increasing concentration of CO2 can regulate and elevate the impacts of local
environmental stress factors like changing patterns of salinity and tidal inundation
upon the mangrove encroachment into nearby ecosystems (Saintilan and Rogers
2015). Enhanced concentrations of CO2 are likely to increase water use efficiency
and productivity of mangroves under certain environmental stress conditions. This
could also enhance biomass, change biotic interaction and increase extent and
coverage of mangroves (McKee and Rooth 2008; Cherry et al. 2009; Langley
et al. 2009; McKee et al. 2012; Saintilan and Rogers 2015; Lovelock et al. 2016;
Reef et al. 2016). These influences are also depending upon additional factors
like salinity regimes, availability of nutrients and interactions with biotic
factors (Lovelock et al. 2016; Reef et al. 2016). Elevated CO2 may increase the
mangrove growth; however, responses vary according to species. In many species,
these responses are confounded by changes in water use efficiency, availability of
nutrients and soil salinity. The Rhizophora mangle, Rhizophora apiculata and
Rhizophora stylosa showed enhanced growth at elevated CO2 level and less salt
concentration, but not in the case of higher salt concentration. Moreover, all three
species maturated earlier as compared to control plants (Ball et al. 1997). Net
primary productivity of A. germinans, R. mangle and Conocarpus erectus was not
influenced by elevated CO2 level; however, productivity of L. racemosa was
declined. Furthermore, all four species showed enhanced transpiration efficiency;
but with increasing CO2 concentration decline in transpiration and stomatal conduc-
tance was recorded (Snedaker and Araújo 1998). These outcomes suggest that
responses of mangroves to increasing CO2 will be complex, as some species
blooming whereas others show declines or little or even no change. The effects of
the interaction of elevated CO2 with temperature, humidity, nutrient availability and
salinity imply that location of the coastal area might be a key determinant in the
mangrove response. For example, patterns of species in the estuaries can vary
depending upon the capability of species to respond to temporal as well as spatial
244 A. Kumari and M. S. Rathore

variations in the availability of nutrients, salinity and other factors in association to


elevated CO2 levels. The effects of double CO2 concentration were analysed on the
seedlings of Rhizophora mangle by Farnsworth et al. (1996). Seedling grown in
double CO2 concentration exhibited significant increment in the leaf area, branching
activity, total stem length and biomass in comparison to seedling grown in normal
CO2 levels. In this experiment, under a high concentration of CO2, the reproduction
of R. mangle was attained only after 1 year of growth, whilst it naturally takes
complete 2 years to become able for reproduction during field conditions. Hence,
enhanced CO2 also emerged to hasten the maturation of mangroves along with
growth. It is predicted that the increased atmospheric CO2 concentration elevates
mangrove growth but it will not be enough to compensate the harmful effects of sea
level rise. Other effects of elevated CO2 and temperature on mangrove are the coral
reefs degradation triggered by impaired growth and mass bleaching. Impairment of
coral reefs can negatively affect the mangrove ecosystems that depend upon the reefs
for providing shelter from wave action (where mangrove—reef couple system
occurs).

10.4.6 Biotic Factors like Pests

Some of the animals and plants residing in the mangrove forest/mangal are severe
pests. These pests cause impairment of the mangroves, reduction in the growth and
productivity and killing the trees in extreme cases (Kathiresan and Bingham 2001).
Certain harmful species may not harm directly to the mangroves rather than cause
injury by competing for limited resources. Interspecific competition is the most
common process in the forest of mangrove which is confirmed by allelopathic
interaction among the species of mangrove. For instance, leaf litter of certain
mangroves (R. apiculata, Ceriops decandra and Lumnitzera racemosa) secrete
toxic leachates that impede the seedling growth of R. mucronata and R. apiculata
(Kathiresan 1993). Generally, the situation of osmotic stress causes suberization and
lignification in mangroves that prohibit the development of undergrowth of dense
herbaceous in the forests of mangroves. Thus, they reduce intense competition
between non-mangrove plants and mangroves. The Acrostichum (mangrove fern)
is a weedy pest that causing substantial damages to mangrove forests (Kathiresan
and Bingham 2001). Some pest injures the mangroves just by residing on their
surface such as Chiracanthium and Tetragnatha nitens living on Rhizophora. On the
leaves, they lay the eggs, which cause chlorosis, rolling of leaf, and wilting. Severe
invasions may kill the mangroves. The Phthirusa maritime (semi-parasitic mistletoe)
possess more intense and direct impacts on the trees. Higher rate of transpiration,
lower water use efficiency, and rate of CO2 assimilation were detected in Coccoloba
uvifera and Conocarpus erectus when infected with this parasite (Orozco et al.
1990). Due to the feeding action of herbivores, most severe and extensive damage
has been reported in the mangroves. The tidal regime and current are the main factors
that controlled the distribution of pests. Indeed, insects are the most destructive
among the animals feed on the canopy of mangroves. Some insect herbivores merely
10 Roles of Mangroves in Combating the Climate Change 245

utilize mangroves as a substitute host since they are serious crop pests. Some
examples of pests have the apparent choice for mangroves, viz. Antestiopsis on
Avicennia, Calliphara on Excoecaria, Glaucias on Lumnitzera and Mictis on
Sonneratia. Mangroves stands can be fully defoliated by insect herbivores
(Kathiresan and Bingham 2001). When Aspidiotus destructor (scale insects)
attacked on the leaves of Rhizophora, leaves turn yellow at the site of infection,
subsequently brown and finally necrosis happens. In case of severe infection, leaves
dry up, fall down and even the death of complete seedlings occurs. Certain
herbivores especially infect to the seeds and reproductive tissues of mangroves.
The Afrocypholaelaps africana (mite) feeds on the pollen of mangroves. Likewise,
crabs are the main predators of seed that damage the seeds severely. However,
mangrove forests cover a smaller portion of tropical forest; latest research
demonstrated that herbivory levels of these forests were similar to other forests of
the tropics and temperate. Although the survivals of mangroves are very well in
response to an invasion because they have distinct phytochemicals such as
derivatives of polyphenol, phenols, terpenoids, alkaloids and occurrence of numer-
ous antioxidant defence mechanism (Bandaranayake 2002). Mutual association
between mangroves and endophytes Aspergillus flavus (e.g. E. agallocha L.,
R. mucronata Lam., A. officinalis L., K. candel Druce) improves the defence
responses of mangroves against different abiotic and biotic stresses (Ravindran
et al. 2012; Thatoi et al. 2014). The defence system of plants in response to pathogen
attack includes oxidative burst process resulting in ROS production. Plants activated
multiple defence process against biotic stress that prompts ROS scavenging system
of plants. Insect herbivores can exhibit priority amongst mangrove hosts. For
instance, the Oiketicus kirbyi (bagworm) eliminated around 5% foliage of the
Laguncularia racemosa, 10% of the Conocarpus erectus and 80% of the Avicennia
germinans in an Ecuadorian mangal (Gara et al. 1990). The physico-chemical
properties may correlate with the susceptibility of individual mangroves plants or
species of mangroves. More toughness of leaf decreases digestibility and palatabil-
ity. Tannins also protect from infection with herbivores. Species of Rhizophora have
higher levels of tannin, thus suffer less damage from herbivore attacks, than that of
Avicennia species (having less content of tannins). The health of mangroves may
also influence the feeding choice of insects. Trees enriched with nutrients tend to
more suffering with herbivore attacks. Rhizophora mangle when treated with P and
NPK exhibits significantly higher infection with herbivores Marmara (mines stem)
and Ecdytolopha (feed on apical buds). However, treatment with N only did not
enhance the attack of herbivory, but their seeds can also be infected by insects. The
propagules of Avicennia marina are likely to injure by insect borers, but this does not
kill them (Robertson and Duke 1990).
246 A. Kumari and M. S. Rathore

10.5 Ecological Importance of Mangroves

Mangroves are one of the most productive and valuable ecosystems in the world.
They provide a lot of benefits to the local communities in addition to the ecology and
environment surrounding them. In response to environmental inconstancy,
mangroves show a higher range of community persistence and ecological stability.
They provide an extensive array of economic and ecosystem services like resources
of fuel and food; nursery ground for aquatic and terrestrial fauna; cycling of
nutrients; formation of soil and wood; carbon sequestration and ecotourism (Alongi
2012). Mangroves stabilize and protect the coastline, enrich the waters of coastal
area, supported fisheries of coastal area and yield commercial forest products. They
are biodiversity-rich and protect from coastal erosion that happens because of
intense tropical storm and tsunamis. The rates of carbon production in the mangroves
ecosystem are equal to the tropical humid forest. Mangroves exhibit higher ratios of
carbon mass in below- to above-ground than terrestrial trees and allocate corre-
spondingly higher carbon in the below-ground part (Alongi 2012). Most carbon-rich
biomes are mangroves, having an average of 937 tC ha 1. They accelerate the rate of
sediment accumulation (~5 mm year 1), the burial of carbon (174 gC m 2 year 1)
and facilitating more accretion of fine particles (Alongi 2012). Mangroves contribute
just around 1% (13.5 Gt year 1) carbon sequestration of the world’s forests.
However, being habitats of the coastal area they contribute 14% carbon sequestra-
tion through the ocean. If the stocks of carbon are disturbed, it may result a huge
emissions of gas (Alongi 2012). Mangroves modify the environment of soil which in
turn influences the growth and survival of each and every functional variety of
anaerobic as well as aerobic bacteria. Energetically efficient and highly evolved
interactions of plant–soil–microbes are key factor account for high productivity of
mangroves during the very harsh situation. The mangroves that are distinctly
adjusted themselves to face the continuous tidal movement are also able to endure
stronger energy of wind and force of waves happen during extreme events of
weather. Based on the ecological situation, mangroves absorb minimum 70–90%
of the waves energy and act like physical buffers between the shore and elements
(UNEP-WCMC 2006). Organic matter and roots of mangroves absorb floodwater by
working as a sponge and also play a role in to trap sediment. Roots of mangroves
also act as filters to sort out pollutants that come to sea from inland waters. So, they
help in improving the water quality reaching the ecosystem of the sea. The plant
removes CO2 from the atmosphere and accumulates it as biomass through a process
that is recognized as carbon sequestration. That is why ocean and plants are known
as carbon sinks. It is projected that a large number of carbons are sequestered by
mangroves, around 25.5 million tonnes of carbon in a year (Eong 1993). There is
also the anticipation that mangroves deliver greater than 10% of vital dissolved
organic carbon that is provided to the world ocean from land (Dittmar et al. 2006).
Due to the tangled and extensive supportive root systems of mangroves, they trap
sediments more efficiently and prevent them from wash into the sea. The timber of
mangrove is utilized to make and build furniture, houses, boats, fences and rafters all
over the world. Around 300,000m3 wood of mangrove is extracted every year from
10 Roles of Mangroves in Combating the Climate Change 247

Sundarbans (Miththapala 2008). The wood of mangroves is also employed as


fuelwood and yet supply 90% of fuel utilized in Vietnam. The screw pine (panda-
nus) and mangrove palm (Nypa) leaves are used in weaving and thatching. The
Cerbera manghas wood is used to crave puppets and masks in Sri Lanka. Due to the
presence of salt glands, mangroves serve as a source of sodium. The ash of certain
species like Avicennia is utilized as soap. The aerenchyma tissues of the several
Sonneratia spp with their breathing roots are applied in the production of fish floats
and corks. Tannins and gums are produced from the bark of various species that are
yet utilized in the Indian subcontinents to cure fish nets and leather (Miththapala
2008). In India and Bangladesh, an important local industry is the production of
honey from mangroves. They produce almost 20 tonnes of honey yearly from
200,000 ha of mangroves. Shoots, roots, leaves and fruits of mangroves serve as
edible fruits and vegetables in several parts of the world. Moreover, drinks and
sugars are extracted from the Sonneratia species (Miththapala 2008). Approximately
70 distinct mangrove plants are recorded to possess traditional medicinal utilization
in the treatment of several diseases and ailments (Bandaranayake 1998). For
instance, Lumnitzera, Rhizophora and Bruguiera are applied in many ailments like
angina, diarrhoea and blood pressure (Upadhyay et al. 2002).

10.6 Future Prospective and Conclusion

Mangroves are a unique group of plants, located at the interface between land and
sea. They grow luxuriantly in extreme environmental conditions like high salinity,
extreme temperature and tides, flood, strong winds and anaerobic soil. For the
growth and survival during harsh conditions, they developed several adaptations
such as morphological, physiological and ecological. These adaptations make them
more resistant and resilient to the adverse impacts of the harsh conditions.
Mangroves provide a lot of economic and environmental services like nutrient
cycling, carbon sequestration, soil formation, enrichment of biodiversity of coastal
areas, supporting fisheries and ecotourism. They also protect coastal ecosystems
from extreme environmental conditions. To overcome the deleterious impacts of
salinity, mangroves possess well-developed ion sequestration, salt accumulation and
salt excreting properties. To eliminate the excess ROS produced during stress
conditions, they also have a variable extent of antioxidant enzyme and
non-enzyme activity. These antioxidants help in ROS detoxification, protect the
cellular and photosynthetic functions from the ROS induced oxidative stress and
maintain healthy growth and productivity in the harsh conditions. Since very little
information about the molecular mechanism of mangroves against climate change
are available. A deep understanding and knowledge of the key stress-related gene
and cross talk among the various signalling components must remain a strong
research area in the future. By getting the complete picture of the fundamental
mechanism of mangroves to cope with several biotic and abiotic stresses may be
utilized in a logical and systemic way that will provide great vision in regard to their
potential application. Moreover, the development of climate-resilient crops using the
248 A. Kumari and M. S. Rathore

major stress-responsive genes from mangroves will be a better choice to overcome


the problem of food security for ever-growing populations.

Acknowledgements CSIR-CSMCRI PRIS 172/2020.


The authors are heartily thankful to Council of Scientific and Industrial Research (CSIR), New
Delhi for financial support in the form of infrastructure and different plan projects. AK is thankful to
the Department of Biotechnology (DBT), Government of India and Academy of Scientific and
Innovation Research (AcSIR), Ghaziabad for fellowship in the form of JRF/SRF and registration in
the Ph.D program.

References
Abood JK, Lösel DM (2003) Changes in carbohydrate composition of cucumber leaves during the
development of powdery mildew infection. Plant Pathology, 52(2):256–265.
Alongi D (2009) The energetics of mangrove forests. Springer Science & Business Media. Springer
Netherlands.
Alongi DM (2002) Present state and future of the world’s mangrove forests. Environ Conserv 29
(3):331–349.
Alongi DM (2008) Mangrove forests: resilience, protection from tsunamis, and responses to global
climate change. Estuar Coast Shelf Sci 76(1):1–13.
Alongi DM (2012) Carbon sequestration in mangrove forests. Carbon Manage 3(3):313–322.
Alongi DM (2015) The impact of climate change on mangrove forests. Curr Clim Change Rep 1
(1):30–39.
Alongi DM, Sasekumar A, Tirendi F, Dixon P (1998) The influence of stand age on benthic
decomposition and recycling of organic matter in managed mangrove forests of Malaysia. J
Exp Mar Biol Ecol 225(2):197–218.
Andrews TJ, Clough BF, Muller GJ (1984) Photosynthetic gas exchange properties and carbon
isotope ratios of some mangroves in North Queensland. In Physiology and Management of
Mangroves (pp. 15–23). Springer, Dordrecht.
Atreya A, Vartak V, Bhargava S (2009) Salt priming improves tolerance to desiccation stress and to
extreme salt stress in Bruguiera cylindrica. Inter J Integr Biol 6(2):68–73.
Ball MC, Sobrado MA (2002) Ecophysiology of mangroves: challenges in linking physiological
processes with patterns in forest structure. In Physiological Plant Ecology: 39th Symposium of
the British Ecological Society (Vol. 39, pp. 331–346), Cambridge University Press.
Ball MC, Pidsley SM (1995) Growth responses to salinity in relation to distribution of two
mangrove species, Sonneratia alba and S. lanceolata, in northern Australia. Func Ecol 9:77–85.
Ball MC, Cochrane MJ, Rawson HM (1997) Growth and water use of the mangroves Rhizophora
apiculata and R. stylosa in response to salinity and humidity under ambient and elevated
concentrations of atmospheric CO2. Plant Cell Environ 20(9):1158–1166.
Bandaranayake W (1998) Traditional and medicinal uses of mangroves. Mangroves and Salt
Marshes 2(3):133–148.
Bandaranayake WM (2002) Bioactivities, bioactive compounds and chemical constituents of
mangrove plants. Wet Ecol Manage 10(6):421–452.
Barnabás B, Jäger K, Fehér A (2008) The effect of drought and heat stress on reproductive
processes in cereals. Plant cell Environ 31(1):11–38.
Bengtsson L, Hodges KI, Roeckner E (2006) Storm tracks and climate change. J Clim 19
(15):3518–3543.
Bhargava S, Sawant K (2013) Drought stress adaptation: metabolic adjustment and regulation of
gene expression. Plant Breeding 132(1):21–32.
Bruun P (1988) The Bruun rule of erosion by sea-level rise: a discussion on large-scale two-and
three-dimensional usages. J Coast Res 4:627–648.
10 Roles of Mangroves in Combating the Climate Change 249

Burns WC (2001) The possible impacts of climate change on Pacific island state ecosystems. Inter J
Global Environ Issues 1(1):56–72.
Cahoon DR, Hensel P (2006) High-resolution global assessment of mangrove responses to sea-level
rise: a review. In Proceedings of the Symposium on Mangrove Responses to Relative Sea Level
Rise and Other Climate Change Effects (Vol. 13, pp. 9–17).
Cahoon DR, Hensel P, Rybczyk J, McKee KL, Proffitt CE, Perez BC (2003) Mass tree mortality
leads to mangrove peat collapse at Bay Islands, Honduras after Hurricane Mitch. J Ecol 91
(6):1093–1105.
Cahoon DR, Hensel PF, Spencer T, Reed DJ, McKee KL, Saintilan N (2006) Coastal wetland
vulnerability to relative sea-level rise: wetland elevation trends and process controls. In
Wetlands and Natural Resource Management (pp. 271–292). Springer, Berlin, Heidelberg.
Caregnato FF, Koller CE, MacFarlane GR, Moreira JC (2008) The glutathione antioxidant system
as a biomarker suite for the assessment of heavy metal exposure and effect in the grey mangrove,
Avicennia marina (Forsk.) Vierh. Mar Poll Bull 56(6):1119–1127.
Cavanaugh KC, Kellner JR, Forde AJ, Gruner DS, Parker JD, Rodriguez W, Feller IC (2014)
Poleward expansion of mangroves is a threshold response to decreased frequency of extreme
cold events. Proc Nat Acad Sci 111(2):723–727.
Chakraborty K, Bhaduri D, Uprety DC, Patra AK (2014) Differential response of plant and soil
processes under climate change: a mini-review on recent understandings. Proc Nat Acad Sci,
India Section B: Biol Sci 84(2):201–214.
Chapman VJ (1976) Mangrove vegetation. Vaduz, Liechtenstein: Cramer, 447.
Chen R, Twilley RR (1998) A gap dynamic model of mangrove forest development along gradients
of soil salinity and nutrient resources. J Ecol 86(1):37–51.
Cherian S, Reddy MP, Pandya JB (1999) Studies on salt tolerance in Avicennia marina (Forstk.)
Vierh.: effect of NaCl salinity on growth, ion accumulation and enzyme activity. Indian J Plant
Physiol 4(4):266–270.
Cherry JA, McKee KL, Grace JB (2009) Elevated CO2 enhances biological contributions to
elevation change in coastal wetlands by offsetting stressors associated with sea-level rise. J
Ecol 97(1):67–77.
Christensen JH, Hewitson B, Busuioc A, Chen A, Gao X, Held R, Jones R, Kolli RK, Kwon WK,
Laprise R, Magaña Rueda V, Mearns L, Menéndez CG, Räisänen J, Rinke A, Sarr A, Whetton P
(2007) Regional climate projections. In Climate Change, 2007: The Physical Science Basis.
Contribution of Working group I to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change, University Press, Cambridge, Chapter 11 (pp. 847–940).
Cambridge/New York.
Church JA, White NJ (2006) A 20th century acceleration in global sea-level rise. Geophys Res Lett
33(1): https://doi.org/10.1029/2005GL024826.
Church JA, Gregory JM, Huybrechts P, Kuhn M, Lambeck K, Nhuan MT, Qin D, Woodworth PL
(2001) Changes in sea level. In: JT Houghton, Y Ding, DJ Griggs, M Noguer, PJ Van der
Linden, X. Dai, K. Maskell, and CA Johnson (eds.): Climate Change 2001: The Scientific Basis:
Contribution of Working Group I to the Third Assessment Report of the Intergovernmental
Panel (pp. 639–694). Cambridge University Press, Cambridge, United Kingdom, and
New York, NY, USA.
Dahdouh-Guebas F, Koedam N, Danielsen F, Sørensen MK, Olwig MF, Selvam V, Parish F,
Burgess ND, Topp-Jørgensen E, Hiraishi T, Karunagaran VM, Rasmussen MS, Hansen LB,
Quarto A, Suryadiputra N (2006) Coastal vegetation and the Asian tsunami. Science 311:37–38.
Dai A (2011) Drought under global warming: a review. Wiley Interdisciplinary Reviews: Climate
Change 2(1):45–65.
Danielsen F, Sørensen MK, Olwig MF, Selvam V, Parish F, Burgess ND, Hiraishi T, Karunagaran
VM, Rasmussen MS, Hansen LB, Quarto A (2005) The Asian tsunami: a protective role for
coastal vegetation. Science 310(5748):643–643.
Das PK, Chakravarti V, Dutta A, Maity S (1995) Leaf anatomy and chlorophyll estimates in some
mangroves. Indian Forester 121(4):289–294.
250 A. Kumari and M. S. Rathore

Das S, Ghose M (1993) Morphology of stomata and leaf hairs of some halophytes from Sundarbans,
West Bengal. Phytomorphol 43(1–2):59–70.
Das SK, Patra JK, Thatoi H (2016) Antioxidative response to abiotic and biotic stresses in
mangrove plants: A review. Internat Rev Hydrobiol 101(1–2):3–19.
Dasgupta N, Nandy P, Sengupta C, Das S (2012) Protein and enzymes regulations towards salt
tolerance of some Indian mangroves in relation to adaptation. Trees 26(2):377–391.
De Ollas C, Morillón R, Fotopoulos V, Puértolas J, Ollitrault P, Gómez-Cadenas A, Arbona V
(2019) Facing climate change: biotechnology of iconic Mediterranean woody crops. Front Plant
Sci 10:427.
Dittmar T, Hertkorn N, Kattner G, Lara RJ (2006) Mangroves, a major source of dissolved organic
carbon to the oceans. Global Biogeochem Cyc 20(1):GB1012, doi: https://doi.org/10.1029/
2005GB002570.
Doganlar ZB, Demir K, Basak H, Gul I (2010) Effects of salt stress on pigment and total soluble
protein contents of three different tomato cultivars. Afr J Agri Res 5(15):2056–2065.
Duke N, Ball M, Ellison J (1998) Factors influencing biodiversity and distributional gradients in
mangroves. Glob Ecol Biogeogr Lett 7(1):27–47.
Duke NC (1992) Mangrove floristics and biogeography. Tropical mangrove ecosystems,
(pp. 63–100). American Geophysical Union, Washington, D.C., USA.
Duke NC, Kovacs JM, Griffiths AD, Preece L, Hill DJ, Van Oosterzee P, Mackenzie J, Morning
HS, Burrows D (2017) Large-scale dieback of mangroves in Australia’s Gulf of Carpentaria: a
severe ecosystem response, coincidental with an unusually extreme weather event. Mar Fresh-
water Res 68(10):1816–1829.
Duke NC, Meynecke JO, Dittmann S, Ellison AM, Anger K, Berger U, Cannicci S, Diele K, Ewel
KC, Field CD, Koedam N (2007) A world without mangroves?. Science 317(5834):41–42.
Ellison JC, Stoddart DR (1991) Mangrove ecosystem collapse during predicted sea-level rise:
Holocene analogues and implications. J Coast Res 7:151–165.
Ellison JC (2000) How South Pacific mangroves may respond to predicted climate change and
sea-level rise. In Climate Change in the South Pacific: Impacts and Responses in Australia,
New Zealand, and Small Island States (pp. 289–300). Springer, Dordrecht.
Eong OJ (1993) Mangroves-a carbon source and sink. Chemosphere 27(6):1097–1107.
Eslami-Andargoli L, Dale PER, Sipe N, Chaseling J (2009) Mangrove expansion and rainfall
patterns in Moreton Bay, southeast Queensland, Australia. Estuar Coast Shelf Sci 85
(2):292–298.
Fancy NN, Bahlmann AK, Loake GJ (2017) Nitric oxide function in plant abiotic stress. Plant Cell
Environ 40(4):462–472.
Farnsworth EJ, Ellison AM, Gong WK (1996) Elevated CO2 alters anatomy, physiology, growth,
and reproduction of red mangrove (Rhizophora mangle L.). Oecologia 108(4):599–609.
Feller IC (1995) Effects of nutrient enrichment on growth and herbivory of dwarf red mangrove
(Rhizophora mangle). Ecol Monogr 65(4):477–505.
Field CD (1995) Impact of expected climate change on mangroves. In Asia-Pacific Symposium on
Mangrove Ecosystems (pp. 75–81). Springer, Dordrecht.
Fleta-Soriano E, Munné-Bosch S (2016) Stress memory and the inevitable effects of drought: a
physiological perspective. Front Plant Sci 7:143.
Food and Agriculture Organization of the United Nations, (2007) The world’s mangroves
1980–2005. FAO Forestry Paper 153.
Forbes MG, Dunton KH (2006) Response of a subtropical estuarine marsh to local climatic change
in the southwestern Gulf of Mexico. Estuar Coast 29(6):1242–1254.
Foyer CH, Shigeoka S (2011) Understanding oxidative stress and antioxidant functions to enhance
photosynthesis. Plant physiol 155(1):93–100.
Gara RI, Sarango A, Cannon PG (1990) Defoliation of an Ecuadorian mangrove forest by the
bagworm, Oiketicus kirbyi Guilding (Lepidoptera: Psychidae). J Tropic For Sci 3(2):181–186.
10 Roles of Mangroves in Combating the Climate Change 251

García-Sánchez F, Syvertsen JP, Gimeno V, Botía P, Perez-Perez JG (2007) Responses to flooding


and drought stress by two citrus rootstock seedlings with different water-use efficiency. Physiol
Plant 130(4):532–542.
Geng Q, Sugi M (2003) Possible change of extratropical cyclone activity due to enhanced
greenhouse gases and sulfate aerosols-Study with a high-resolution AGCM. J Clim 16
(13):2262–2274.
Gilman EL, Ellison J, Duke NC, Field C (2008) Threats to mangroves from climate change and
adaptation options: a review. Aquat Bot 89(2):237–250.
Gopal B, Chauhan M (2006) Biodiversity and its conservation in the Sundarban mangrove
ecosystem. Aquat Sci 68(3):338–354.
Hamilton SE, Casey D (2016) Creation of a high spatio-temporal resolution global database of
continuous mangrove forest cover for the 21st century (CGMFC-21). Glob Ecol Biogeogr 25
(6):729–738.
Hirayama T, Shinozaki K (2010) Research on plant abiotic stress responses in the post-genome era:
Past, present and future. The Plant J 61(6):1041–1052.
Holgate SJ, Woodworth PL (2004) Evidence for enhanced coastal sea level rise during the 1990s.
Geophys Res Lett 31(7).
Houghton J, Ding Y, Griggs D, Noguer M, van der Linden P, Dai X, Maskell K, Johnson C (Eds.),
2001. Climate Change 2001: The Scientific Basis (Published for the Intergovernmental Panel on
Climate Change). Cambridge University Press, Cambridge, United Kingdom, and New York,
NY, USA.
Huang GY, Wang YS (2010) Expression and characterization analysis of type 2 metallothionein
from grey mangrove species (Avicennia marina) in response to metal stress. Aquat Toxicol 99
(1):86–92.
Huang GY, Wang YS, Sun CC, Dong JD, Sun ZX (2010) The effect of multiple heavy metals on
ascorbate, glutathione and related enzymes in two mangrove plant seedlings (Kandelia candel
and Bruguiera gymnorrhiza). Oceanol Hydrobiol Stud 39(1):11–25.
Hutchings P, Saenger P (1987) Ecology of mangroves. Ecology of mangroves. University of
Queensland Press, St. Lucia.
IPCC (2007) Climate change 2007: synthesis report. Contribution of Working Groups I, II and III to
the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. IPCC,
Geneva, Switzerland, p. 104.
Jaleel CA, Manivannan P, Wahid A, Farooq M, Al-Juburi HJ, Somasundaram R, Panneerselvam R
(2009) Drought stress in plants: a review on morphological characteristics and pigments
composition. Inter J Agri Biol 11(1):100–105.
Jithesh MN, Prashanth SR, Sivaprakash KR, Parida AK (2006) Antioxidative response mechanisms
in halophytes: their role in stress defence. J Genet 85(3):237.
Joshi R, Wani SH, Singh B, Bohra A, Dar ZA, Lone AA, Pareek A, Singla-Pareek SL (2016)
Transcription factors and plants response to drought stress: current understanding and future
directions. Front Plant Sci 7:1029.
Kaly UL, Eugelink G, Robertson AI (1997) Soil conditions in damaged North Queensland
mangroves. Estuaries 20(2):291–300.
Kandasamy K (2017) Mangroves in India and climate change: an overview. In: Participatory
Mangrove Management in a Changing Climate (pp. 31–57). Springer, Tokyo.
Kathiresan K, Bingham BL (2001) Biology of mangroves and mangrove ecosystems. Adv Mar Biol
40:84–254.
Kathiresan K, Rajendran N (2005) Coastal mangrove forests mitigated tsunami. Estuar Coast Shelf
Sci 65(3):601–606.
Kathiresan K (1993) Dangerous pest on nursery seedlings of Rhizophora. Indian Forester 119(12).
Keyvan S (2010) The effects of drought stress on yield, relative water content, proline, soluble
carbohydrates and chlorophyll of bread wheat cultivars. J Ani Plant Sci 8(3):1051–1060.
Koch MS, Snedaker SC (1997) Factors influencing Rhizophora mangle L. seedling development in
Everglades carbonate soils. Aquat Bot 59(1–2):87–98.
252 A. Kumari and M. S. Rathore

Korres NE, Norsworthy JK, Tehranchian P, Gitsopoulos TK, Loka DA, Oosterhuis DM, Gealy DR,
Moss SR, Burgos NR, Miller MR, Palhano M (2016) Cultivars to face climate change effects on
crops and weeds: a review. Agro Sustain Develop 36(1):12.
Kosová K, Prášil IT, Vítámvás P (2013) Protein contribution to plant salinity response and tolerance
acquisition. Internat J Mole Sci 14(4):6757–6789.
Krauss KW, Allen JA, Cahoon DR (2003) Differential rates of vertical accretion and elevation
change among aerial root types in Micronesian mangrove forests. Estuar Coast Shelf Sci 56
(2):251–259.
Kundzewicz ZW, Mata LJ, Arnell NW, Doll P, Kabat P, Jimenez B, Miller K, Oki T, Zekai S,
Shiklomanov I (2007) Freshwater resources and their management. In Climate change 2007:
impacts, adaptation and vulnerability contribution of Working Group II to the Fourth Assess-
ment Report of the Intergovernmental Panel on Climate Change. (pp. 173–210) Cambridge
University Press, Cambridge.
Lacerda LD, Ittekkot V, Patchineelam SR (1995) Biogeochemistry of Mangrove Soil Organic
Matter: a Comparison Between Rhizophora and Avicennia Soils in South-eastern Brazil. Estuar
Coast Shelf Sci 40(6):713–720.
Langley JA, McKee KL, Cahoon DR, Cherry JA, Megonigal JP (2009) Elevated CO2 stimulates
marsh elevation gain, counterbalancing sea-level rise. Proce Nat Acad Sci 106(15):6182–6186.
Li MS, Lee SY (1997) Mangroves of China: a brief review. Forest Ecol Manage 96(3):241–259.
List JH, Sallenger Jr AH, Hansen ME, Jaffe BE (1997) Accelerated relative sea-level rise and rapid
coastal erosion: testing a causal relationship for the Louisiana barrier islands. Mar Geol 140
(3–4):347–365.
Lovelock CE, Ellison JC, (2007) Vulnerability of mangroves and tidal wetlands of the Great Barrier
Reef to climate change. In: Climate Change and the Great Barrier Reef: A Vulnerability
Assessment. (pp. 237–269) Great Barrier Reef Marine Park Authority and Australian Green-
house Office, Australia.
Lovelock CE, Krauss KW, Osland MJ, Reef R, Ball MC (2016) The physiology of mangrove trees
with changing climate. In Tropical Tree Physiology: Adaptations and Responses in a Changing
Environment, (pp. 149–179). Springer, New York, New York, USA.
MacFarlane GR, Burchett MD (2001) Photosynthetic pigments and peroxidase activity as indicators
of heavy metal stress in the grey mangrove, Avicennia marina (Forsk.) Vierh. Mar Pollut Bull 42
(3):233–240.
MacNae W (1969) A general account of the fauna and flora of mangrove swamps and forests in the
Indo-West-Pacific region. In: Advances in Marine Biology (Vol. 6, pp. 73–270). Academic
Press.
Mandura AS (1997) A mangrove stand under sewage pollution stress: Red Sea. Mangroves and Salt
Marshes, 1(4):255–262.
McKee K, Rogers K, Saintilan N (2012) Response of salt marsh and mangrove wetlands to changes
in atmospheric CO2, climate, and sea level. In Global Change and the Function and Distribution
of Wetlands (pp. 63–96). Springer, Dordrecht.
McKee KL, Rooth JE (2008) Where temperate meets tropical: multi-factorial effects of elevated
CO2, nitrogen enrichment, and competition on a mangrove-salt marsh community. Glob Change
Biol 14:971–984.
McKee KL (1993) Soil physicochemical patterns and mangrove species distribution—reciprocal
effects?. J Ecol 81:477–487.
McKee KL, Cahoon DR, Feller IC (2007) Caribbean mangroves adjust to rising sea level through
biotic controls on change in soil elevation. Glob Ecol Biogeogr 16(5):545–556.
Miller PC, Hom J, Poole DK (1975) Water relations of three mangrove species in south Florida.
Oecol Plant Fr Da 10:355–367.
Miththapala S (2008) Mangroves. Coastal ecosystems series, Volume 2. Mangroves. Coastal
ecosystems series, Volume 2. IUCN, Sri Lanka.
10 Roles of Mangroves in Combating the Climate Change 253

Mitra A, Gangopadhyay A, Dube A, Schmidt AC, Banerjee K (2009) Observed changes in water
mass properties in the Indian Sundarbans (northwestern Bay of Bengal) during 1980–2007. Curr
Sci 1445–1452.
Nabeelah Bibi S, Fawzi MM, Gokhan Z, Rajesh J, Nadeem N, Rengasamy Kannan RR,
Albuquerque RDDG, Pandian SK (2019) Ethnopharmacology, phytochemistry, and global
distribution of mangroves-A comprehensive review. Marine Drugs, 17(4):231.
Nakashima K, Ito Y, Yamaguchi-Shinozaki K (2009) Transcriptional regulatory networks in
response to abiotic stresses in Arabidopsis and grasses. Plant Physiol 149(1):88–95.
Nikalje GC, Srivastava AK, Pandey GK, Suprasanna P (2018) Halophytes in biosaline agriculture:
Mechanism, utilization, and value addition. Land Degr Develop 29(4):1081–1095.
Orozco A, Rada F, Azocar A, Goldstein G (1990) How does a mistletoe affect the water, nitrogen
and carbon balance of two mangrove ecosystem species?. Plant Cell Environ 13(9):941–947.
Osland MJ, Enwright N, Stagg CL (2014) Freshwater availability and coastal wetland foundation
species: ecological transitions along a rainfall gradient. Ecology 95(10):2789–2802.
Osland MJ, Enwright N, Day RH, Doyle TW (2013) Winter climate change and coastal wetland
foundation species: salt marshes vs. mangrove forests in the southeastern United States. Glob
Change Biol 19(5):1482–1494.
Parida AK, Jha B (2010) Salt tolerance mechanisms in mangroves: a review. Trees 24(2):199–217.
Parida AK, Das AB, Mohanty P (2004) Investigations on the antioxidative defence responses to
NaCl stress in a mangrove, Bruguiera parviflora: differential regulations of isoforms of some
antioxidative enzymes. Plant Growth Regulation 42(3):213–226.
Pilkey OH, Cooper JAG (2004) Society and sea level rise. Science 303(5665):1781–1782.
Piou C, Feller IC, Berger U, Chi F (2006) Zonation patterns of Belizean offshore mangrove forests
41 years after a catastrophic hurricane 1. Biotropica: J Biol Conserv 38(3):365–374.
Qadir M, Tubeileh A, Akhtar J, Larbi A, Minhas PS, Khan MA (2008) Productivity enhancement of
salt-affected environments through crop diversification. Land Degr Develop 19(4):429–453.
Rai V, Vajpayee P, Singh SN, Mehrotra S (2004) Effect of chromium accumulation on photosyn-
thetic pigments, oxidative stress defense system, nitrate reduction, proline level and eugenol
content of Ocimum tenuiflorum L. Plant Sci 167(5):1159–1169.
Ravindran C, Naveenan T, Varatharajan GR, Rajasabapathy R, Meena RM (2012) Antioxidants in
mangrove plants and endophytic fungal associations. Bot Mar 55(3):269–279.
Reef R, Slot M, Motro U, Motro M, Motro Y, Adame MF, Garcia M, Aranda J, Lovelock CE,
Winter K (2016) The effects of CO2 and nutrient fertilisation on the growth and temperature
response of the mangrove Avicennia germinans. Photosynth Res 129(2):159–170.
Reid PC, Beaugrand G (2012) Global synchrony of an accelerating rise in sea surface temperature.
Marine Biological Association of the United Kingdom, 92(7):1435.
Robertson AI, Duke NC (1990) Mangrove fish-communities in tropical Queensland, Australia:
spatial and temporal patterns in densities, biomass and community structure. Mar Biol 104
(3):369–379.
Rogers K, Saintilan N, Cahoon D (2005) Surface elevation dynamics in a regenerating mangrove
forest at Homebush Bay, Australia. Wet Ecol Manage 13(5):587–598.
Saenger P (2002) Mangrove ecology, silviculture and conservation. Springer Science & Business
Media Berlin, Germany.
Saintilan N, Rogers K (2015) Woody plant encroachment of grasslands: a comparison of terrestrial
and wetland settings. New Phytol 205(3):1062–1070.
Saintilan N, Williams RJ (1999) Mangrove transgression into saltmarsh environments in south-east
Australia. Glob Ecol Biogeogr 8(2):117–124.
Saintilan N, Wilson NC, Rogers K, Rajkaran A, Krauss KW (2014) Mangrove expansion and salt
marsh decline at mangrove poleward limits. Glob Change Biol 20(1):147–157.
Sapeta H, Costa JM, Lourenco T, Maroco J, Van der Linde P, Oliveira MM (2013) Drought stress
response in Jatropha curcas: growth and physiology. Environ Exp Bot 85:76–84.
Schneider P (2011) The discovery of tropical mangroves in graeco-roman antiquity: Science and
wonder. Hakluyt Society pp. 1–16.
254 A. Kumari and M. S. Rathore

Scholander PF (1968) How mangroves desalinate seawater. Physiol Plant 21(1):251–261.


Scholander PF, Hammel HT, Hemmingsen EA, Bradstreet ED (1964) Hydrostatic pressure and
osmotic potential in leaves of mangroves and some other plants. Proce Nat Acad Sci, USA 52
(1):119.
Sebastiani L, Scebba F, Tognetti R (2004) Heavy metal accumulation and growth responses in
poplar clones Eridano (Populus deltoides x maximowiczii) and I-214 (P. x euramericana)
exposed to industrial waste. Environ Exp Bot 52(1):79–88.
Singh AK, Ansari A, Kumar D, Sarkar UK (2012) Status, biodiversity and distribution of
mangroves in India: an overview. Uttar Pradesh State Biodiversity Board. Marine Biodiversity:
One Ocean, Many Worlds of Life, 59–67.
Snedaker S (1993) Impact on mangroves. In: Climate Change in the Intra-American Seas:
Implications of Future Climate Change on the Ecosystems and Socio-economic Structure of
the Marine and Coastal Regimes of the Caribbean Sea, Gulf of Mexico, Bahamas and N. E.
Coast of South America. (pp. 282–305) Edward Arnold, London.
Snedaker SC, Araújo RJ (1998) Stomatal conductance and gas exchange in four species of
Caribbean mangroves exposed to ambient and increased CO2. Mar Freshwat Res 49
(4):325–327.
Snedaker SC (1995) Mangroves and climate change in the Florida and Caribbean region: scenarios
and hypotheses. Hydrobiologia 295(1–3):43–49.
Sobrado MA (2005) Leaf characteristics and gas exchange of the mangrove Laguncularia racemosa
as affected by salinity. Photosynthetica 43(2):217–221.
Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt KB, Tignor M, Miller HL (2007)
Climate change 2007: The physical science basis. Intergovernmental Panel on Climate Change
(IPCC), Cambridge University Press, Cambridge.
Spaulding M, Kainuma M, Collins L (2010) World Atlas of Mangroves. A collaborative project of
ITTO, ISME, FAO, UNEP-WCMC, Earthscan: London, UK.
Suárez N, Sobrado MA (2000) Adjustments in leaf water relations of mangrove (Avicennia
germinans) seedlings grown in a salinity gradient. Tree Physiol 20(4):277–282.
Takemura T, Hanagata N, Sugihara K, Baba S, Karube I, Dubinsky Z (2000) Physiological and
biochemical responses to salt stress in the mangrove, Bruguiera gymnorrhiza. Aquat Bot 68
(1):15–28.
Tam NFY, Wong YS (1996) Retention of wastewater-borne nitrogen and phosphorus in mangrove
soils. Environ Technol 17(8):851–859.
Thatoi H, Samantaray D, Das SK (2016) The genus Avicennia, a pioneer group of dominant
mangrove plant species with potential medicinal values: a review. Front Life Sci 9(4):267–291.
Thatoi HN, Patra JK, Das SK (2014) Free radical scavenging and antioxidant potential of mangrove
plants: a review. Acta Physiol Plant 36(3):561–579.
Thomas N, Lucas R, Bunting P, Hardy A, Rosenqvist A, Simard M (2017) Distribution and drivers
of global mangrove forest change, 1996–2010. PloS one 12(6):e0179302.
Thomas R, Rignot E, Casassa G, Kanagaratnam P, Acuña C, Akins T, Brecher H, Frederick E,
Gogineni P, Krabill W, Manizade S (2004) Accelerated sea-level rise from West Antarctica.
Science 306(5694):255–258.
Tomlinson PB (1986) The Botany of Mangroves. Cambridge University Press, United Kingdom.
Triwilaida Intari SE (1990) Factors affecting the death of mangrove trees in the Pedada Strait,
Indragiri Hilir, Riau, with reference to the site conditions. Buletin Penelitian Hutan (531):33–48.
UNEP-WCMC (2006) In the front line: shoreline protection and other ecosystem services from
mangroves and coral reefs. UNEP-WCMC, Cambridge, UK. pp. 33.
Upadhyay VP, Ranjan R, Singh JS (2002) Human-mangrove conflicts: The way out. Curr Sci 83
(11):1328–1336.
Van Lavieren H, Spalding M, Alongi DM, Kainuma M, Clüsener-Godt M, Adeel Z (2012) Securing
the future of mangroves. A policy brief. United Nations University, Institute for Water,
Environment and Health, UNESCO-MAB with ISME, ITTO, FAO, UNEP-WCMC and
TNC. p. 53, Canada.
10 Roles of Mangroves in Combating the Climate Change 255

Verma S, Dubey RS (2003) Lead toxicity induces lipid peroxidation and alters the activities of
antioxidant enzymes in growing rice plants. Plant Sci 164(4):645–655.
Vinit-Dunand F, Epron D, Alaoui-Sossé B, Badot PM (2002) Effects of copper on growth and on
photosynthesis of mature and expanding leaves in cucumber plants. Plant Sci 163(1):53–58.
Vyas P (2012) Biodiversity conservation in Indian Sundarban in the context of anthropogenic
pressures and strategies for impact mitigation. Saurashtra University, Rajkot, India.
Whelan KR, Smith TJ, Cahoon DR, Lynch JC, Anderson GH (2005) Groundwater control of
mangrove surface elevation: Shrink and swell varies with soil depth. Estuaries 28(6):833–843.
Wong YS, Tam NFY, Lan CY (1997) Mangrove wetlands as wastewater treatment facility: a field
trial. In Asia-Pacific Conference on Science and Management of Coastal Environment
(pp. 49–59). Springer, Dordrecht.
Yadav SK (2010) Heavy metals toxicity in plants: an overview on the role of glutathione and
phytochelatins in heavy metal stress tolerance of plants. South Afr J Bot 76(2):167–179.
Yin JH (2005) A consistent poleward shift of the storm tracks in simulations of 21st century climate.
Geophys Res Lett 32(18).
Zhang FQ, Wang YS, Lou ZP, Dong JD (2007) Effect of heavy metal stress on antioxidative
enzymes and lipid peroxidation in leaves and roots of two mangrove plant seedlings (Kandelia
candel and Bruguiera gymnorrhiza). Chemosphere 67(1):44–50.
Zheng WJ, Wang WQ, Lin, P., (1999) Dynamics of element contents during the development of
hypocotyles and leaves of certain mangrove species. J Exp Mar Biol Ecol 233(2):247–257.
Zornoza P, Vázquez S, Esteban E, Fernández-Pascual M, Carpena R (2002) Cadmium-stress in
nodulated white lupin: strategies to avoid toxicity. Plant Physiol Biochemi 40(12):1003–1009.
Role of Mangroves in Pollution Abatement
11
Arumugam Sundaramanickam, Ajith Nithin, and
Thangavel Balasubramanian

Abstract

Mangroves are one of the significant categories of coastal vegetation, distributed


in the shorelines of tropical and subtropical regions and performing a dynamic
role in coastal ecosystems. The recent developments of coastal cities, industriali-
zation, unplanned recreational activities and expansion of aquaculture at man-
grove areas in various regions of the world have generated a threat on such
significant ecosystems. Nevertheless, the mangrove vegetation has some adaptive
features to mitigate the pollution. These ecosystems act as physical, chemical and
biological barriers for the transference of pollutants. They play a vital role in
trapping sediments; assimilate the excess nutrients by phyto- or bioremediation of
toxic substances. The earlier studies revealed that mangroves can act as possible
phytoremediators. They absorb a considerable quantity of toxic metal ions and
store them in various parts such as stems and roots, consequently evade the
transmission of heavy metal ions. In addition, the mangrove associated
microorganisms are responsible for remediation of numerous toxic contaminants
present within the mangroves. In view of the above, the present chapter offers a
comprehensive discussion on the role of mangrove vegetation on abatement of
pollutants.

Keywords

Mangrove ecosystem · Heavy metals · Plastics · Hydrocarbons

A. Sundaramanickam (*) · A. Nithin · T. Balasubramanian


Faculty of Marine Sciences, CAS in Marine Biology, Annamalai University, Parangipettai, India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 257
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_11
258 A. Sundaramanickam et al.

11.1 Introduction

Mangroves forests are composed of 73 species of trees and shrubs that grow along
the coasts of temperate regions and are known for their ecological and socioeco-
nomic importance (FAO 2007; Sandilyan and Kathiresan 2012). These comprise of
woody trees, shrubs and flowering trees which are well adapted to the estuarine
conditions (Spalding et al. 2010). Mangroves have highly adapted morphological
and physiological features that acclimatize to the saline and anoxic environmental
conditions, which enable the increase of biodiversity and biomass productivity
proving to be ecologically and economically beneficial (Mukherjee et al. 2014;
Duke and Schmitt 2015). These benefits include fisheries, maintaining coastal
water quality, shoreline protection, primary production, aquaculture, tourism and
recreation, whereas the most important global service is its role in pollution abate-
ment (Duke et al. 2007; Richards and Fries 2016).
Mangroves have several ecological functions, which contribute to economy and
improve quality of human life (Rodríguez-Rodríguez et al. 2016). Mangrove
ecologies are regarded as barriers prohibiting the release of terrestrial contaminants
into the oceans (Li et al. 2019). Mangroves play an essential role in climate
regulation, soil stabilization and controlling erosion, regulating nutrient cycles and
protecting seagrasses as well as coral reefs (Alongi 2014).
Mangroves act as protective barriers against several natural disasters however this
role is dependent on several factors, viz. tree height, density and species composi-
tion, dominant species make up (roots, stems, branches and foliage), diameter of
roots and trunks, and habitats elevation, nature of channel, pools as well as status of
ecological degradation of the forests (McIvor et al. 2012). These natural barriers are
also effective against storms, cyclones and tidal waves in several tropical regions
(as reviewed by Sandilyan and Kathiresan 2015).
Over the past half century several diseases have occurred globally where the chief
causative agents have been heavy metals. Heavy metal pollution especially in the
marine environments is a serious threat owing to their persistence, toxicity and
non-degradable nature. Heavy metals originate from mining, aquaculture, smelting,
agriculture, printing and from industrial sources, viz. petrochemical and electronic
industry while discharge of untreated municipal wastes into the marine environment
is also one of the sources (Paz-Alberto and Sigua 2013). All discharges from these
sources enter the marine habitats where they get accumulated in the organisms and
are biomagnified to higher organisms (Rainbow and Luoma 2011). The primary
sources for pollution in the mangroves include trace metals, oil residues, pesticides
and untreated wastewater. These pollutants affect photosynthesis and growth of
plants ultimately leading to mortality (Lovelock et al. 2009).
In short, it can be stated that mangroves offer numerous benefits; however, these
may become retarded due to anthropogenic activities that affect mangroves. The
following section describes some of the services rendered by the mangroves.
11 Role of Mangroves in Pollution Abatement 259

11.2 Solutions from the Mangroves to Pollution Problems

Mangrove ecosystem is of vast ecological, traditional and socioeconomic values


(Kathiresan and Bingham 2001). They are actively involved in controlling of marine
toxicants, such as heavy metals, pesticides, oils & greases, nutrients and plastics
(Wang Q et al. 2019b). They are also involved in carbon sequestration and coastal
production. Furthermore, the mangrove residing microorganisms are actively
involved in degradation of a wide array of pollutants (Fig. 11.1).

11.2.1 Reducing Nutrient Loads

The ocean waves and currents constantly change the shapes of coastlines as sediment
accretion and removal occurs during each wave. This may cause soil erosion along
the coasts and this is where mangroves prove to be essential as they help to prevent
soil erosion and bind the soil. The mangroves decrease wave intensity thereby
reducing sediment discharge into the oceans. Decreased wave flow enables the
deposition of bulk mass of sediment along the shores. Mangroves provide organic
matter such as leaves and twigs which bind the sediments together and prevent
subsidence (Spalding et al. 2014).
The domestic and industrial effluents are discharged into the estuaries; however,
the mangroves are capable of purifying and improving the water quality thereby
minimizing the effects of these discharges. However, if the pollutant concentrations
in these discharges increase, it may harm the mangroves and surrounding
environments (Sri Dattatreya et al. 2018). Mangroves can maintain the quality of

Fig. 11.1 Mangroves on pollution abatements


260 A. Sundaramanickam et al.

Table 11.1 Mangrove sources in carbon sequestration


Role in pollution
Source Common/Scientific name abatement References
Plant Ceriops decandra Carbon Clough (1998), Alongi
sequestration et al. (2000)
Plant Rhizophora stylosa and Carbon Alongi et al. (2003)
Avicennia marina sequestration
Plant Sonneratia caseolaris Carbon Alongi et al. (2004b)
sequestration
Plant Rhizophora apiculata, Carbon Alongi et al. (2004a),
sequestration Alongi (2009)
Plant Kandelia candel Carbon Alongi et al. (2005)
sequestration
Soil – Carbon Ren et al. (2010)
sequestration
Sediment Mangrove peat Carbon Ezcurra et al. (2016)
sequestration

estuaries by eliminating pollutants from effluents and transforming them into useful
compounds (Wu et al. 2008).
The sedimentary characteristics such as soil grain, salinity and soil nutrients are
also essential in regulating the benthic macro-organisms (Safahieh et al. 2012).
Regions with fluctuating soil nutrients prevent growth of soil associated fauna.
Hence, the soil nutrients are also detrimental in the mangrove productivity
(Kathiresan and Bingham 2001). Dissanayake and Chandrasekara (2014) clearly
explained the importance of physicochemical characteristics on the abundance of
benthic organisms.
Nitrates and ammonia are two important compounds that have a role to play in
regulating water quality. Nitrates are regularly maintained by the nitrogen cycle
where nitrates are converted to nitrites, which are consumed by several organisms as
nitrogen source. Excess concentrations of ammonia may prove toxic to the
surrounding. During nitrification, ammonia is converted to nitrates by ammonia
oxidizing bacteria such as Nitrosomonas sp., Nitrococcus sp. and Nitrosospira
sp. (Purkhold et al. 2000; Ward and O’Mullan 2002; Purkhold et al. 2003). These
species are abundantly found in all ecosystems including the mangroves. Moreover,
Nitrosopumilus maritimus also plays a decisive role in eliminating ammonia from
the mangrove ecosystems (Cao et al. 2011) (Table 11.1).

11.2.2 Reduction of Petroleum Hydrocarbons

Mangroves are known to attract tourists; therefore, the estuaries associated to the
mangroves are subjected to constant boating activities apart from regular fishing.
Such activities may yield oil spillage. Effluents from the paper and petroleum
industries also contribute to oil pollution in mangroves (Lovelock et al. 2009).
11 Role of Mangroves in Pollution Abatement 261

Additionally, accidental spillage from oil containers also contributes to contamina-


tion of the mangrove system. These contaminations might cause several environ-
mental effects on natural resources. The severity of oil spillages is dependent on
volume of discharge, type of discharge and environmental conditions, which varies
from season to season (Wang and Stout 2007). Total petroleum hydrocarbons
(TPHs) are toxic to the aquatic organisms as they belong to one of the highly toxic
group of persistent organic carbons. TPHs are released into various environments as
a consequence of extraction, refining and consumption carried out by the petroleum
industries (McNioll and Baweja 1995; USEPA 2000). Adequate and effective
strategies are required to combat these hydrocarbons which can be achieved by
attenuation, biostimulation and bioaugmentation within the soil which is possible in
the mangroves. Mangrove areas are subjected to oil spills which affect the soils,
plant surfaces and cause sub-lethal and lethal effects (Swan et al. 1994; Duke et al.
1999; Duke and Burns 2003; NOAA 2014). Such disruptions may disturb the
socioeconomic benefits rendered by the mangroves.
The ubiquitous hydrocarbons are persistent in the environment due to anthropo-
genic activities which include oil and gas operations. Polycyclic aromatic
hydrocarbons (PAHs) are the most common form of organic pollutant persistent in
the environment (Tam and Wong 2008; Lu et al. 2011). Oil spillage may transpire
during transportation of crude oil, discharges from petroleum products and during
emissions from industries (Tam and Wong 2008). These petroleum hydrocarbons
adhere to soil particles, hence degradation is tedious (Barathi and Vasudevan 2001).
Bioremediation of TPHs is a time consuming process since most of the TPH
compounds are of high molecular weight (McNioll and Baweja 1995; Huang et al.
2005). Numerous environmental and physicochemical factors including nutrient
availability, sunlight, temperature, soil grain size as well as microbial community
influence the bioremediation of TPHs (Colombo et al. 2005; Lacerda 2006; Moreira
et al. 2011). Some TPH removal by ex-situ methods such as Thermal Desorption,
Chemical Oxidation and Incineration have been conducted with limited success. In
situ technology involving air sparging, land farming, biosparging, bioventing, bio-
remediation, and phytoremediation has been carried out successfully (Seabra 2008).
Among these, phytoremediation in combination with biotechnology has provided
excellent results (Espinosa et al. 2005; Huang et al. 2005; Parrish et al. 2005;
Doumett et al. 2008). Mangrove sediments enable to capture, transform and
rhizodegrade TPHs during phytoremediation. Mangrove soils contain microbial
communities suitable for rhizodegradation of TPHs and make it less bio-available
to plants and other biota (Kamath et al. 2004). A former study (Moreira et al. 2011)
has exhibited that phytoremediation using mangrove plants is more efficient than
bioremediation.
Once an oil spill has occurred, the plants take some time to grow. However, plant
recovery is faster than animal recovery as animals require the toxicity to be
completely eliminated (Salmo and Duke 2010). Similarly, new seedlings can be
planted only after the complete removal of toxicity (Duke 2001). Erosion and
elevation are also important components which influence mangrove growth after
an oil spill (Lewis 2005). Biotic and abiotic components are also essential to
262 A. Sundaramanickam et al.

determine the mangrove recovery after each oil spillage. The biotic components
include propagules and tolerances of the species towards oil spillage, whereas the
chief abiotic components include erosion and elevation (Duke 2016).

11.2.3 Pesticides Degradation

Plants have been progressively used in bioremediation as a probable solution to


pollution. Certain plants including mangroves are known to eliminate heavy metals
from soil and water, while some grasses can provide a remedy for problems due to
petroleum hydrocarbons (USEPA 2000). One such problem is pesticide spillage
which may occur during its manufacture, distribution, formulation and application in
the field. Excess levels of pesticide in soil may cause several problems such as
leaching and running off into ground and surface water, respectively, while also
affecting the soil organisms.
It is well known that bioremediation using plants is a time consuming process.
During this extended time the pesticides may gradually be eliminated or become less
bio-available to the biota. In certain cases, the pesticides may transform the soil
structure which may cause drastic changes to such environments. A study (Belden
et al. 2004) was conducted to upsurge the phytoremediation efficiency by employing
prairie grasses along with mulberry trees. The study showed increased efficiency
which suggested that similar techniques can be employed to the mangroves ecology
to decrease the pollution due to pesticides.
Persistent Organic Pollutants (POPs), such as DDTs, chlordanes, endosulfans,
etc. are ubiquitously found in waters or sediments in mangroves of the tropical areas.
Even though these contaminants have been detected in various ecosystems globally,
very little knowledge is available on the consequences of these compounds on the
mangrove ecosystems (Bayen 2012).
The estuaries bring in these toxicants that are sprayed on to the agricultural fields.
Since some estuaries inundate the mangroves, certain pesticides are deposited
(Chilundo et al. 2008; Nagelkerken et al. 2008). However, the exact composition
of these pollutants and their effects on various metabolic processes of the mangroves
are largely unknown. This is where biosurfactants come into and play to combat the
effects of several pollutants. Biosurfactants are bio-molecules that are retrieved from
bacteria and fungi (Maier and Soberon-Chavez 2000; Mukherjee et al. 2006) and aid
as natural choices to remove surface chemicals (Banat et al. 2010; Cameotra et al.
2010). Some studies have proved the ability of bio-surfactant as an effective surface
agent (Kheiralla et al. 2013). However, their significance in complete degradation of
pesticides is still under research.

11.2.4 Heavy Metals Remediation

Mangroves are essential in combating pollutants as they are known to sink anthro-
pogenic pollutants and toxicants. In the mangrove soils, metal contaminants
11 Role of Mangroves in Pollution Abatement 263

generally accumulate in the superficial water and pore water, and in solid phases like
abiotic and biotic components (Lewis et al. 2011). A former study has highlighted
the depletion of oxygen owing to the inundation of mangroves (Bayen 2012). The
presence of sulphides and decreased oxygen concentration may contribute to
co-precipitation of heavy metals in mangrove soils. In addition to sulphides, physi-
cochemical concentrations are also associated with trace metals (Bayen 2012). Some
mangrove plants like Rhizophora sp. can accumulate up to 95% more trace metals
than other plants, thereby enabling the mangroves to retain and prevent heavy metals
from infiltration into the environment (Silva et al. 1990). Salinity exchanges in the
estuaries are also responsible for metal retention. The major cations such as Na, K,
Ca and Mg intensify in the water during increased salinity and these ions bind with
heavy metals (Laing et al. 2009).
Phytoremediation employing mangroves is used to remove/detoxify contamina-
tion due to heavy metals. The mangroves are employed to eliminate heavy metals
and certain harmful organic compounds from soil while some volatile-organic
compounds can be removed from the groundwater (Paz-Alberto and Sigua 2013).
An aggregate of 33 species of mangrove plants are adapted at consuming heavy
metals. Avicenna marina has the utmost potential to degrade heavy metals than other
mangrove plants. In general, roots are known to take up more heavy metals than the
aerial parts, hence the upper plant tissues are not good indicators of metal pollution.
However, in a study A. marina leaves had consumed 10% more metals than roots.
The study had revealed that A. marina plants act as an indicator of Cu, Zn and Pb, as
these metals were preferred (MacFarlane and Burchett 2002; MacFarlane et al.
2003).
The biodiversity of mangroves is also dependent on magnitude of pollution in the
particular environment. Increased pollution causes the rise of pollution tolerant
species while other species do not grow in these areas. A. marina plants are identified
as superlative pollution tolerant species hence may be abundantly present in areas of
high pollution. In simple words, it can be said that pollution hampers the mangrove
biodiversity (Maiti and Chowdhury 2013).
The dispersion of metals varies with depth, coastal distance and typology of
above ground mangrove vegetation (Kehrig et al. 2003; Marchand et al. 2006;
Chatterjee et al. 2009). Trace metals may arise from natural and anthropogenic
sources hence the exact source of the respective metals is difficult to differentiate.
The anthropogenic sources of metals include mining (Marchand et al. 2006), gas
(Antizar-Ladislao et al. 2011; Kruitwagen et al. 2008) and textile dye industries
(Machado et al. 2002). The metals are brought to mangroves by rivers (Kehrig et al.
2003), or tidal exchange (Kruitwagen et al. 2008) or by atmospheric transfer
(Rumbold et al. 2011). Mangroves efficiently retain the metals in plant tissues,
hence reduce the metal bioavailability to the fauna.
264 A. Sundaramanickam et al.

11.2.5 Act as a Plastics Trap

Plastics are widely used in our day to day lives which have led to the improper
disposal of plastics and the resultant mismanagement of wastes. The mismanaged
plastic wastes accumulate in the beaches, mangroves and other aquatic ecosystems
(Lebreton et al. 2017). Plastic debris has always garnered attention due to its low
degradation rate, accumulation and bioavailability to the aquatic organisms (Iñiguez
et al. 2016). Approximately 60–90% of marine debris constitutes plastics owing to
its high production (Li et al. 2016). The plastic conundrum arises from mismanage-
ment which according to a study ranged between 4.8 and 12.7 million metric tons in
2010 (Jambeck et al. 2015). In nature, compounds such as phthalates and persistent
organic pollutants may adhere to plastic debris by virtue of which bioaccumulation
and biomagnifications may occur (Moore and Phillip 2011; Jang et al. 2016; Clukey
et al. 2018).
The larger plastic debris persists in these marine systems for a long period of time,
thereby undergoing breakdown to minor particles referred to as microplastics which
pose a bigger threat than larger plastics (Li et al. 2016). Microplastics are ingested by
various organisms such as mollusks, crustaceans, fishes and others due to its smaller
size and attractive colours (Wright et al. 2013; Antão-Barboza et al. 2018).
Microplastics adsorb POPs and heavy metals leading to the bioaccumulations and
biomagnifications, thereby causing their entry into the food chain (Ríos et al. 2007;
Andrady 2011; Kühn et al. 2015; Bennecke et al. 2016; Wang et al. 2016; Massos
and Turner 2017). Microplastics can release toxic materials which are consciously
added during the manufacture of plastics, thereby causing environmental threat
(Gallo et al. 2018). Moreover, microplastics are potential transporters of exotic
species (Rech et al. 2016) and pathogenic microorganisms (Kovač Viršeka et al.
2017). However, information on the complete effects of microplastics on the marine
environment is still insufficient (Auta et al. 2017).
Estuaries provide several ecological benefits to man (Barbier et al. 2011); how-
ever, they are constantly polluted due to domestic sewage discharges which act as an
entry pathway for microplastics into the marine environment (Browne et al. 2011;
Cesa et al. 2017). Cordeiro and Costa (2010) explained that the litter in the
mangroves originated from terrestrial and freshwater sources. Mangroves are unique
with regard to their retention and transformations capabilities of land-based litter and
pollutants due to intense biomass and productivity (Fourqurean et al. 2012; Li et al.
2016; Booth and Sear 2018), hence are garnering much research attention (Martin
et al. 2019). Some of the terrestrial origins of microplastics include personal care
goods, for example, facial scrubs, cleansers, creams and toothpaste which reach the
aquatic systems through municipal sewage and wastewater treatment plants (Auta
et al. 2017; Murphy et al. 2016).
Mangroves have a complex root system consisting of pneumatophores and prop
roots which enhance turbulence and wave energy (Horstman et al. 2014; Norris et al.
2017). These roots act as a filter system which traps materials transported by waves
and currents. An earlier report exhibited the different retention capacities of various
objects dumped in the mangrove forests (Ivar do Sul et al. 2014).
11 Role of Mangroves in Pollution Abatement 265

Plastics trapped by mangrove roots cause external damage which affects the tree
and the dependent fauna. These plastics also prevent gas exchange, meanwhile
release absorbed chemicals to the mangrove ecosystem (Cole et al. 2011). A number
of natural factors such as estuarine hydrodynamics, mangrove characteristics, sedi-
ment grain size along with human activities such as tourism, mariculture and coastal
dumping are some of the determining factors for plastic distribution and retention in
the mangroves (Ivar do Sul et al. 2014; Lima et al. 2016). Ivar do Sul et al. (2014)
explained that retention capacity of the mangroves is reliant on the hydrodynamics
of the plastic materials while also mentioned that plastic bags are retained the most
owing to its hydrodynamic property. Thus the hydrodynamics of semi-enclosed
mangrove habitats are a regulating factor for the distribution of plastics (Boelens
et al. 2018).
Polycyclic aromatic hydrocarbon (PAH) is a type of organic contaminant found
in sediments which are transferred to organisms (Ramdine et al. 2012). The hydro-
phobic nature of plastics enables POPs such as dichlorodiphenyltrichloroethane
(DDTs), polycyclic aromatic hydrocarbons (PAHs) and polychlorinated biphenyls
(PCBs) to get adsorbed onto microplastic surfaces (Endo et al. 2005; Teuten et al.
2007; Frias et al. 2010; Rochman et al. 2013). These adsorbed compounds are
consumed when the microplastics are ingested and released into the tissues of
organisms (Engler 2012). Similarly, toxic monomers (Saido et al. 2009) and some
plastic additives like bisphenol A and phthalates (Fries et al. 2013) leach from the
microplastics and affect the marine organisms and environment.
Polyethylene (PE), Polypropylene (PP), Polyvinyl chloride (PVC), rayon and
polyester (PES) are the some of the regular polymers recovered from the mangroves
(Li et al. 2018; Zhu et al. 2019). However, Ajith et al. (2020) reviewed that apart
from these Polystyrene (PS) and nylon are also commonly found in the marine
environment and this could enter the mangroves. PP has a wide use in industrial and
household applications which include textiles, stationery, packaging, automobile
parts and laboratory equipment (Gewert et al. 2015). PE is used in packaging, plastic
bottles, bags (Zhang et al. 2016) and fishing gears (Chen et al. 2018; Wang T et al.
2019a). The other polymers are also dominant, but their origins are not documented
well (Ajith et al. 2020).
Plastic polymers are recalcitrant in nature which prohibits rapid microbial degra-
dation enabling the long-term persistence of plastics in the environment (Longo et al.
2011). Additionally, the plastic surfaces act as substrates for microbial colonization
(Dussud and Ghiglione 2014). During colonization, microbes initiate biodegradation
by secreting extracellular enzymes (Eich et al. 2015; Sekhar et al. 2016). Certain
microorganisms produce biofilms which also influence biodegradation of plastics
(Sheik et al. 2015; Eich et al. 2015; Jeon and Kim 2016). Microbial enzymes are
important in biodegradation of plastic substances, since some enzymes utilize
plastics as carbon source (Auta et al. 2018).
266 A. Sundaramanickam et al.

11.2.6 Mangrove Residing Microorganisms

Mangroves occur at the interface of marine and terrestrial ecosystems which exem-
plify a prosperous biodiversity. Microorganisms are essential components of the
mangrove ecosystem; they play crucial roles in the creation and upholding of this
biosphere. They also supply biotechnologically potential and valuable products
(Thatoi et al. 2013). This ecosystem offers a typical environment harbouring differ-
ent assemblages of microbes, for instance, actinomycetes, bacteria, cyanobacteria,
fungi, protozoa, algae, etc. (Sen and Naskar 2003). Among them, bacteria and fungi
provide a noteworthy contribution towards reducing the pollutants (Fig. 11.2). These
microbes contributing in different steps of decomposition and mineralization of
mangrove leaves among other litters significantly contribute to the productivity of
the mangrove environment. They play an important role in recycling nutrients,
involve in carbon sequestration and devastate pollutants. The microorganisms
degrade the pollutants by their enzymatic action while enzymes act as biocatalysts
facilitating degradation. Among the mangrove residing microbes, the bacteria are the
most dominated group as they are several fold higher than those of fungi (Kathiresan
and Qasim 2005). Microbes utilize a broad range of substances to produce energy
and nutrients from the surrounding environment for the growth and multiplications
of cells. The mangroves are the extremely productive ecosystems hold rich nutrients
which support the growth and multiplication of diverse groups of microorganisms.
Domestic wastes and industrial effluents are expelled into the estuaries and
channels associated with the mangroves which may release excess nutrients in the
surface water and soil. This may instigate eutrophication and algal blooms whilst
excess nitrogenous wastes may also be harmful. Certain aquatic biota including
bacteria and algae enable mangroves to eradicate the harmful nitrogenous wastes and
unused nutrients present on the surface water and soil. This purified water can be
reused in aquaculture (Spalding et al. 2014).
Mangroves are one of the main ‘hotspots’ for marine fungi (Shearer et al. 2007).
Although the trunks and aerating roots are permanently or intermittently submerged
the other parts remain free of salt water interference. The lichens and terrestrial fungi

Fig. 11.2 Mangrove residing


microorganisms and their
dynamic role in pollutants
reduction
11 Role of Mangroves in Pollution Abatement 267

inhabit the dry parts of plant while marine groups live in the bottom region; middle
region overlies the marine fungi (Sarma and Hyde 2001).
Bacteria are well known to degrade petroleum hydrocarbons as they have been
employed successfully during oil spills (Rahman et al. 2003; Brooijmans et al.
2009). Biodegradation of petroleum hydrocarbons relies on a number of factors,
such as the type of compound and availability to microorganisms. Some bacteria
consume hydrocarbons as their primary source of supplement (Yakimov et al. 2007).
However, high molecular weight hydrocarbons can never be degraded (Atlas and
Bragg 2009). Mangrove sediments host a variety of bacteria and fungi among which
certain strains may have ability to degrade hydrocarbons. Some of these species
include Bacillus subtilis, Pseudomonas fluorescens and Rhodococcus erythropolis
(Duke 2016).
The B. subtilis is capable of degrading hydrocarbons and can survive in extreme
conditions of heat. These spores are advantageous as they have the potential to
produce a natural biosurfactant called surfactin which is synthesized in combination
with crude oil (Queiroga et al. 2003). The strain P. fluorescens is also known to
produce biosurfactants which have the potential to degrade hydrocarbons (Kaczorek
and Olszanowski 2011). The bacterium R. erythropolis has a unique metabolism
which consumes carbon from hydrocarbons, thereby degrading various compounds
such as chlorinated phenols, steroids, lignin, coal and crude oil (Brandao et al. 2003).
The mangrove fungus has attracted attention of scientists due to their wealthy and
diverse potentials. Wu (1993) screened 15 species of fungi from mangroves regions
of Tanshui Estuary, Taiwan. He observed that they can secrete different enzymes
that have potential to decompose mangrove litter. Xin et al. (2002) found that the
marine white-rot fungus can degrade 50% of lignin incubated with entire sugarcanes.
Microorganisms adapt to all kinds of environment (Brooks et al. 2011; Aujoulat
et al. 2012) and possess the potential to transform plastic polymers. The mangrove
residing microbes can metabolize the pollutants which might enhance biotransfor-
mation and subsequent degradation (Luigi et al. 2007). Several studies have articu-
lated their results on microbial degradation of plastic polymers. PET was degraded
by Ideonella sakaiensis 201-F6 (Yoshida et al. 2016); degradation polystyrene by
Pseudomonas sp. and Bacillus sp. (Mohan et al. 2016); the attenuation of molecular
weight of PE under test conditions by a fungus Zalerion maritimum (Paco et al.
2017); degradation of PE by Bacillus cereus (Sowmya et al. 2014); Kocuria
palustris M16, Bacillus subtilis H1584 and Bacillus pumilus M27 (Harshvardhan
and Jha 2013). A number of other bacteria are also involved in polymer-degradation
which includes Pseudomonas stutzeri, Alcaligenes faecalis, Pseudomonas putida,
Staphylococcus sp., Streptomyces sp. and Brevibacillus borstelensis (Ghosh et al.
2013; Caruso 2015). Therefore, it may be affirmed that microbial degradation of
plastics is an eco-friendly approach (Boelens et al. 2018). Mangroves host a wide
variety of microbes (Kathiresan 2003; Thatoi et al. 2012) due to the suitable
conditions of temperature, pH and salinity (Ghizelini et al. 2012).
268 A. Sundaramanickam et al.

11.2.7 Carbon Sequestration

Mangroves are the principal carbon-rich habitats of the tropic regions (Kristensen
et al. 2008; Donato et al. 2011; Pendleton et al. 2012; Siikamäki et al. 2012b; Alongi
2014). Typically, mangroves garner quadruple amounts of carbon than other floral
habitats (Donato et al. 2011). However, mangrove deforestation occurs at a signifi-
cantly higher rate than the loss of other types of forests (FAO 2007). Mangrove
deforestation causes carbon emission and reduces carbon sequestration (Alongi
2014). Carbon emissions coupled with other greenhouse gases (CH4, N2O) may
have contributed to climate change (IPCC 2014). To combat climate change, it is
necessary to prevent deforestation of mangroves and subsequent carbon emission
(Kristensen et al. 2008; Donato et al. 2011; Houghton 2012; Alongi 2014).
Blue carbon emissions can be reduced significantly by the mangrove restoration
programs (Pendleton et al. 2012; Siikamäki et al. 2012a). Through mangrove
restoration, several environmental activities such as photosynthesis, respiration,
water purification, decomposition and predation may also be restored (Cardinale
et al. 2011; Parrotta et al. 2012). Blue carbon sequestration through mangrove
restoration is gaining popularity due to its importance in climate change (Le 2008;
Joffre and Schmitt 2010). Blue carbon sequestration enables to promote and safe-
guard mangroves as well as provide a solution to climate change (Robledo et al.
2004). However, mangrove restoration is possible only with an ecologically benefi-
cial design and socioeconomic support for its construction (Mazda et al. 2006).
In the early days, carbon accumulation was conducted as short-term studies and
the results expressed that a large amount of carbon is stocked in the soil. However,
over time it was proved that short-term studies do not portray the actual picture of
carbon sequestration. This is evidenced in a study where short-term assessment of
137Cs, 222Th and 210Pb radionuclides suggested a high carbon accumulation rate
of 1.25 tC ha 1 year 1 (Alongi et al. 2004). However, ground-truth studies and
imaging studies have showed a lesser carbon accumulation rate of 0.5 tC ha 1
year 1 (Alongi et al. 2004). This is due to simultaneous accretion and erosion of
mangroves forests. The actual picture of carbon sequestration can be projected only
by carrying out long-term studies (Alongi 2011).
Despite the numerous services provided by the mangroves, they are constantly
subjected to human pressures as over 120 million people are known to inhabit areas
surrounding the mangroves (UNEP 2014). A large area of mangroves is converted to
aquaculture as this provides an income and means of food. The loss of mangroves
may lead increased concentration of carbon in the atmosphere. Mangroves are
essential sources of blue carbon sinks. Blue carbon is the organic carbon stored,
sequestered, and released from mangroves, salt marshes, seagrasses and other marine
environments (Nellemann et al. 2009; McLeod et al. 2011; Murray et al. 2011;
Pendleton et al. 2012; Duarte et al. 2013; Alongi 2014). Annually, mangroves
contribute to 30% of global carbon sequestration (Siikamäki et al. 2012a).
Mangroves need to be restored by carrying out effective restoration programs such
as Reduced Emissions during deforestation and degradation (REDD) (Ahmed and
Glaser 2016) (Table 11.2).
11 Role of Mangroves in Pollution Abatement 269

Table 11.2 Mangrove sources in regulation of nutrient load/water quality


Source Common/Scientific name Role in pollution abatement References
Soil – Treatment of water quality Tam and Wong
(1995)
Plant Kandelia candel, Bruguiera Salinity and nutrient Ye et al. (2001)
gymnorrhiza removal
Plant Kandelia candel Treatment of municipal Wu et al. (2008)
wastewater
Soil Glomalin related protein Improves water quality Wang et al.
(2018)

11.2.8 Coastal Protection

Mangroves act as natural barriers that prevent storms and cyclones or resist their
intensity, thereby minimizing the damage on the inner areas. The constant exchange
of sediments between oceans and mangroves strengthens the mangrove soil thereby
preventing soil erosion. The dense mangroves growing along the coastal regions can
act as a natural shield, help to reduce the impact of natural calamities like cyclone,
tsunami and coastal storm by engrossing the waves’ energy because of the flexibility
of the mangrove stem. The study conducted by Kathiresan and Rajendran (2005)
scientifically proved that the mangroves play a significant role in coastal protection.
During excessive high tides or tsunamis, the coasts are protected by the mangroves,
which reduce the wave intensity, hence the waves do not enter too far into the
terrestrial areas (Spalding et al. 2014).

11.3 Conclusion

Though the mangroves provide solutions for several pollutants, the tolerance of
mangroves towards toxicity necessitates testing to comprehend the toxicity threshold
level of mangroves. This may be accomplished by testing the various stages of
mangroves right from their juvenile stages. By such experiments, the information on
tolerance of mangroves towards toxicity can be documented. Such information
extends knowledge of the scientific community to apprehend effects of chemical
and non-chemical stressors.

References
Ahmed N, Glaser M (2016) Coastal aquaculture, mangrove deforestation and blue carbon
emissions: is REDD+ a solution?. Mar Policy 66:58–66.
Ajith N, Arumugam S, Parthasarathy S, Manupoori S, Janakiraman S (2020) Global distribution of
microplastics and its impact on marine environment—a review. Environ Sci Pollut Res
27:25970–25986.
Alongi DM (2011) Carbon payments for mangrove conservation: ecosystem constraints and
uncertainties of sequestration potential. Environ Sci Pol 14:462–470.
270 A. Sundaramanickam et al.

Alongi DM (2014) Carbon cycling and storage in mangrove forests. Annu Rev Mar Sci 6:195–219.
Alongi DM, Sasekumar A, Chong VC, Pfitzner J, Trott LA, Tirendi F, Dixon P, Brunskill GJ (2004)
Sediment accumulation and organic material flux in a managed mangrove ecosystem: estimates
of land–ocean–atmosphere exchange in peninsular Malaysia. Mar Geol 208:383–402.
Alongi DM, Tirendi F, Trott LA, Xuan TT (2000) Benthic decomposition rates and pathways in
plantations of the mangrove Rhizophora apiculata in the Mekong delta, Vietnam. Mar Ecol
Prog Ser 194:87–101.
Alongi DM (2009) The energetics of Mangrove forests. Springer, Amsterdam, The Netherlands.
Alongi DM, Clough BF, Dixon P, Tirendi F (2003) Nutrient partitioning and storage in arid-zone
forests of the mangroves Rhizophora stylosa and Avicennia marina. Trees 17(1):51–60.
Alongi DM, Pfitzner J, Trott LA, Tirendi F, Dixon P, Klumpp DW (2005) Rapid sediment
accumulation and microbial mineralization in forests of the mangrove Kandelia candel in the
Jiulongjiang estuary, China. Est Coast Shelf Sci 63:605–618.
Alongi DM, Wattayakorn G, Tirendi F, Dixon P (2004a) Nutrient capital in different aged forests of
the mangrove Rhizophora apiculata. Bot Mar 47(2):116–24.
Alongi DM, Sasekumar A, Chong VC, Pfitzner J, Trott LA, Tirendi F, Dixon P, Brunskill GJ
(2004b) Sediment accumulation and organic material flux in a managed mangrove ecosystem:
estimates of land–ocean–atmosphere exchange in peninsular Malaysia. Mar Geol 208:383–402.
Andrady A (2011) Microplastic in the marine environment. Mar Pollut Bull 62(8):1596–1605.
Antão-Barboza L, Vethaak AD, Lavorante B, Lundebye A, Guilhermino L (2018) Marine
microplastic debris: an emerging issue for food security, food safety and human health. Mar
Pollut Bull 133:336–348.
Antizar-Ladislao B, Sarkar SK, Anderson P, Peshkur T, Bhattacharya BD, Chatterjee M, Satpathy
KK (2011) Baseline of butyltin contamination in sediments of Sundarban mangrove wetland
and adjacent coastal regions, India. Ecotoxicol 20(8):1975–1983.
Atlas RM, Bragg J (2009) Bioremediation of marine oil spills: when and when not-the Exxon
Valdez experience. Microb Biotechnol 2(2):213–221.
Aujoulat F, Roger F, Bourdier A, Lotte A, Lami B, Marchandin H, Jumas-Bilak E (2012) From
environment to man: gene evolution and adaptation of human opportunistic bacterial pathogens.
Genes 3:191–232.
Auta HS, Emenike CU, Fauziah SH (2017) Distribution and importance of microplastics in the
marine environment: a review of the sources, fate, effects, and potential solutions. Environ Int
102:165–176.
Auta HS, Emenike CU, Jayanthi B, Fauziah SH (2018) Growth kinetics and biodeterioration of
polypropylene microplastics by Bacillus sp. and Rhodococcus sp. isolated from mangrove
sediment. Mar Pollut Bull 127:15–21.
Banat I, Franzetti A, Gandolfi I, Bestetti G, Martinotti MG, Fracchia L, Smyth TJ, Marchant R
(2010) Microbial biosurfactants production, applications and future potential. Appl Microbiol
Biotechnol 87:427–444.
Barathi S, Vasudevan N (2001) Utilization of petroleum hydrocarbons by Pseudomonas fluorescens
isolated from a petroleum-contaminated soil. Environ Intern 26(5–6):413–416.
Barbier EB, Hacker SD, Kennedy C, Koch EW, Stier AC, Silliman BR (2011) The value of
estuarine and coastal ecosystem services. Ecol Monogr 81:169–193.
Bayen S (2012) Occurrence, bioavailability and toxic effects of trace metals and organic
contaminants in mangrove ecosystems: A review. Environ Intern 48:84–101.
Belden JB, Clark BW, Phillips TA, Henderson KL, Arthur EL (2004) IOWA State University,
Digital Repository, Entomology Publicationshttp, pp. 156–157.
Bennecke D, Duarte B, Paiva F, Cacador I, Canning-Clode J (2016) Microplastic as vector of heavy
metal contamination from the marine environment. Estuar Coast Shelf Sci 178:189–195.
Boelens T, Schuttelaars H, Schramkowski G, Mulder TD (2018). The effect of geometry and tidal
forcing on hydrodynamics and net sediment transport in semi-enclosed tidal basins. Ocean
Dynam. 68:1285–1309.
11 Role of Mangroves in Pollution Abatement 271

Booth DJ, Sear J (2018) Coral expansion in Sydney and associated coral-reef fishes. Coral Reefs
37:995.
Brandao PF, Clapp JP, Bull AT (2003) Diversity of nitrile hydratase and amidase enzyme genes in
Rhodococcus erythropolis recovered from geographically distinct habitats. Appl Environ
Microbiol 69:5754–5766.
Brooijmans RJW, Pastink MI, Siezen RJ (2009) Hydrocarbon-degrading bacteria: the oil-spill
clean-up crew. Microb Biotechnol 2(6):587–594.
Brooks AN, Turkarslan S, Beer KD, Lo FY, Baliga NS (2011) Adaptation of new cells to new
environments. Wiley Interdiscip. Rev Syst Biol Med 3(5):544–561.
Browne MA, Crump P, Niven SJ, Teuten E, Tonkin A, Galloway T, Thompson R (2011)
Accumulation of microplastic on shorelines worldwide: sources and sinks. Environ Sci Technol
45:9175–9179.
Cameotra SS, Makkar RS, Kaur J, Mehta SK (2010) Synthesis of biosurfactants and their
advantages to microorganisms and mankind. Adv Exp Med Biol 672:261–280.
Cao H, Li M, Hong Y, Ji-Dong Gu (2011) Diversity and abundance of ammonia-oxidizing archaea
and bacteria in polluted mangrove sediment. Syst Appl Microbiol 34:513–523.
Cardinale BJ, Matulich KL, Hooper DU, Byrnes JE, Duffy E, Gamfeldt L, Balvanera P, O’connor
MI, Gonzalez A (2011) The functional role of producer diversity in ecosystems. Am J Bot 98
(3):572–592.
Caruso G (2015) Plastic degrading microorganisms as a tool for the bioremediation of plastic
contamination in aquatic environments. J Pollut Eff Control 3:112.
Cesa FS, Turra A, Baruque-Ramos J (2017) Synthetic fibers as microplastics in the marine
environment: a review from textile perspective with a focus on domestic washings. Sci Total
Environ 598:1116–1129.
Chatterjee M, Massolo S, Sarkar SK, Bhattacharya AK, Bhattacharya BD, Satpathy KK Saha S
(2009) An assessment of trace element contamination in intertidal sediment cores of Sunderban
mangrove wetland, India for evaluating sediment quality guidelines. Environ Monit Assess 150
(1–4):307.
Chen M, Jin M, Tao P, Wang Z, Xie W, Yu X, Wang K (2018) Assessment of microplastics derived
from mariculture in Xiangshan Bay, China. Environ Pollut 242:1146–1156.
Chilundo M, Kelderman P, O’Keeffe JHO (2008) Design of a water quality monitoring network for
the Limpopo river basin in Mozambique. Phys Chem Earth 33(8–13):655–665.
Clough BF (1998) Mangrove forest productivity and biomass accumulation in Hinchinbrook
Channel, Australia. Mangroves Salt Marshes 2:191–198.
Clukey KE, Lepczyk CA, Balazs GH, Work TM, Li QX, Bachman MJ, Lynch JM (2018) Persistent
organic pollutant in fat of three species of Pacific pelagic long line caught sea turtles: Accumu-
lation in relation to ingested plastics marine debris. Sci Total Environ 610:402–411.
Cole M, Lindeque P, Halsband C, Galloway TS (2011) Microplastics as contaminants in the marine
environment: a review. Mar Pollut Bull 62:2588–2597.
Colombo JC, Barreda A, Bilos C, Cappelletti N, Migoya MC, Skorupka C (2005) Oil spill in the
Rıo de la Plata Estuary, Argentina: 2-hydrocarbon disappearance rates in sediments and soils,
Environ Pollut 134:267–276.
Cordeiro CAMM, Costa TM (2010) Evaluation of solid residues removed from a mangrove swamp
in the S~ao Vicente Estuary, SP, Brazil. Mar Pollut Bull 60(10):1762–1767.
Dissanayake N, Chandrasekara U (2014) Effects of Mangrove Zonation and the Physicochemical
Parameters of Soil on the Distribution of Macrobenthic Fauna in Kadolkele Mangrove Forest, a
Tropical Mangrove Forest in Sri Lanka. Adv Ecol. https://doi.org/10.1155/2014/564056
Donato DC, Kauffman JB, Murdiyarso D, Kurnianto S, Stidham M, Kanninen M (2011)
Mangroves among the most carbon-rich forests in the tropics. Nat Geosci 4(5):293–297.
Doumett S, Lamperi L, Checchini L, Azzarello E, Mugnai S, Mancuso S, Petruzzelli G, Del Bubba
M (2008) Heavy metal distribution between contaminated soil and Paulownia tomentosa, in a
pilot-scale assisted phytoremediation study: Influence of different complexing agents.
Chemosphere 72:1481–1490.
272 A. Sundaramanickam et al.

Duarte CM, Losada IJ, Hendriks IE, Mazarrasa I Marbà N (2013) The role of coastal plant
communities for climate change mitigation and adaptation. Nat Clim Change 3(11):961–968.
Duke NC (2001) Gap creation and regenerative processes driving diversity and structure of
mangrove ecosystems. Wetl Ecol Manag 9:267–279.
Duke NC (2016) Oil spill impacts on mangroves: recommendations for operational planning and
action based on a global review. Mar Pollut Bull 109(2):700–715.
Duke NC, Burns KA (2003) Fate and effects of oil and dispersed oil on mangrove ecosystems in
Australia. Environmental Implications of Offshore Oil and Gas Development in Australia:
Further Research. A Compilation of Three Scientific Marine Studies. Australian Petroleum
Production and Exploration Association (APPEA), Canberra, pp. 232–363.
Duke NC, Burns KA, Swannell RPJ (1999) Research into the bioremediation of oil spills in tropical
Australia: with particular emphasis on oiled mangrove and saltmarsh habitat. Report to the
Australian Maritime Safety Authority and the Great Barrier Reef Marine Park Authority.
Duke NC, Meynecke JO, Dittmann S, Ellison AM, Anger K, Berger U, Cannicci S, Diele K, Ewel
KC, Field CD, Koedam N, Lee SY, Marchand C, Nordhaus I, Dahdouh-Guebas F (2007) A
world without mangroves. Science 317:41–42.
Duke NC, Schmitt K (2015) Mangroves: Unusual forests at the seas edge. In: Pancel, L., Köhl,
M. (Eds.), Tropical forest handbook. Springer-Verlag, Berlin, Heidelberg.
Dussud C, Ghiglione JF (2014) Bacterial Degradation of Synthetic Plastics- CIESM Workshop
Monograph, France.
Ezcurra P, Ezcurra E, Garcillán PP, Costa MT, Aburto-Oropeza O (2016) Coastal landforms and
accumulation of mangrove peat increase carbon sequestration and storage. Proc Nat Acad Sci
113(16):4404–4409.
Eich A, Mildenberger T, Laforsch C, Weber M (2015) Biofilm and diatom succession on polyeth-
ylene (PE) and biodegradable plastic bags in two marine habitats: early signs of degradation in
the pelagic and benthic zone? PloS one 10 (9):0137201.
Endo S, Takizawa R, Okuda K, Takada H, Chiba K, Kanehiro H, Date T (2005) Concentration of
polychlorinated biphenyls (PCBs) in beached resin pellets: variability among individual
particles and regional differences. Mar Pollut Bull 50(10):1103–1114.
Engler RE (2012) The complex interaction between marine debris and toxic chemicals in the ocean.
Environ Sci Technol 46(22):12302–12315.
Espinosa E, Martinez ME, Torres EF, Rojas MG (2005) Improvement of the hydrocarbon
phytoremediation rate by Cyperus laxus Lam. inoculated with a microbial consortium in a
model system. Chemosph 59:405–413.
FAO (2007) The World’s Mangroves 1980–2005. FAO, Rome.
Fourqurean JW, Duarte CM, Kennedy H, Marba N, Holmer M, Mateo MA, Apostolaki ET,
Kendrick GA, Krause-Jensen D, McGlathery KJ, Serrano O (2012) Seagrass ecosystems as a
globally significant carbon stock. Nat Geosci 5:505–509.
Frias JP, Sobral P, Ferreira AM (2010) Organic pollutants in microplastics from two beaches of the
Portuguese coast. Mar Pollut Bull 60(11):1988–1992.
Fries E, Dekiff JH, Willmeyer J, Nuelle MT, Ebert M, Remy D (2013) Identification of polymer
types and additives in marine microplastic particles using pyrolysis-GC/MS and scanning
electron microscopy. Environm Sci Proc Impacts 15(10):1949–1956.
Gallo F, Fossi C, Weber R, Santillo D, Sousa J, Ingram I, Nadal A, Romano D (2018) Marine litter
plastics and microplastics and their toxic chemicals components: the need for urgent preventive
measures. Environ Sci Eur 30.
Gewert B, Plassmann MM, MacLeod M (2015) Pathways for degradation of plastic polymers
floating in the marine environment. Environ Sci Process Impact 17:1513.
Ghizelini AM, Mendonça-Hagler LCS, Macrae A (2012) Microbial diversity in Brazilian mangrove
sediments e a mini review. Brazilian J Microbiol 1242–1254.
Ghosh SK, Pal S, Ray S (2013) Study of microbes having potentiality for biodegradation of plastics.
Environ Sci Pollut Res 20:4339–4355.
11 Role of Mangroves in Pollution Abatement 273

Harshvardhan K, Jha B (2013) Biodegradation of low-density polyethylene by marine bacteria from


pelagic waters, Arabian Sea, India. Mar Pollut Bull 77(1):100–106.
Horstman EM, Dohmen-Janssen CM, Narra PMF, Van den Berg NJF, Siemerink M, Hulscher SJ
(2014) Wave attenuation in mangroves: A quantitative approach to field observations. Coast
Eng 94:47–62.
Houghton RA (2012) Carbon emissions and the drivers of deforestation and forest degradation in
the tropics. Curr Opin Environ Sustain 4597–603.
Huang XD, El-Alawi Y, Gurska J, Glick BR, Greenberg BM (2005) A multi-process
phytoremediation system for decontamination of persistent total petroleum hydrocarbons
(TPHs) from soils. Microchem J 81:139–147.
Iñiguez ME, Conesa JA, Fullana A (2016) Marine debris occurrence and treatment: A review.
Renew Sustain Energy Rev 64:394–402.
IPCC, Climate Change (2014) Synthesis Report–Summary for Policy makers, Intergovernmental
Panel on Climate Change, Valencia.
Ivar do Sul JA, Costa MF, Silva-Cavalcanti JS, Araújo MCB (2014) Plastic debris retention and
exportation by a mangrove forest patch. Mar Pollut Bull 78(1–2):252–257.
Jambeck JR, Geyer R, Wilcox C, Siegler TR, Perryman M, Andrady A, Narayan R, Law KL (2015)
Plastic waste inputs from land into the ocean. Science 347:768–771.
Jang M, Shim WJ, Han GM, Rani M, Song YK, Hong SH (2016) Styrofoam debris as a source of
hazardous additives for marine organisms. Environ Sci Technol 50(10):4951–4960.
Jeon HJ, Kim MN (2016) Isolation of mesophilic bacterium for biodegradation of polypropylene.
Int Biodeterior Biodegrad 115:244–249.
Joffre OM, Schmitt K (2010) Community livelihood and patterns of natural resources uses in the
shrimp-farm impacted Mekong Delta. Aquacult Res 41:1855–1866.
Kaczorek E, Olszanowski A (2011) Uptake of hydrocarbon by Pseudomonas fluorescens (P1) and
Pseudomonas putida (K1) strains in the presence of surfactants: a cell surface modification.
Water Air Soil Pollut 214:451–459.
Kamath R, Rentz JA, Schnoor JL, Alvare PJJ (2004) Phytoremediation of hydrocarbon
contaminated soils: principles and applications, in: R. Vazquez-Duhalt, R. Quintero-Ramirez
(Eds.), Studies in Surface Science and Catalysis, Elsevier.
Kathiresan K (2003) Polythene and plastic-degrading microbes from the mangrove soil. Rev Biol
Trop 51(3):629–634.
Kathiresan K, Bingham L (2001) Biology of mangroves and mangrove ecosystems. Adv Mar Biol
40:82–251.
Kathiresan K Qasim SZ (2005) Biodiversity of mangrove ecosystems. Hindustan Publishing, New
Delhi(India)
Kathiresan K, Rajendran N (2005) Coastal mangrove forests mitigated tsunami. Estuar Coast Shelf
Sci 65:601–606.
Kehrig HA, Pinto FN, Moreira I, Malm O (2003) Heavy metals and methyl mercury in a tropical
coastal estuary and a mangrove in Brazil. Org Geochem 34:661–669.
Kheiralla ZH, Ashour SM, Rushdy AA, Ahmed HA (2013) Characterization of Biosurfactants
Produced by Halobacillus Dabanensis and Pontibacillus Chungwhensi Isolated from
Oil_Contaminated Mangrove Ecosystem in Egypt. Appl Biochem Microbiol 49(3):263–269.
Kovač Viršeka M, Nika Lovšin M, Koren Š, Kržan A, Peterlin M (2017) Microplastics as a vector
for the transport of the bacterial fish pathogen species Aeromonas salmonicida. Mar Pollut Bull
125(1–2):301–309.
Kristensen E, Bouillon S, Dittmar T, Marchand C (2008) Organic carbon dynamics in mangrove
ecosystems: a review. Aquat Bot 89:201–219.
Kruitwagen G, Pratap HB, Covaci A, Bonga SEW (2008) Status of pollution in mangrove
ecosystems along the coast of Tanzania. Mar Pollut Bull 56:1022–1031.
Kühn S, Bravo EL, Van Franeker JA (2015) Deleterious effects of litter on marine life. In:
Bergmann, M., Gutow, L., Klages, M. (Eds.), Marine Anthropogenic Litter. Springer, Berlin,
pp. 75–116.
274 A. Sundaramanickam et al.

Lacerda LD (2006) Manguezais do nordeste e mudanças ambientais, Ciência Hoje 39:24–28.


Laing GD, Rinklebe J, Vandecasteele B, Meers E, Tack FMG (2009) Trace Metal Behavior in
Estuarine and Riverine Flood Plain Soils and Sediments: A Review. Sci Total Environ 407
(13):3972–3985.
Le H (2008) Economic reforms and mangrove forests in central Vietnam. Soc Nat Resour
21:106–119.
Lebreton L, van der Zwet J, Damsteeg J, Slat B, Andrady A, Reisser J (2017) River plastic
emissions to the world’s oceans. Nat Commun 8:15611.
Lewis III, RR (2005) Ecological engineering for successful management and restoration of man-
grove forests. Ecol Eng 24:403–418.
Lewis M, Pryor R, Wilking L (2011) Fate and Effects of Anthropogenic Chemicals in Mangrove
Ecosystems: A Review. Environ Pollut 159(10): 2328–2346.
Li J, Zhang H, Zhang K, Yang R, Li R, Li Y (2018) Characterization, source, and retention of
microplastic in sandy beaches and mangrove wetlands of the Qinzhou Bay, China. Mar Pollut
Bull 136:401–406.
Li R, Zhang L, Xue B, Wang Y (2019). Abundance and characteristics of microplastics in the
mangrove sediment of the semi-enclosed Maowei Sea of the south China sea: New implications
for location, rhizosphere, and sediment compositions. Environ Pollut 244: 685–692.
Li WC, Tse HF, Fok L (2016) Plastic waste in the marine environment: a review of source,
occurrence and effects. Sci Total Environ 566–567:333–349.
Lima ARA, Barletta M, Costa MF, Ramos JAA, Dantas DV, Melo PAMC, Justio AKS, Ferreira
GVB (2016) Changes in the composition of ichthyoplankton assemblage and plastic debris in
mangrove creeks relative to moon phases. J Fish Biol 89(1):619–640.
Longo C, Savaris M, Zeni M, Brandalise RN, Grisa AMC (2011) Degradation study of polypro-
pylene (PP) and bioriented polypropylene (BOPP) in the environment. Mater Res 14
(4):442–448.
Lovelock CE, Ball MC, Martin KC, Feller IC (2009) Nutrient enrichment increases mortality of
mangroves. PLoS One 4(5):5600.
Lu H, Zhang Y, Liu B, Liu J, Ye J, Yan C (2011) Rhizodegradation gradients of phenanthrene and
pyrene in sediment of mangrove (Kandelia candel (L.) Druce). J Hazard Mat 196:263–269.
Luigi M, Gaetano DM, Vivia B, Angelina LG (2007) Biodegradative potential and characterization
of psychrotolerant polychlorinated biphenyl-degrading marine bacteria isolated from a coastal
station in the Terra Nova Bay (Ross Sea, Antarctica). Mar Pollut Bull 54:1754–1761.
MacFarlane GR, Burchett MD (2002) Toxicity, growth and accumulation relationships of copper,
lead and zinc in the grey mangrove Avicennia marina (Forsk.) Vierh. Mar Environ Res 54
(1):65–84.
MacFarlane GR, Pulkownik A, Burchett MD (2003) Accumulation and Distribution of Heavy
Metals in the Grey Mangrove, Avicennia marina (Forsk.) Vierh: Biological Indication Potential.
Environ Pollut 123(1):139–151.
Machado W, Moscatelli M, Rezende LG, Lacerda LD (2002) Mercury, zinc, and copper accumula-
tion in mangrove sediments surrounding a large landfill in southeast Brazil. Environ Pollut
120:455–461.
Maier RM, Soberon-Chavez G (2000) Pseudomonas aeruginosa rhamnolipids: biosynthesis and
potential applications. Appl Microbiol Biotechnol 54:625–633.
Maiti SK, Chowdhury A (2013) Effects of Anthropogenic Pollution on Mangrove Biodiversity: A
Review. J Environ Prot 4:1428–1434.
Marchand C, Lallier-Verges E, Baltzer F, Alberic P, Cossa D, Baillif P (2006) Heavy metals
distribution in mangrove sediments along the mobile coastline of French Guiana. Mar Chem
98:1–17.
Martin C, Almahasheer H, Duarte CM (2019) Mangrove forests as traps for marine litter. Environ
Pollut 247:499–508.
Massos A, Turner A (2017) Cadmium, lead, and bromine in beached microplastics. Environ Pollut
227:139–145.
11 Role of Mangroves in Pollution Abatement 275

Mazda Y, Magi M, Ikeda Y Kurokawa T, Asano T (2006) Wave reduction in a mangrove forest
dominated by Sonneratia sp. Wetl Ecol Manag 14:365–378.
McIvor AL, Spencer T, Möller I, Spalding M (2012) Storm surge reduction by mangroves. Natural
Coastal Protection Series: Report 2. Cambridge Coastal Research Unit Working Paper 35. ISSN
2050–7941.
Mcleod E, Chmura GL, Bouillon S, Salm R, Björk M, Duarte CM, Lovelock CE, Schlesinger WH,
Silliman BR (2011) A blueprint for blue carbon: toward an improved understanding of the role
of vegetated coastal habitats in sequestering CO2. Front Ecol Environ 9(10):552–560.
McNicoll DM, Baweja AS (1995) Bioremediation of Petroleum-contaminated Soils: An Innova-
tive, Environmental Friendly Technology, Environment Canada.
Mohan AJ, Sekhar VC, Bhaskar T, Nampoothiri KM (2016) Microbial assisted high impact
polystyrene (HIPS) degradation. Bioresour Technol 213:204–207.
Moore C, Phillip C (2011) Plastic Ocean: How a Sea Captain’s Chance Discovery Launched a
Determined Quest to Save the Oceans. First ed. Penguin Group (USA) Inc., New York.
Moreira ITA, Oliveira OMC, Triguis JA, dos Santos AMP, Queiroz AFS, Martins CMS, Silva CS,
Jesus RS (2011) Phytoremediation using Rizophora mangle L. in mangrove sediments
contaminated by persistent total petroleum hydrocarbons (TPH’s). Microchem J 99:376–382.
Mukherjee N, Sutherland WJ, Dicks L, Hugé J, Koedam N, Dahdouh-Guebas F (2014) Ecosystem
service valuations of mangrove ecosystems to inform decision making and future valuation
exercises. PLoS One 9:1–9.
Mukherjee S, Das P, Sen R (2006) Towards commercial production of microbial surfactants. Trends
Biotechnol 24:509–515.
Murphy F, Ewins C, Carbonnier F, Quinn B (2016) Wastewater treatment works (WwTW) as a
source of microplastics in the aquatic environment. Environ Sci Technol 50(11):5800–5808.
Murray BC, Pendleton L, Jenkins WA, Sifleet S (2011) Green Payments for Blue Carbon:
Economic Incentives for Protecting Threatened Coastal Habitats, Nicholas Institute for Envi-
ronmental Policy Solutions, Duke University, North Carolina.
Nagelkerken I, Blaber SJM, Bouillon S, Green P, Haywood M, Kirton LG, Meynecke JO, Pawlik J,
Penrose HM, Sasekumar A, Somerfield PJ (2008) The habitat function of mangroves for
terrestrial and marine fauna: a review. Aquat Bot 89(2):155–185.
Nellemann C, Corcoran E, Duarte CM, Valdés L, De Young C, Fonseca L, Grimsditch G (2009)
Blue Carbon: The Role of Healthy Oceans in Binding Carbon. Blue Carbon. A Rapid Response
Assessment. United Nations Environment Programme, GRID-Arendal.
NOAA (2014) Oil Spills in Mangroves. Planning & Response Considerations. US Department of
Commerce, National Oceanic and Atmospheric Administration (NOAA), National Ocean
Service, Office of Response and Restoration, Seattle, Washington.
Norris BK, Mullarney JC, Bryan KR, Henderson SM (2017) The effect of pneumatophore density
on turbulence: a field study in a Sonneratia-dominated mangrove forest, Vietnam. Cont Shelf
Res 147:114–127.
Paco A, Duarte K, da Costa JP, Santos PSM, Pereira R, Pereira ME, Freitas AC, Duarte AC, Rocha-
Santos, TAP (2017) Biodegradation of polyethylene microplastics by the marine fungus
Zalerion maritimum. Sci Total Environ 586:10–15.
Parrish ZD, Banks MK, Schwab AP (2005) Assessment of contaminant liability during
phytoremediation of polycyclic aromatic hydrocarbon impacted soil. Environ Pollut
137:187–197.
Parrotta JA, Wildburger C, Mansourian S (2012) Understanding Relationships between Biodiver-
sity, Carbon, Forests and People: the Key to Achieving REDD+ Objectives. International Union
of Forest Research Organizations, Vienna.
Paz-Alberto AM, Sigua GC (2013) Phytoremediation: A Green Technology to Remove Environ-
mental Pollutants. Am J Clim Change 2:71–86.
Pendleton L, Donato DC, Murray BC, Crooks S, Jenkins WA, Sifleet S, Craft C, Fourqurean JW,
Kauffman JB, Marbà N, Megonigal P (2012) Estimating global “blue carbon” emissions from
conversion and degradation of vegetated coastal ecosystems. PloS One 7(9):43542.
276 A. Sundaramanickam et al.

Purkhold U, Pommerening-Roser A, Juretschko S, Schmid MC, Koops HP, Wagner M (2000)


Phylogeny of all recognized species of ammonia oxidizers based on comparative 16S rRNA and
amoA sequence analysis: implications for molecular diversity surveys. Appl Environ Microbiol
66:5368–5382.
Purkhold U, Wagner M, Timmermann G, Pommerening-Roser A, Koops HP (2003) 16S rRNA and
amoA-based phylogeny of 12 novel betaproteobacterial ammonia-oxidizing isolates: extension
of the dataset and proposal of a new lineage within the nitrosomonads. Int J Syst Evol Microbiol
53:1485–1494.
Queiroga CL, Nascimento LR, Serra GE (2003) Evaluation of paraffin biodegradation and
biosurfactant production by Bacillus subtilis in the presence of crude oil. Brazilian J Microbiol
34(4):321–324.
Rahman KSM, Rahman TJ, Kourkoutas Y, Petsas I, Marchant R, Banat IM (2003) Enhanced
bioremediation of n-alkane in petroleum sludge using bacterial consortium amended with
rhamnolipid and micronutrients. Bioproc Technol 90(2):159–168.
Rainbow PS, Luoma SN (2011) Metal toxicity, uptake and bioaccumulation in aquatic
invertebrates—modelling zinc in crustaceans. Aqua Toxicol 105(3–4):455–465.
Ramdine G, Fichet D, Louis M, Lemoine S (2012) Polycyclic aromatic hydrocarbons (PAHs) in
surface sediment and oysters (Crassostrea rhizophorae) from mangrove of Guadeloupe: levels,
bioavailability, and effects. Ecotoxicol Environ Saf 79:80–89.
Rech S, Borrell Y, García-Vázquez E (2016) Marine litter as vector of non-native species: what we
need to know. Mar Pollut Bull 113(1–2):40–43.
Ren H, Chen H, Li Z, Han W (2010) Biomass accumulation and carbon storage of four different
aged Sonneratia apetala plantations in Southern China. Plant Soil 327:279–291.
Richards DR, Fries DA (2016) Rates and drivers of mangrove deforestation in Southeast Asia,
2000–2012. PNAS 113 (2):344–349.
Ríos LM, Moore C, Jones PR (2007) Persistent organic pollutants carried by synthetic polymers in
the ocean environment. Mar Pollut Bull 54(8):1230–1237.
Robledo C, Fischler M, Patiño A (2004) Increasing the resilience of hill side communities in
Bolivia: has vulnerability to climate change been reduced as a result of previous sustainable
development cooperation? Mt Res Dev 24:14–18.
Rochman CM, Hoh E, Hentschel BT, Kaye S (2013) Long-term field measurement of sorption of
organic contaminants to five types of plastic pellets: implications for plastic marine debris.
Environ Sci Technol 47(3):1646–1654.
Rodríguez-Rodríguez JA, Sierra-Correa PC, Gómez-Cubillos MC, Villanueva LVL (eds) (2016)
Mangroves of Colombia. In: Finlayson, C., Milton, G., Prentice, R., Davidson, N. The Wetland
Book. Springer, Dordrecht.
Rumbold DG, Evans DW, Niemczyk S, Fink LE, Laine KA, Howard N, Krabbenhoft DP, Zucker
M (2011) Source identification of Florida Bay’s methyl mercury problem: mainland runoff
versus atmospheric deposition and in situ production. Est Coast 34(3):494–513.
Safahieh A, Nabavi MB, Vazirizadeh, A, Ronagh MT, Kamalifar R (2012). Horizontal zonation in
macrofauna community of Bardestan mangrove Creek, Persian Gulf. W J Fish Mar Sci 4(2):
142–149.
Saido K, Itagaki T, Sato H, Kodera Y, Abe O, Ogawa N, Miyashita K (2009) New contamination
derived from marine debris plastics. In: Paper presented at the 238th National Meeting of the
American Chemical Society (ACS).
Salmo III, SG, Duke NC (2010) Establishing mollusk colonization and assemblage patterns in
planted mangrove stands of different ages in Lingayen Gulf, Philippines. Wetl Ecol Manage 18:
745–754.
Sandilyan S, Kathiresan K (2012) Mangrove conservation: a global perspective. Biodivers Conserv
21(14):3523–3542.
Sandilyan S, Kathiresan K (2015) Mangroves as bioshield: An undisputable fact. Ocean Coast
Manage 103:94–96.
11 Role of Mangroves in Pollution Abatement 277

Sarma VV, Hyde KD (2001) A review on frequently occurring fungi in mangroves. Fungal Divers
8:1–34.
Seabra PN (2008) Biorremediação de solos contaminados por petróleo e derivados, in: (Ed.),
Microbiologia Ambiental.
Sekhar VC, Nampoothiria KM, Mohana AJ, Naira NR, Bhaskarb T, Pandey A (2016) Microbial
degradation of high impact polystyrene (HIPS), an e-plastic with decabromodiphenyl oxide and
antimony trioxide. J Hazard Mater 318:347–354.
Sen N, Naskar K (2003) Algal flora of Sundarbans Mangal. Daya, New Delhi.
Sheik, S, Chandrashekar KR, Swaroop K, Somashekarappa HM (2015) Biodegradation of gamma
irradiated low density polyethylene and polypropylene by endophytic fungi. Int Biodeterior
Biodegrad 105:21–29.
Shearer CA, Descals E, Kohlmeyer B, Kohlmeyer J, Marvanova L, Padgett DE, Porter D, Raja HA,
Schmit JP, Thorton HA, Voglymayr H (2007) Fungal biodiversity in aquatic habitats. Biodivers
Conserv 16:49–67.
Siikamäki J, Sanchirico JN, Jardine S, McLaughlin D, Morris DF (2012a) Blue Carbon: Global
options for reducing emissions from the degradation and development of coastal ecosystems,
resources for the future, Washington DC.
Siikamäki J, Sanchirico JN, Jardine SL (2012b) Global economic potential for reducing carbon
dioxide emissions from mangrove loss. Proc Nat Acad Sci 109:14369–14374.
Silva CAR, Lacerda LD, Rezende CE (1990) Metals Reservoir in a Red Mangrove Forest.
Biotropica 4:339–345.
Sowmya HV, Ramalingappa KM, Thippeswamy B (2014) Biodegradation of polyethylene by
Bacillus cereus. Adv Polym Sci Technol 4(2):28–32.
Spalding M, McIvor A, Tonneijck FH, Tol S and van Eijk P (2014) Mangroves for coastal defence.
Guidelines for coastal managers & policy makers. Published by Wetlands International and The
Nature Conservancy. p42.
Spalding M, Kainuma M, Collins L (2010) World Atlas of Mangroves. Earthscan, London.
Sri Dattatreya P, Madhavi K, Satyanarayana B, Adnan Amin, Harini C (2018) Assessment of
Physico-chemical Characteristics of Mangrove Region in the Krishnapatnam Coast, India. Int J
Curr Microbiol App Sci 7(5):2326–2342.
Swan JM, Neff JM, Young PC (1994) Energy Research and Development Corporation (Australia),
Australian Petroleum Production Exploration Association. Environmental Implications of Off-
shore Oil and Gas Development in Australia: the Findings of an Independent Scientific Review,
Sydney.
Tam NFY, Wong YS (2008) Effectiveness of bacterial inoculum and mangrove plants on remedia-
tion of sediment contaminated with polycyclic aromatic hydrocarbons. Mar Pollut Bull
57:716–726.
Tam NFY, Wong YS (1995) Mangrove soils as sinks for wastewater-borne pollutants Hydrobiol
295:231–241.
Teuten EL, Rowland SJ, Galloway TS, Thompson, RC (2007) Potential for plastics to transport
hydrophobic contaminants. Environ Sci Technol 41(22):7759–7764.
Thatoi H, Behera BC, Mishra RR, Dutta SK (2013) Biodiversity and biotechnological potential of
microorganisms from mangrove ecosystems: a review. Annal Microbiol 63(1):1–19.
Thatoi HN, Behera BC, Dangar TK, Mishra RR (2012) Microbial biodiversity in mangrove soils of
Bhitarkanika, Odisha, India.
UNEP (2014) The Importance of Mangroves to People: a Call to Action, UNEP World Conserva-
tion Monitoring Centre, Cambridge.
USEPA (2000) Introduction to Phytoremediation.Washington, DC. EPA/600/R-99/107.
Wang Z, Stout SA (2007) Oil Spill Environmental Forensics. Academic Press, California.
Wang M, Cao W, Jiang C, Yan Y, Guan Q (2018) Potential ecosystem service values of mangrove
forests in southeastern China using high-resolution satellite data. Estuarine Coastal Shelf Sci.,
209: 30-40.
Wang T Zou X Li B, Yao Y, Zang Z, Li Y, Yu W, Wang W (2019a) Preliminary study of the source
apportionment and diversity of microplastics: taking floating microplastics in the South China
Sea as an example. Environ Pollut 245:965–974.
278 A. Sundaramanickam et al.

Wang J, Tan Z, Peng J, Qiu Q, Li M (2016) The behaviors of microplastics in the marine
environment. Mar Environ Res 113:7–17.
Wang Q, Mei D, Chen J, Lin Y, Liu J, Lu H, Yan C (2019b) Sequestration of heavy metal by
glomalin-related soil protein: implication for water quality improvement in mangrove wetlands.
Water Res 148:142–52.
Ward BB, O’Mullan GD (2002) Worldwide distribution of Nitrosococcus oceani, a marine
ammonia-oxidizing gamma-proteobacterium, detected by PCR and sequencing of 16S rRNA
and amoA genes. Appl Environ Microbiol 68:4153–4157.
Wright S, Thompson R, Galloway T (2013) The physical impacts of microplastic on marine
organisms: a review. Environ Pollut 178:483–492.
Wu Y, Chung A, Tam NFY, Pi N, Wong MH (2008) Constructed mangrove wetland as secondary
treatment system for municipal wastewater. Ecol Eng 34:137–146.
Wu RY (1993) Studies on the microbial ecology of the Tansui Estuary. Bot Bull Acad Sin
34:13–30.
Xin, Li, Ryuichiro, K and Kokki, S. 2002. Biodegradation of sugarcane bagasse with marine fungus
phlebia sp. MG–60. J Wood Sci 48: 159–162.
Yakimov MM, Timmis KN, Golyshin PN (2007) Obligate oil-degrading marine bacteria. Curr Opin
Biotechnol 18(3):257–266.
Ye Y, Tam NFY, Wong YS (2001) Livestock wastewater treatment by a mangrove pot- cultivation
system and the effect of salinity on the nutrient removal efficiency. Mar Pollut Bull 42
(6):513–521.
Yoshida S, Hiraga K, Takehana T, Taniguchi I, Yamaji H, Maeda Y, Toyohara K, Miyamoto K,
Kimura Y, Oda K (2016) A bacterium that degrades and assimilates poly (ethylene terephthal-
ate). Science 351(6278):1196–1199.
Zhang K, Su J, Xiong X, Wu X, Wu C, Liu J (2016) Microplastic pollution of lakeshore sediments
from remote lakes in Tibet plateau, China. Environ Pollut 219:450–455.
Zhu J, Zhang Q, Li Y, Tan S, Kang Z, Yu X, Lan W, Cai L, Wang J, Shi H (2019) Microplastic
pollution in the Maowei Sea, a typical mariculture bay of China. Sci Total Environ 658:62–68.
Measurement and Modeling
of Above-Ground Root Systems 12
as Attributes of Flow and Wave Attenuation
Function of Mangroves

Masaya Yoshikai, Takashi Nakamura, Rempei Suwa, Rene Rollon,


and Kazuo Nadaoka

Abstract

Mangrove forests protect the coasts from natural disasters such as storm surges
and tsunamis. These ecosystems also deposit sediments—a key process in carbon
sequestration and adapting to rising sea level. These services provided by
mangroves are related to their drag effect that significantly attenuates flow and
waves, and enhances sedimentation. The drag force exerted by mangroves is due
to the complex structures of their above-ground roots. A key parameter for the
quantitative assessment of drag force is the projected area of vegetation. In this
chapter, we focus on the above-ground root system (prop roots) of Rhizophora—
the most dominant genus in the Asia-Pacific region and likely to exhibit the
highest drag among mangrove species—in exploring the projected area of vege-
tation. We describe the methods of field measurement and an empirical model for
the projected area of the Rhizophora prop root system. The results show the
allometric relationships between the prop root projected area and tree size, and
how the allometric relationships vary depending on the sites. The developed
model shows its great ability as well as some limitations in accurately predicting
the projected area of prop roots. We then discuss the environmental factors that
may affect the allometric relationship, and prospects in developing the universal

M. Yoshikai (*) · T. Nakamura · K. Nadaoka


School of Environment and Society, Tokyo Institute of Technology, Tokyo, Japan
e-mail: [email protected]; [email protected]; [email protected]
R. Suwa
Forestry Division, Japan International Research Center for Agricultural Sciences (JIRCAS),
Tsukuba, Japan
R. Rollon
Institute of Environmental Science and Meteorology, College of Science, University of the
Philippines, Quezon City, Philippines

# The Author(s), under exclusive license to Springer Nature Singapore Pte 279
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_12
280 M. Yoshikai et al.

model, which can predict the prop root projected area in any type of environment.
We also discuss perspectives in the measurement and modeling of other types of
mangrove above-ground root system such as pneumatophores of Avicennia and
Sonneratia species.

Keywords

Drag · Rhizophora · Root system · Allometry · Modeling · Morphological


plasticity

12.1 Introduction

Mangroves are known to form complex structures in their above-ground roots.


Examples are the prop roots of Rhizophora species (Fig. 12.1a), the pneumatophores
of Avicennia and Sonneratia species that are usually denoted as pencil roots

Fig. 12.1 (a) Prop root system of Rhizophora stands and (b) pneumatophores of Avicennia stands
with vertical distribution of the projected area of vegetation (Yoshikai et al., unpublished data). z:
height from the ground, A: projected area of vegetation per unit ground area per 1 cm vertical
interval. The data on A was collected in a planted mangrove forest in Panay Island, the Philippines
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 281

(Fig. 12.1b), and the root knees and buttresses that are seen in Bruguiera species and
some other species (Srikanth et al. 2016; Tomlinson 2016). These are considered to
have evolutionally developed to deal with the anaerobic nature of substrates
generated by the recurrent tidal water flooding by oxygenating the absorbing roots
(fine roots) below-ground through the above-ground roots that are exposed in the
atmosphere (Srikanth et al. 2016; Tomlinson 2016).
The complex structures of the mangrove above-ground root system exert drag
force that substantially influences tidal flow (Furukawa et al. 1997; Mazda et al.
2005; Horstman et al. 2015; Chen et al. 2016). It has been recognized that drag
effects of mangroves have an important role in protecting shorelines against
damages by tsunami-induced waves (Dahdouh-Guebas et al. 2005; Danielsen et al.
2005; Yanagisawa et al. 2010) and storm surges (Das and Vincent 2009; Mclvor
et al. 2012; Menéndez et al. 2018). While it is expected that the risk of coastal
flooding is enhanced by increase in the occurrence of more intense tropical cyclones
and sea-level rise in the future (Woodruff et al. 2013), ecosystem-based coastal
protection has been proposed as sustainable and cost-effective option which could be
alternative to gray structures such as seawalls (Temmerman et al. 2013). In this
regard, the quantification of the effects of mangroves on wave attenuation and
hydrodynamic load reduction is of great interest to properly assess the mangroves’
function as coastal defense (Maza et al. 2017, 2019; Montgomery et al. 2019;
Tomiczek et al. 2020).
The drag effects of mangroves also facilitate sedimentation by flow attenuation
and sediment trapping by the surface of vegetation (Horstman et al. 2015; Willemsen
et al. 2016; Chen et al. 2018). While mangroves have been threatened by sea-level
rise, the vertical accretion of mangroves through sedimentation is considered as a
key process in the survival of mangroves (Krauss et al. 2014; Lovelock et al. 2015;
Woodroffe et al. 2016; Saintilan et al. 2020). Sedimentation also contributes to the
mangrove carbon storage by trapping allochthonous carbon (Xiong et al. 2018), and
thus may play an important role in carbon sequestration.
Several studies highlighted the importance of the quantification of mangrove drag
effects. A key parameter is the projected area of vegetation as suggested in the
following equation for the drag force exerted by vegetation:

1
F D ¼ CD A f U 2 ð12:1Þ
2
where CD is the drag coefficient, Af is the total projected area of the submerged
vegetation per unit volume, and U is the depth-averaged flow velocity (Chen et al.
2016; Maza et al. 2017). An example of the vertical profile of the projected area of
vegetation of Rhizophora and Avicennia stands per unit ground is provided in
Fig. 12.1, which was manually measured in Bakhawan Ecopark in Panay Island,
the Philippines (Yoshikai et al., unpublished data), showing the significance of the
above-ground root system on the projected area of vegetation. This kind of data is
needed for estimating the forest-scale drag force considering the wide areal coverage
of the target mangrove forest. On the other hand, the projected area of vegetation
282 M. Yoshikai et al.

may vary significantly with the mangrove conditions such as tree density, age, and
size of the individual trees. It may also be influenced by species composition as
suggested by the different patterns in the vertical profile of the projected area
between the prop root system and pneumatophores (Fig. 12.1). However, individual
measurements of the above-ground root morphology for the whole mangrove forest
are impractical. In this regard, data collection on the above-ground root morphologi-
cal traits of different species in various sites, and development of a predictive model
based on the collected data are effective ways to estimate the projected area of
vegetation in a mangrove forest, thus moving forward our understanding on the
forest-scale mangrove drag effects.
In this chapter, we describe the methods of field measurement and an empirical
model used to estimate the projected area of the above-ground root system. We
especially focus on the prop root system of Rhizophora, some knowledge and
information on the morphological trait of which have been obtained by recent studies
and the empirical models proposed (Ohira et al. 2013; Yoshikai et al. 2021). This
species is also the most dominant species in the Asia-Pacific region (Ong et al. 2004)
and likely to exhibit the highest drag among mangrove species (Horstman et al.
2014). We also discuss some perspectives on the field measurement and model
development for the other above-ground root types such as pneumatophores of
Avicennia and Sonneratia species, whose morphology have not been well
investigated.

12.2 Field Measurement of Mangrove Above-Ground Root


Morphology

In this section, we review the methods of field measurement of mangrove above-


ground root morphology, including both manual and ground-based remote-sensing
techniques. Although studies on the mangrove above-ground morphology are quite
limited, there are several research works that investigated the Rhizophora prop root
system and the pneumatophores of Avicennia and Sonneratia species. We therefore
targeted these two root systems for the review presented here.

12.2.1 Manual Measurement

12.2.1.1 Prop Root System of Rhizophora


The projected area of Rhizophora stands varies vertically (Fig. 12.1a), and obtaining
and predicting such vertical profile is the objective of the study. In this regard, we
partitioned the prop root system into 0.1 m-thick vertical layers, and obtained the
vertical profile of the projected area in each layer (Ai, where i is the layer number
shown in Fig. 12.2) from field data.
Ohira et al. (2013) proposed a methodology of manual measurement of the prop
root structure in the field to estimate Ai by approximating the individual prop root
shapes as quadratic curves. For the approximation, three parameters are required,
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 283

Fig. 12.2 Vertical layering of the prop root system with layer variables—number of prop roots (Ni)
and projected area (Ai), numbering of individual prop roots, and geometric parameters of prop roots
(HRk: height, Lk: horizontal distance, θk: angle, Φk: diameter of k-th root), and stem (DBH: diameter
at breast height). t-th root is the smallest prop root, thus t indicates the total number of prop roots.
Modified from Yoshikai et al. (2021)

which are root height (HR), horizontal distance (L ), and angle (θ) of individual prop
roots (Fig. 12.2). The shape of each prop root was thus predicted as:

HR þ L tan θ 2
y¼ x þ ð tan θÞx þ HR ð12:2Þ
L2
where y  0 and is the root height from the ground at position x, and x is the
horizontal distance from the emergent point of the prop root. For example, for a
primary (first-order) root that emerges from the stem, x ¼ 0 refers to the position of
the stem. The individual prop root projected area can be obtained by multiplying the
prop root length obtained from Eq. (12.2) and the diameter (Φ). The whole-tree prop
root projected area, which is a sum of the projected areas of individual prop roots,
and its vertical distribution (Ai) can then be calculated. Note that Eq. (12.2) expresses
the projected prop root area from the side view which does not consider the azimuth
angle of prop roots to water flow direction. When the information on the azimuth
angles of the individual prop roots are available, the projected areas of the individual
prop roots from the flow direction can be calculated as:
 2  
HR þ L tan θ x x
y¼ þ ð tan θÞ þ HR ð12:3Þ
L2 cos ψ cos ψ

where ψ is the azimuth angle of the individual prop roots relative to the flow
direction. The methodology proposed by Ohira et al. (2013) quantifies the individual
root shapes, and therefore has the advantage of providing the three-dimensional
284 M. Yoshikai et al.

Fig. 12.3 The fractions of


the projected areas of the first-
to sixth-order prop roots to the
whole-tree prop root projected
area for 22 trees sampled from
a natural mangrove forest in
Ishigaki Island, Japan. The
color indicates the DBHs of
the individual trees. (Data
source: Yoshikai et al. 2021)

(3D) information, which is useful when reconstructing 3D structures from the data.
For example, the model Rhizophora stands used for the laboratory flume
experiments done in Maza et al. (2017, 2019), Shan et al. (2019), and Tomiczek
et al. (2020) were built by referring to the data collected by Ohira et al. (2013).
While the measurement by Ohira et al. (2013) is limited to the first-order roots
only, Yoshikai et al. (2021) have measured the four geometric parameters (HR, L, θ,
Φ; Fig. 12.2) for all prop roots, including the second-, third-, and even higher-order
roots—where the root order indicates the level of branching from the stem—from
one sampled tree. The whole-tree prop root projected area and its vertical distribution
was shown for the first time, and the significance of the prop roots higher than the
first-order in the whole-tree prop root projected area was revealed. Figure 12.3 shows
the fractions of the projected areas of the first- to sixth-order prop roots to the whole-
tree projected area. The data shows that branching rate of prop roots increases
as DBH increases, and up to six-order prop roots were observed in trees with
DBH > 0.13 m. As a result, the presence of prop roots higher than the first-order
becomes more significant in the whole-tree prop root projected area, with up to 80%
significance in a tree with DBH ¼ 0.14 m. This indicates that the non-inclusion of
the prop roots higher than first-order will fall short in obtaining the actual projected
area of the prop root system.
However, measuring the four prop root parameters (HR, L, θ, Φ) for all prop roots
in a tree remains to be very laborious. For example, measurement of the prop root
system of a tree with 30 prop roots by two persons would take 30–40 min, and it is
not uncommon that some trees have more than 100 prop roots depending on tree
size. Therefore, collecting a sufficient number of samples on the prop root system for
developing a predictive model in a site would be labor-intensive and time-
consuming. In this regard, the measurement of the prop root system was simplified
from measuring the four prop root parameters to one parameter—HR in Yoshikai
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 285

et al. (2021) for the purpose in estimating the vertical distribution of Ai without
considering the individual prop root shapes. This simplification is possible using the
strong relationship between the vertical distributions of the number of prop roots (Ni)
and the prop root projected area (Ai), which can be expressed as a linear equation
with intercept fixed at zero (coefficient of determination, R2 ¼ 0.9632):

Ai ¼ 0:0043N i ð12:4Þ

where the unit of Ai is m2 (see Yoshikai et al. 2021 for the plot). This equation
indicates that the averaged projected area of one prop root in vertical layers can be
considered uniform with an area of 0.0043 m2. The vertical profile of Ni can be
obtained from the height data (HR) by counting the number of prop roots in each
vertical layer. With this simplification, the time for measuring the prop root system
morphology can be shortened by around one-third compared to measuring the four
parameters. This was further simplified by measuring only the highest root height
and the total number of prop roots in a tree to predict Ni, and thereby Ai (Yoshikai
et al. 2021); this is described in Sect. 12.3.7. This further simplification greatly
reduces the amount of field work needed to quantify the prop root system of a tree. It
should be noted, however, that this method is valid only for statistically predicting Ni
and Ai, which does not consider the detailed actual 3D structure of the prop root
system.

12.2.1.2 Pneumatophores of Avicennia and Sonneratia Species


There are several studies that have measured the morphology of pneumatophores—
aerial roots of Avicennia and Sonneratia species—in the field. The pneumatophore
morphology such as spatial density (m2), height, and diameter are usually surveyed
within quadrats of 0.5 m  0.5 m or 1 m  1 m with some replicates to characterize
the structural complexity in this type of forest (Dahdouh-Guebas et al. 2007;
Horstman et al. 2014; Lienard et al. 2016; Zhang et al. 2019). This methodology is
used because the pneumatophores usually distribute relatively uniform in space in
the Avicennia- and Sonneratia-dominated forests, unlike the prop roots of
Rhizophora stands.
The individual shape of the pneumatophore is not very complicated. It has been
approximated as a cylindrical shape (Horstman et al. 2014) or truncated cone shape
(Zhang et al. 2015a; Lienard et al. 2016) when quantifying pneumatophore
parameters such as volume and projected area. To validate this approximation, we
have measured pneumatophore geometry (height, diameters at the base, one-third
and two-third heights, and at the tip) in the field, and sampled the pneumatophore for
laboratory analysis. The volume and the lateral surface area of the pneumatophore
were geometrically estimated by assuming a truncated cone shape as in Zhang et al.
(2015a) and Lienard et al. (2016). In the laboratory analysis, the volume of the
sampled pneumatophore was measured using the water displacement method. The
lateral surface area was measured using the aluminum foil method for surface area
determination (Marsh 1970), which is commonly used for measuring the surface
area of corals. The geometrically estimated volume (125 cm3) and lateral surface
286 M. Yoshikai et al.

area (152 cm2) turned out to be very close to the measured values (120 cm3 and
144 cm2, respectively) with around 5% error, suggesting the validity of this approxi-
mation. Figure 12.1b shows the field-measured vertical distribution of the projected
area of vegetation of Avicennia stands in Bakhawan Ecopark, Panay Island, the
Philippines (Yoshikai et al. unpublished data). The near-ground area (around up to
0.3 m height) is densely vegetated with pneumatophores, and shows a high projected
area of vegetation. Because the pneumatophore height is limited up to 0.4 m, the
projected area of vegetation higher than that height is primarily due to the stem parts.
However, the above-mentioned studies are limited to the quadrat-scale only, and
quantification of the whole-tree pneumatophore morphological parameters such as
total pneumatophore projected area, volume, and extent of individual trees remains
unresolved. Therefore, it is still unclear how the projected area per unit ground
shown in Fig. 12.1b will change with tree size and density.
The Avicennia and Sonneratia root system is developed below-ground through
the cable roots. This makes it very difficult to identify which pneumatophore
originates from which tree in the Avicennia- and Sonneratia-dominated forests,
unlike the prop roots system of Rhizophora trees. Investigation of the pneumato-
phore root system requires a below-ground approach. Although some schematic
descriptions on the above- and below-ground root system can be found in Gill and
Tomlinson (1977), Ezcurra et al. (2016), and Tomlinson (2016), few studies have
focused on the below-ground root system. Destructive measurement such as the full
excavation of the root system is one traditional way of quantification (Danjon et al.
2005; Smith et al. 2014), but it is labor-intensive, time-consuming, and poses risks to
the forest ecosystem. While some non-destructive methodologies have been devel-
oped for detecting the below-ground root system in terrestrial ecosystem such as
ground penetrating radar (GPR, Zhu et al. 2014; Hardiman et al. 2017) and electric
resistivity imaging (ERI, Amato et al. 2008; Rossi et al. 2011), there are no studies to
our knowledge that have applied such techniques in mangrove forests. Unlike the
substrate conditions for the terrestrial ecosystem, the substrate of mangrove forests is
under the influence of saline water intrusion that may affect radar attenuation and
electric resistivity below-ground. The effects of saline waters in soil should be
examined for possible application of these non-destructive techniques. All these
factors make the investigation of the pneumatophore root system challenging. Yet,
some studies have added meaningful insights in this field. Yando (2018) measured
the pneumatophore extent of Avicennia germinans at the mangrove-salt marsh
boundary where the extent of pneumatophore of the individual trees can be visibly
recognized, and found a correlation between the maximum pneumatophore extent
and tree height. Vovides et al. (2016) developed a simple non-destructive methodol-
ogy to trace the below-ground cable roots from tree trunk using an ultrasonic
Doppler fetal monitor and steel rods. With this method, they measured the pneumat-
ophore extent of Avicennia germinans stands, and obtained a significant relationship
with tree size as with Yando (2018). These results suggest the scaling relations in the
individual pneumatophore root system and the possibilities for the development of a
predictive model. More comprehensive data such as the number of cable roots per
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 287

tree, branching of the cable roots, and pneumatophore densities specific to the cable
root surface area should be collected for realizing the model development.

12.2.2 Ground-Based Remote-Sensing Techniques

As described in Sect. 12.2.1.1, manual measurement of the detailed morphological


structures of the above-ground root system (e.g., the four parameters HR, L, θ, and Φ
of all the prop roots in a tree) could be labor-intensive and time-consuming. While
the simplification of the measurement of the prop root system could reduce the
amount of field works, it sacrifices some information on the morphological
parameters such as the shape of the individual prop roots and the three-dimensional
spatial extent of the prop root system. On the other hand, applications of novel
ground-based remote-sensing techniques such as terrestrial laser scanning (TLS, also
known as terrestrial LiDAR) and photogrammetry in the characterization of ecosys-
tem structures have recently advanced. These techniques provide the 3D information
on the morphological traits with low time and labor requirements, thus having the
potential to overcome the limitations of manual measurement.
The TLS instrument emits small footprint laser pulses at a high rate and accu-
rately measures the distance between the sensor and a target based on the elapsed
time between the emission and the return of laser pulses (Yao et al. 2011). Hence, the
3D coordinates of the point cloud that represents object surfaces are obtained with
great spatial resolution. The TLS has been used in numerous ecological studies,
including forest biomass estimation (Yao et al. 2011; Hosoi et al. 2013), canopy
structures (Greaves et al. 2015; Olsoy et al. 2016), and more complicated parameters
like leaf angle distribution (Liu et al. 2019). There are some studies that have applied
TLS for mangroves—Olagoke et al. (2016) used TLS for estimating the volume and
biomass of large mangrove trees that are usually difficult to directly measure;
Paynter et al. (2016) used TLS for capturing the 3D structure of a complex prop
root system of Rhizophora mangle and reconstructed a 3D model. In Bakhawan
Ecopark, Panay Island, the Philippines, a field survey using a portable-TLS instru-
ment to scan the Rhizophora apiculata stands was conducted (Fig. 12.4a). Using the
instrument, point cloud of the stands in an area wider than 5 m  5 m was rapidly
and easily acquired (Fig. 12.4b; Baloloy et al. personal communication).
Figure 12.4c shows the scaled up point cloud of a tree, and a 3D model of the
prop root system of this tree was reconstructed using a freeware MeshLab
(Fig. 12.4d). The 3D model well represents the detailed morphological traits of the
complex prop root system that cannot be described by the information taken by the
manual measurement.
Photogrammetry is also often used to quantify the complex structure of mangrove
ecosystems. This method produces a 3D model from a series of overlapping
photographs taken from multiple perspectives using structure-from-motion algo-
rithm (SfM) (Figueira et al. 2015). There are several free and open source software
(e.g., VisualSfM, Regard3D) that allow the easy application of the algorithm for 3D
model reconstruction. This method is also cost-effective as it technically requires
288 M. Yoshikai et al.

Fig. 12.4 (a) Scanning of Rhizophora stands using a portable-TLS instrument in the field, (b)
obtained point cloud (Baloloy et al., personal communication), (c) point cloud upscaled to a tree, (d)
3D model of the prop root system reconstructed using the point cloud, (e) 3D-printed model of the
prop root system, and (f) flume experiment using the 3D-printed models

only a digital camera and a reference scale for data collection in the field. In a large-
scale, this method has been used to map the 3D canopy structures of mangroves from
the photographs taken from UAVs (Navarro et al. 2019; Yaney-Keller et al. 2019).
In a small-scale, Zhang et al. (2015b) has applied photogrammetry to capture the 3D
information of a young Rhizophora stylosa tree with a relatively simple structure.
Lienard et al. (2016) and Norris et al. (2017) applied photogrammetry to the
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 289

pneumatophores of Sonneratia stands in deriving the projected area of vegetation


and volume. Lienard et al. (2016) compared the photogrammetry-derived and
manually measured pneumatophore morphological parameters, and showed a rea-
sonable agreement. They also stressed the ability of photogrammetry in deriving the
morphological parameters of pneumatophores having complex geometry due to the
numerous barnacles attached to the pneumatophore surface, which cannot be
resolved by manual measurement. However, the photogrammetry requires a number
of photographs taken from various perspectives for one target object, thus making it
difficult to collect 3D information of a relatively wide area in a short period of time
compared to TLS (Fig. 12.4b).
Another 3D scanning technique is the use of RGB-D (Red, Green, Blue-Depth)
sensor which is available at low cost (Kamal et al. 2014). Kamal et al. (2014) used
this sensor to scan the 3D structure of a prop root system of R. stylosa and
pneumatophores of Avicennia marina. From the reconstructed 3D models, they
analyzed the fractal dimension in the root system. Yanagisawa and Miyagi (2020)
also used the RGB-D sensor to scan the prop root systems of Rhizophora apiculata
trees with different ages. Their results showed its successful application to both
relatively simple and complex structures of prop root systems.
The studies that have utilized the ground-based remote-sensing techniques
showed remarkable abilities in acquiring the detailed 3D information of the complex
morphological structures of the above-ground root system. These are quite
promising methods for the rapid data collection of the morphological parameters
of above-ground roots with very high accuracy. Also, these have great potential for
quantifying other types of mangrove above-ground root system, like root knees and
buttresses of Bruguiera gymnnorrhiza, the morphological structure of which has not
been investigated yet. Another advantage of these techniques is the possibility of
creating 3D-printed models of the above-ground root system from the data
(Fig. 12.4e). This makes it feasible to conduct flume experiments using more
realistic physical mangrove models for investigating drag effects and hydrodynam-
ics (Fig. 12.4f).

12.3 Modeling the Morphological Structure of the Prop Root


System

In this section, we describe the models for the prop root system developed based on
the field data. The models were designed to predict the vertical distribution of the
projected area Ai using scaling relations. We review two models—a model for a prop
root system with only primary roots developed by Ohira et al. (2013), and a model
which can be applied to prop root system with multiple-order of prop roots devel-
oped by Yoshikai et al. (2021). We then discuss the effects of the environmental
conditions on the complexity of the prop root system.
290 M. Yoshikai et al.

12.3.1 Model for the Prop Root System with Only Primary Roots

Ohira et al. (2013) proposed an empirical model of the prop root system with only
the primary roots. The model predicts the four parameters of the individual prop
roots (HR, L, θ, Φ) from DBH, which is needed to quantify the shapes and the
projected areas of the individual prop roots. The vertical distribution of whole-tree
prop root projected area (Ai; Fig. 12.2) is then obtained. Based on the field data for
the prop root system of Rhizophora mucronata and R. apiculata collected from
Thailand, they found the following relationships: DBH–HRmax, HRmax–t, HRk–Lk,
HRk–θk, and (HRmax, DBH)–Φk relationships, where a uniform vertical interval of
each root height in the tree was assumed. This indicates that the shapes of all the prop
roots of a tree can be predicted only from DBH, thereby Ai. However, it should be
noted that these relationships have been confirmed only for the primary prop roots;
therefore, the model by Ohira et al. (2013) may not be valid for the prop root system
with multiple-order prop roots which can be seen in any mangroves. When looking
at the HRk–Lk relationship for the prop root system in a mangrove forest in Ishigaki
Island, Japan, significant correlation can be seen for the primary roots (Fig. 12.5a),
but the relationship is no longer significant for the higher-order prop roots
(Fig. 12.5b), highlighting the complicated structures of the prop root system with
multiple-order prop roots. This fact limits the applicability of the model by Ohira
et al. (2013) only to the prop root system with primary roots.

a b
1.6 1.6
y = 0.6329x + 0.0462 y = 0.4523x + 0.0638
R² = 0.5036 R² = 0.1057
1.2 1.2
L (m)

0.8 0.8

0.4 0.4

0 0
0 0.4 0.8 0 0.4 0.8

HR (m) HR (m)

Fig. 12.5 Relationship between height (HR) and horizontal distance (L ) for (a) primary roots, and
(b) second-order prop roots of 22 trees sampled from a natural mangrove forest in Ishigaki Island,
Japan. (Data source: Yoshikai et al. 2021)
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 291

12.3.2 Model for the Prop Root System with Multiple-Order


Prop Roots

To characterize the complexity of the prop root system with multiple-order prop
roots and predict Ai, a new model was proposed in Yoshikai et al. (2021). The model
was designed to specifically predict the vertical distribution of the number of prop
roots (Ni)—a good predictor of Ai as suggested by Eq. (12.4). A parameter—scaling
factor (S), which determines the prop root structure—was introduced in the model.
Here, the individual prop roots in a prop root system was numbered in the order of
root height, from the highest to the lowest; the first root refers to the highest root, the
second root refers to the second-highest, and k-th root refers to the k-th highest in the
prop root system (Fig. 12.2). Then the height distribution of the prop roots is
approximated as:

HRkþ1
¼S ð12:5Þ
HRk
where HRk and HRk + 1 are the height of the k-th and (k + 1)-th root (m), and S is the
scaling factor ranging from 0 to 1. Note that the height ratio of the k-th and (k + 1)-th
root could be the ratio of roots coming from different branches. If the height of the
first root (HR1: hereafter denoted as HRmax), which is the maximum root height in a
prop root system, is given, the height of the k-th root can be calculated as:

HRk ¼ HRmax Sðk1Þ ð12:6Þ

The minimum root height is determined by a critical height (HRmin, m). If the t-th
root is the one with the minimum height in a prop root system, t is the largest integer
number that satisfies the following expression:

HRt ¼ HRmax Sðt1Þ  HRmin ð12:7Þ

In this way, the height of prop roots in a tree can be modeled from the parameters
HRmax, S, and HRmin. Ni is then predicted by counting the number of modeled prop
roots in each vertical layer. Using fixed values of HRmax and HRmin, the number of
prop roots increases as the value of S increases toward 1, which implies more
primary (first-order) roots and/or more branching of prop roots (higher-order
roots). Among the three model parameters, HRmax and HRmin can be directly
measured in the field, however, HRmin was treated as a constant with a value of
0.05 m for simplicity. The variations in the values of HRmin are small with a mean
value of 0.062  0.039 m, based on the data collected from various mangroves in
Indonesia, the Philippines, and Japan in Yoshikai et al. (2021), and these variations
do not largely affect the model output. The values of S for the individual trees are
unknown, but can be determined by searching the optimum value (optimized S) with
which the modeled number of prop roots in the vertical layer fits best with the field-
measured number of prop roots. Figure 12.6 shows some examples of the
292 M. Yoshikai et al.

16 18
Site: Bak1 Site: Bak2
vertical layer i
12
DBH: 0.11 m DBH: 0.11 m
HRmax: 1.65 m 12 HRmax: 1.75 m
8 Optimized S: 0.973 Optimized S: 0.921
R2 = 0.984 6 R2 = 0.967
4

0 0
0 50 100 150 0 20 40

9
12 Site: Kii Site: Kar
DBH: 0.06 m DBH: 0.03 m
vertical layer i

8
HRmax: 1.28 m 6 HRmax: 0.85 m
Optimized S: 0.894 Optimized S: 0.984
R2 = 0.957 3 R2 = 0.992
4

0 0
0 10 20 30 40 0 60 120 180
Ni Ni

Fig. 12.6 Comparison of measured and modeled vertical distribution of the number of prop roots
(Ni) for selected trees from different mangroves—Bak1 and Bak2: ~17 and ~ 30 years planted
stands in Bakhawan Ecopark, Panay Island, the Philippines, respectively; Kii: planted stands in KII
Ecopark in Panay Island, the Philippines; Kar: mix of natural and planted stands in Karimunjawa
Island, central Java, Indonesia. The measured values for HRmax, a constant value for HRmin
(0.05 m), and the optimized values for S are used to predict Ni for individual trees. (Data source:
Yoshikai et al. 2021)

comparison of the vertical distribution of the field-measured and modeled number of


prop roots for some selected trees sampled from different mangroves. The number of
prop roots generally increases from the top of the prop root system to the bottom in
the form of a power function. It was shown that the model can reproduce such
patterns with high accuracy by adjusting the value of S. Among the 156 trees
sampled in Yoshikai et al. (2021), the vertical distributions of Ni of 120 trees were
reproduced by the model with R2 values higher than 0.95. This suggests that the
complexity of the prop root system with multiple-order prop roots can be well
represented by the parameter S.

12.3.3 Analysis of the Scaling Relation in the Prop Root System

The optimized S values described in Sect. 12.3.2 have a strong relationship with tree
size such as DBH, which can be expressed as:
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 293

a b
180 10

Whole-tree prop root


projected area (m2)
measured Ni

120 1

60 0.1

y = 1.1504x - 0.0118
R² = 0.8824
0 0.01
0 60 120 180 0.01 0.1

modeled Ni DBH (m)

Fig. 12.7 (a) Comparison of measured and modeled number of prop roots (Ni) in vertical layers for
22 trees sampled from a natural mangrove forest in Ishigaki Island, Japan. All vertical layers with
prop roots were used in the plot. (b) Comparison of measured and modeled whole-tree prop root
projected area for the 22 trees. (Data source: Yoshikai et al. 2021)

S ¼ 1  αDBHα1 ð12:8Þ

where α is a constant, and α1 is the scaling exponent; the unit of DBH is meter. The
value “1” in this equation represents the asymptotic maximum value of S. Therefore,
the S values of the individual trees, which cannot be directly measured in the field,
can be predicted from DBH—the most fundamental and easy-to-measure parameter.
It should be noted that the scaling relations represented by α and α1 may significantly
vary among sites and/or species; this is discussed in the following section.
HRmax of the individual prop root system also have a strong relationship with
DBH as shown by Ohira et al. (2013), Mendez-Alonzo et al. (2015), and Yoshikai
et al. (2021), which can be expressed as:

HRmax ¼ β1 DBH þ β ð12:9Þ

where β1 and β are the slope and intercept terms, respectively. Like the case of
optimized S, the relationship of HRmax with DBH may vary among sites and/or
species.
These suggest that once the scaling relations for S and HRmax for a species in a site
represented by Eqs. (12.8) and (12.9) are obtained, the values of S and HRmax,
thereby Ni, of trees with various DBHs for the species in the site can be predicted
from DBH. An example of the model prediction is provided in Fig. 12.7a which
shows the comparison between measured and modeled Ni and the whole-tree prop
root projected area for 22 trees sampled in site Fuk, where the scaling relations for
S and HRmax were obtained as:
294 M. Yoshikai et al.

S ¼ 1  0.0006DBH1.755 (R2 ¼ 0.92) and


HRmax ¼ 2.71DBH + 0.50 (R2 ¼ 0.43),

respectively (see Yoshikai et al. 2021 for the plots). The model predicted Ni with a
reasonable accuracy (R2 ¼ 0.8824). Using the relationship between Ni and Ai
(Eq. (12.4)), Ai can also be predicted, and the estimated whole-tree prop root
projected area is plotted against DBH together with the field-measured values
(Fig. 12.7b). The whole-tree prop root projected area shows the log-log relationship
with DBH and the model result agrees well with this trend. The log-log relationship
suggests the allometry in the whole-tree prop root projected area, and this is
consistent with the other studies that showed the allometric relations in the prop
root biomass and DBH (Ong et al. 2004; Comley and McGuinness 2005; Van Vinh
et al. 2019).

12.3.4 Site and Species Differences in the Scaling Relation

The compilation of the S–DBH relationships of R. apiculata and R. mucronata


stands obtained from the different sites is provided in Fig. 12.8. Note that the
relationships in the form of the following equation are shown, which is a rearrange-
ment of Eq. (12.8), after log-transformation:

0 Bak1 (R. a)
Bak2 (R. a)
-0.5
Bat (R. a)
log (1 - S)

Bat (R. m)
-1
Mar (R. m)

Kar (R. a)
-1.5
Kar (R. m)

-2 Ber (R. a, R. m)
-2 -1.5 -1 -0.5
log (DBH)

Fig. 12.8 Compilation of S–DBH relationships of R. apiculata and R. mucronata stands sampled
from the different sites—Bak1 and Bak2: ~17 and ~ 30 years planted stands in Bakhawan Ecopark,
Panay Island, the Philippines, respectively; Bat: ~20 years planted stands in an island in Batan Bay
in Panay Island, Philippines; Mar: natural stands in Maratua Island, east Kalimantan, Indonesia;
Kar: mix of natural and planted stands in Karimunjawa Island, central Java, Indonesia; Ber: natural
stands at coast of Berau Continental Shelf in east Kalimantan. R. a: R. apiculata; R. m:
R. mucronata. The unit of DBH is meter. (Data source: Yoshikai et al. 2021)
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 295

log ð1  SÞ ¼ α0 þ α1 log ðDBHÞ ð12:10Þ

where α1 and α0 are the slope and intercept term, respectively, and α0 ¼ log(α) of
Eq. (12.8). Here and hereafter, log denotes the common logarithm. It clearly shows
how the relationships of S–DBH could vary among sites even if the species is the
same. The effects of the differences in the relationship for S can be seen variation of
the total number of prop roots at the different sites. For example, at site Bak2,
R. apiculata trees with DBH ¼ 0.03 ~ 0.04 m (log DBH ¼ 1.52 ~ 1.40) have
around 10 prop roots (log (1S) ¼ 0.65 ~ 0.59) while trees of the same species and
same size at site Bat have around 15 ~ 30 prop roots (log (1S) ¼ 1.07 ~ 0.77),
and 110 ~ 220 prop roots (log (1S) ¼ 1.62 ~ 1.94) in trees of the same species
and size at site Kar. Therefore, the scaling relation obtained at a site should not be
applied to other mangroves without validation; otherwise, it may lead to a large error
in the prediction of Ai, and ultimately the drag force by the mangroves.
The phenomenon causing the site difference in scaling relation is considered as
morphological plasticity of the prop root system, which may be the result of the
mangroves’ adaptation to the severe environment that they inhabit. When the S–
DBH relationships of co-existing species in sites Kar and Bat—Rhizophora
apiculata and Rhizophora mucronata—are compared, the values of S for
R. apiculata tend to be slightly higher than R. mucronata, which indicates higher
complexity in the prop root system of this species. This highlights the importance of
environmental effects on the morphological plastic response and the interspecific
differences in the prop root system that should be taken into consideration when
predicting the prop root structures.

12.3.5 Representation of the Scaling Relations in Prop Root System

To assess the effects of environmental conditions and species variation, a represen-


tation of the site- and species-specific scaling relation is necessary. Equation (12.10),
which represents the site- and species-specific scaling relation, has two parameters—
the slope α1 and the intercept α0. In Yoshikai et al. (2021), it was found that α1 and
α0 have a linear relationship that can be written as:

α1 ¼ 0.4192α0  0.2067 (R2 ¼ 0.82).

Therefore, by substituting this equation into Eq. (12.10), the site- and species-
specific scaling relation can be expressed as:

log ð1  SÞ ¼ α0 þ ð0:4192α0  0:2067Þ log ðDBHÞ ð12:11Þ

In Yoshikai et al. (2021), the value of α0, which minimizes the total error between the
measured and modeled Ni in the vertical layers for all sampled trees, was derived for
each species and each site, and was denoted as optimum α0. By using the optimum
α0 in Eq. (12.11), it was shown that Ni was reproduced with a reasonable accuracy
296 M. Yoshikai et al.

for all sampled trees. This suggests that the site and species differences in the scaling
relations in the prop root system can be well represented by the single parameter α0.
The values of α0 optimized for each species at each site ranged from 5.38 to
2.23 (Yoshikai et al. 2021), where lower α0 value stands for higher complexity of
the prop roots system (higher S values) for trees having the same DBH. It is
considered that species differences and environmental conditions create the
variations in the optimized α0 values. Compared to the range of the optimized
α0 for each species at each site, the differences of the optimized α0 values between
the two co-existing species—R. apiculata and R. mucronata—were small
(0.20 ~ 0.29 in sites Bat and Kar), with lower values for R. apiculata. Therefore,
the effects of environmental conditions on the prop root system are considered to be
more significant than the species variation, which may be the result of the morpho-
logical plastic response in the prop root system.

12.3.6 Effects of Environmental Conditions on Prop Root System


Complexity

Morphological plasticity is a common phenomenon in plants that is considered as


plants’ adaptive strategy to the environment (Bloom et al. 1985). One of the most
typical morphological plastic responses is the different patterns in the biomass
allocation between shoots and roots in response to nutrient availability (Bloom
et al. 1985; Ågren and Franklin 2003; Mašková and Herben 2018). When nutrient
availability decreases, it is likely that plants’ growths are limited by nutrient rather
than carbon availability through photosynthesis, thus plants allocate relatively more
biomass to roots to compensate for the limited resources (Ågren and Franklin 2003).
It has been shown that mangroves also respond to environments in a similar way—
investing more biomass to below-ground parts under nutrient-poor conditions
(Castañeda-Moya et al. 2011, 2013). Sherman et al. (2003) found the increase in
the ratio of root to above-ground biomass with increased soil (pore-water) salinity,
whose osmotic effects significantly regulate the resource uptake of mangroves in
below-ground (Peters et al. 2014); this is consistent with the mangroves’ response to
nutrient-poor conditions (Castañeda-Moya et al. 2011, 2013). Vovides et al. (2014)
showed that mangrove tree stems tend to be less slender with increasing soil salinity,
which may be explained by the plant’s adaptive strategy to enhance the water uptake
ability by increasing the hydraulic conductivity in the stem.
Morphological responses to resource limitation such as those mentioned may be
applied to the prop root system. In Yoshikai et al. (2021), it was implied that factors
that affect the prop root system complexity are sediment thickness, sediment hard-
ness, and species composition. Sediment thickness correlated well with the
optimized α0—trees on a shallow substrate (around 0.1 m) such as mangroves
formed on a reef flat like sites Kar and Mar (Fig. 12.8) tend to have prop root
systems with higher complexity than trees on thicker substrate. Although the
mangrove below-ground root biomass is concentrated in the shallow part of the
sediment (usually until 0.4 m) (Tamooh et al. 2008; Xiong et al. 2017), shallow
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 297

substrates with 0.1 m thickness may constrain the development of the mangrove’s
below-ground root system in the vertical direction, thereby constraining the fine root
biomass development that has a role in absorbing resources below-ground. This may
lead to the limitation of the plant’s growth due to the lowered potential for below-
ground resource uptake. To compensate for the limited development of the below-
ground root system in the vertical direction, plants may have needed to increase
the number of prop roots anchoring to the ground to increase the fine root biomass in
the limited space below-ground. This may be the mechanism for the formation of the
prop root systems with higher complexity on shallow substrates.
Sediment hardness was not directly correlated with the optimized α0 in Yoshikai
et al. (2021), but the multivariate analysis which involved sediment thickness,
sediment hardness, soil salinity, and species variation as independent variables,
and the optimized α0 as dependent variable revealed that the hard sediments such
as sandy substrates possibly contribute to the higher complexity of the prop root
system. Hence, it was hypothesized that the hydrodynamic force by tidal flow or
waves that may have created the sandy substrate may also be the factor affecting the
prop root system complexity. This should be further investigated in future studies to
explain the site variations in the scaling relations of the prop root system.
Soil salinity, on the other hand, had no significant relationship with the optimized
α0 in Yoshikai et al. (2021), which seems contrary to previous studies that showed
changes in the above- and below-ground biomass ratios with changes in soil salinity
(Sherman et al. 2003; Peters et al. 2014). However, the regulation of the below-
ground resources uptake due to high soil salinity may be alleviated by high nutrient
concentrations; plants may need a smaller amount of water uptake if the pore-water
is rich in nutrients. Similarly, even if the soil salinity is not high, limitation of below-
ground resources could be severe if the nutrient concentration is low. The effects of
soil salinity on the prop root system morphology should be therefore considered
together with the pore-water nutrient concentrations. Thus, in future studies, nutrient
concentrations, which were not measured in Yoshikai et al. (2021), should be
measured to explain the effects of soil salinity and below-ground resource availabil-
ity on the complexity of the prop root system.
The results in Yoshikai et al. (2021) implied that species composition may also
affect the prop root system complexity. In a Sonneratia alba-dominated forest at site
Ber, the prop root system of Rhizophora apiculata and Rhizophora mucronata
showed complexity as high as the sites with the shallow substrates, e.g., sites Mar
and Kar, even though the sediment thickness is more than 5 m. Here, the ground is
intensively covered by pneumatophores of Sonneratia alba even inside the area of
the prop root system of Rhizophora stands; thus, below-ground interspecific compe-
tition between Sonneratia alba and Rhizophora species for resources below-ground
such as water and nutrients may be present (Pranchai et al. 2018; Peters et al. 2020).
The below-ground interspecific competition may lead to below-ground resource
limitation for growth of individual trees; this may have facilitated the production
of prop roots anchoring to the ground to increase the fine root biomass and absorb
more water and nutrients. The effects of species composition on the prop root
morphology should also be investigated in future studies.
298 M. Yoshikai et al.

12.3.7 Prospective of the Development of the Universal Model


for Prop Root System

Yoshikai et al. (2021) applied a multivariate model to predict the species- and site-
specific relations from environmental variables and species variation, but the accu-
racy in predicting Ni was not sufficiently high. Therefore, the model has not yet been
generalized for application to various Rhizophora stands with different environmen-
tal conditions and species variation. Further data collection on the scaling relations
of the prop root system with environmental variables from varying environmental
conditions and species composition is needed for development of a universal model.
For rapid data collection of the scaling relations, it was suggested in Yoshikai et al.
(2021) to use the following equation, which is a rearrangement of Eq. (12.7):

log HRmin  log HRmax


log S ¼ ð12:12Þ
t1
where t is the total number of whole-tree prop roots, and HRmin is a constant with a
value of 0.05 m. Thus, only t and HRmax are required, which are relatively easy to
measure in the field, to predict S values of individual trees.
For the environmental variables, it is desirable to measure the pore-water nutrient
concentrations in addition to the sediment thickness and soil salinity, to properly
assess the effects of below-ground resource limitation on the prop root system
morphology. It is also advantageous to conduct sediment coring inside or near the
area of the prop root system to measure the fine root biomass to confirm the
hypothesis that the increased number of prop roots anchoring to the ground increases
fine root biomass to compensate for the below-ground resource limitation. Also, to
assess the effects of hydrodynamic loads on the prop root system morphology,
parameters such as tidal amplitude, ground elevation, flow velocity, and wave height
might be necessary. Prop root system morphology with and without interspecific
competition in similar environments should also be investigated to quantify the
effects of interspecific competition. These kinds of information may help in devel-
oping a universal model for the prop root system.

12.3.8 Perspectives of Model Development of the Other Types


of the Above-Ground Root System

Due to the limited knowledge and data on the above-ground root system of other
mangrove genus such as those with pneumatophores, it is still difficult to develop a
predictive model for these types of above-ground root system. Development of
methodologies to effectively measure the root system in the field is needed. Some
studies have shown an allometric relationship of below-ground biomass with DBH
in mangrove species including Avicennia and Sonneratia species (Comley and
McGuinness 2005; Komiyama et al. 2005). Because the below-ground biomass is
directly linked to pneumatophores (Gill and Tomlinson 1977; Ezcurra et al. 2016;
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 299

Tomlinson 2016), it is expected that the parameters of pneumatophores such as


volume and projected area have also allometric relationships with DBH. The scaling
relations in the pneumatophore root system were also implied by the significant
relationship between the pneumatophore extent and tree size shown by Vovides et al.
(2016) and Yando (2018). These results suggest the possibility of the development
of the pneumatophore root system using the scaling relations.
On the other hand, Vovides et al. (2016) and Yando (2018) showed different
patterns in the relationships between the pneumatophore extent and tree size—the
extent of pneumatophore increased with tree size at first, but saturated at around
1.7 m in Vovides et al. (2016) while the saturation trend was not clearly seen in
Yando (2018), and extension longer than 4 m was observed for some trees. This
difference may be attributed to the different forest conditions—mixed forest of
A. germinans and other species in Vovides et al. (2016) (where pneumatophore
root development may be interfered by the root system of the neighbor trees), and
monospecific stands of A. germinans at the mangrove-salt marsh boundary in Yando
(2018) (where the interference of root system development may not be very severe).
Dahdouh-Guebas et al. (2007) compiled literature data on pneumatophore density of
Avicennia marina from the various mangroves and showed that the density could
significantly vary among sites. They also revealed that depressions of the
microtopography cause local increase in the pneumatophore density and length,
which is probably a mangrove’s adaptation to the places with longer inundation
periods (Dahdouh-Guebas et al. 2007). These studies suggest that the pneumato-
phore root system also exhibits morphological plasticity to the environments. There-
fore, as with the case of the Rhizophora prop root system, it is necessary to collect
the data on the morphology, together with environmental variables from various
mangroves with different environmental conditions and species composition to
develop a universal model. There is a lack of studies quantifying the morphological
structures not only for the pneumatophore root system, but also for other types of
above-ground root system. We recommend that more research be focused on the
mangroves’ morphological trait as these attributes are significant in understanding
the mangrove drag effects that influence coastal protection, adaptation to sea-level
rise, and carbon dynamics of mangrove forests.

Acknowledgments We are grateful to the Japan International Cooperation Agency (JICA) and
Japan Science and Technology Agency (JST) through the Science and Technology Research
Partnership for Sustainable Development Program (SATREPS) for financially supporting the
Project “Comprehensive Assessment and Conservation of Blue Carbon Ecosystems and their
Services in the Coral Triangle (BlueCARES).” We also thank Alvin Baloloy and Dr. Ariel
C. Blanco for provision of the TLS data, Reginald Argamosa for his help in the field survey, and
Dr. Charissa Ferrera for her help in the laboratory analysis.

References
Ågren G, Franklin O (2003) Root : shoot ratios, optimization and nitrogen productivity. Anna Bot
92:795–800.
300 M. Yoshikai et al.

Amato M, Basso B, Celano G, Bitella G, Morelli G, Rossi R (2008) In situ detection of tree root
distribution and biomass by multi-electrode resistivity imaging. Tree Physiol 28:1441–1448.
Bloom AJ, Chapin III FS, Mooney HA (1985) Resource limitation in plants—an economic analogy.
Annu Rev Ecol Syst 16:363–392.
Castañeda-Moya E, Twilley RR, Rivera-Monroy VH, Marx BD, Coronado-Molina C, Ewe SM
(2011) Patterns of root dynamics in mangrove forests along environmental gradients in the
Florida Coastal Everglades, USA. Ecosystems 14:1178–1195.
Castañeda-Moya E, Twilley RR, Rivera-Monroy V (2013) Allocation of biomass and net primary
productivity of mangrove forests along environmental gradients in the Florida Coastal
Everglades, USA. For Ecol Manag 307:226–241.
Chen Y, Li Y, Cai T, Thompson C, Li Y (2016) A comparison of biohydrodynamic interaction
within mangrove and saltmarsh boundaries. Earth Surf Process Landforms 41:1967–1979.
Chen Y, Li Y, Thompson C, Wang X, Cai T, Chang Y (2018) Differential sediment trapping
abilities of mangrove and saltmarsh vegetation in a subtropical estuary. Geomorphology
318:270–282.
Comley BWT, McGuinness KA (2005) Above and below ground biomass, and allometry of four
common northern Australian mangroves. Aust J Bot 53:431–436.
Dahdouh-Guebas F, Jayatissa LP, Di Nitto D, Bosire JO, Lo Seen D, Koedam N (2005) How
effective were mangroves as a defence against the recent tsunami? Curr Biol 15:R443–R447.
Dahdouh-Guebas F, Kairo JG, De Bondt R, Koedam N (2007) Pneumatophore height and density
in relation to micro-topography in the grey mangrove Avicennia marina. Belg J Bot
40 (2):213–221.
Danielsen F, Sorensen MK, Olwig MF, Selvam V, Parish F, et al (2005) The Asian tsunami: a
protective role for coastal vegetation. Science 310:643.
Danjon F, Fourcaud T, Bert D (2005) Root architecture and wind-firmness of mature Pinus
pinaster. New Phytol 168:387–400.
Das S, Vincent JR (2009) Mangroves protected villages and reduced death toll during Indian super
cyclone. Proc Nat Acad Sci 106:7357–7360.
Ezcurra P, Ezcurra E, Garcillán PP, Costa MT, Aburto-Oropeza O (2016) Coastal landforms and
accumulation of mangrove peat increase carbon sequestration and storage. Proc Nat Acad Sci
113:4404–4409.
Figueira W, Ferrari R, Weatherby E, Porter A, Hawes S, Byrne M (2015) Accuracy and precision of
habitat structural complexity metrics derived from underwater photogrammetry. Remote Sens
7:16883–16900.
Furukawa K, Wolanski E, Mueller H (1997) Currents and sediment transport in mangrove forests.
Estuar Coast Shelf Sci 44:301–310.
Gill A, Tomlinson P (1977) Studies on the growth of red mangrove (Rhizophora mangle L.) 4. The
adult root system. Biotropica 9:145–155.
Greaves H, Vierling L, Eitel J, Boelman NT, Magney T, Prager C, Griffin K (2015) Estimating
aboveground biomass and leaf area of low-stature Arctic shrubs with terrestrial LiDAR. Remote
Sens Environ 164:26–35.
Hardiman BS, Gough CM, Butnor JR, Bohrer G, Detto M, Curtis PS (2017) Coupling fine-scale
root and canopy structure using ground-based remote sensing. Remote Sens 9:182.
Horstman EM, Dohmen-Janssen CM, Narra PMF, Van den Berg NJF, Siemerink M, Hulscher
SJMH (2014) Wave attenuation in mangroves: a quantitative approach to field observations.
Coast Eng 94:47–62.
Horstman EM, Dohmen-Janssen CM, Bouma TJ, Hulscher SJMH (2015) Tidal-scale flow routing
and sedimentation in mangrove forests: combining field data and numerical modelling. Geo-
morphology 228:244–262.
Hosoi F, Nakai Y, Omasa K (2013) 3-D voxel-based solid modeling of a broad-leaved tree for
accurate volume estimation using portable scanning lidar. ISPRS J Photogram Remote Sens
82:41–48.
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 301

Kamal S, Lee SY, Warnken J (2014) Investigating three-dimensional mesoscale habitat complexity
and its ecological implications using low-cost RGB-D sensor technology. Methods Ecol Evol
5:845–853.
Komiyama A, Poungparn S, Kato S (2005) Common allometric equations for estimating the tree
weight of mangroves. J Tropic Ecol 21:471–477.
Krauss KW, McKee KL, Lovelock CE, Cahoon DR, Saintilan N, Reef R, Chen L (2014) How
mangrove forests adjust to rising sea level. New Phytol 202:19–34.
Lienard J, Lynn K, Strigul N, Norris BK, Gatziolis D, Mullarney JC, Bryan KR, Henderson SM
(2016) Efficient three-dimensional reconstruction of aquatic vegetation geometry: estimating
morphological parameters influencing hydrodynamic drag. Estuar Coast Shelf Sci 178:77–85.
Liu J, Skidmore A, Wang T, Zhu X, Premier J, Heurich M, Beudert B, Jones S (2019) Variation of
leaf angle distribution quantified by terrestrial LiDAR in natural European beech forest. ISPRS J
Photogram Remote Sens 148:208–220.
Lovelock CE, Cahoon DR, Friess DA, Guntenspergen GR, Krauss KW, Reef R, Rogers K,
Saunders ML, Sidik F, Swales A, Saintilan N, Thuyen LX, Triet T (2015) The vulnerability
of Indo-Pacific mangrove forests to sea-level rise. Nature 526:559–563.
Marsh JA (1970) Primary productivity of reef-building calcareous red algae. Ecology 51
(2):255–263.
Mašková T, Herben T (2018) Root:shoot ratio in developing seedlings: how seedlings change their
allocation in response to seed mass and ambient nutrient supply. Ecol Evol 8:7143–7150.
Maza M, Adler K, Ramos D, Garcia A, Nepf H (2017) Velocity and drag evolution from the leading
edge of a model mangrove forest. J Geophys Res 122(11):9144–9159.
Maza M, Lara JL, Losada IJ (2019) Experimental analysis of wave attenuation and drag forces in a
realistic fringe Rhizophora mangrove forest. Adv Water Resour 131:103376.
Mazda Y, Kobashi D, Okada S (2005) Tidal-scale hydrodynamics within mangrove swamps. Wetl
Ecol Manag 13:647–655.
Mclvor A, Spencer T, Möller I (2012) Storm surge reduction by mangroves. Nat Coast Prot Ser 35.
Mendez-Alonzo R, Moctezuma C, Ordonez VR, Angeles G, Martinez AJ, Lopez-Portillo J (2015)
Root biomechanics in Rhizophora mangle: anatomy, morphology and ecology of mangrove’s
flying buttresses. Ann Bot 115:833–840.
Menéndez P, Losada IJ, Beck MW, Torres-Ortega S, Espejo A, Narayan S, Díaz-Simal P, Lange
G-M (2018) Valuing the protection services of mangroves at national scale: the T Philippines.
Ecosys Serv 34:24–36.
Montgomery JM, Bryan KR, Mullarney JC, Horstman EM (2019) Attenuation of storm surges by
coastal mangroves. Geophys Res Lett 46:2680–2689.
Navarro JA, Algeet-Abarquero N, Fernández-Landa A, Esteban J, Rodríguez-Noriega P, Guillén-
Climent ML (2019) Integration of UAV, Sentinel-1, and Sentinel-2 Data for mangrove planta-
tion aboveground biomass monitoring in Senegal. Remote Sens 11:77.
Norris B, Mullarney J, Bryan K, Henderson SM (2017) The effect of pneumatophore density on
turbulence: a field study in a Sonneratia-dominated mangrove forest, Vietnam. Continent Shelf
Res 147:114–127.
Ohira W, Honda K, Nagai M, Ratanasuwan A (2013) Mangrove stilt root morphology modelling
for estimating hydraulic drag in tsunami inundation simulation. Trees 27:141–148.
Olagoke A, Proisy C, Féret J, Blanchard E, Fromard F, Mehlig U, Menezes MM, Santos V, Berger
U (2016) Extended biomass allometric equations for large mangrove trees from terrestrial
LiDAR data. Trees 30:935–947.
Olsoy PJ, Mitchell J, Levia D, Clark P, Glenn N (2016) Estimation of big sagebrush leaf area index
with terrestrial laser scanning. Ecol Indicat 61:815–821.
Ong JE, Gong WK, Wong CH (2004) Allometry and partitioning of the mangrove, Rhizophora
apiculata. For Ecol Manag 188(1–3):395–408.
Paynter I, Saenz E, Genest D, Peri F, Erb A, Li Z, Wiggin K, Muir J, Raumonen P, Schaaf ES,
Strahler A, Schaaf C (2016) Observing ecosystems with lightweight, rapid-scanning terrestrial
lidar scanners. Remote Sens Ecol Conserv 2:174–189.
302 M. Yoshikai et al.

Peters R, Vovides AG, Luna S, Grueters U, Berger U (2014) Changes in allometric relations of
mangrove trees due to resource availability—a new mechanistic modelling approach. Ecol
Modell 283:53–61.
Peters R, Walther M, Lovelock C, Jiang J, Berger U (2020) The interplay between vegetation and
water in mangroves: new perspectives for mangrove stand modelling and ecological research.
Wet Ecol Manag 28:697–712.
Pranchai A, Jenke M, Vogt J, Grueters U, Yue L, de Menezes MM, Wagner S, Berger U (2018)
Density-dependent shift from facilitation to competition in a dwarf Avicennia germinans forest.
Wetlands Ecol Manage 26:139.
Rossi R, Amato M, Bitella G, Bochicchio R, Gomes J, Lovelli S, Martorella E, Favale P (2011)
Electrical resistivity tomography as a non-destructive method for mapping root biomass in an
orchard. Eur J Soil Sci 62:206–215.
Saintilan N, Khan NS, Ashe E, Kelleway JJ, Rogers K, Woodroffe CD, Horton BP (2020)
Thresholds of mangrove survival under rapid sea level rise. Science 368:1118–1121.
Shan Y, Liu C, Nepf H (2019) Comparison of drag and velocity in model mangrove forests with
random and in-line tree distributions. J Hydrol 568:735–746.
Sherman RE, Fahey T, Martinez P (2003) Spatial patterns of biomass and aboveground net primary
productivity in a mangrove ecosystem in the Dominican Republic. Ecosystems 6:384–398.
Smith A, Astrup R, Raumonen P, Liski J, Krooks A, Kaasalainen S, Åkerblom M, Kaasalainen M
(2014) Tree root system characterization and volume estimation by terrestrial laser scanning and
quantitative structure modeling. Forests 5:3274–3294.
Srikanth S, Kaihekulani S, Lum Y, Chen Z (2016) Mangrove root: adaptations and ecological
importance. Trees 30:451–465.
Tamooh F, Huxham M, Karachi M, Mencuccini M, Kairo JG, Kirui B (2008) Belowground root
yield and distribution in natural and replanted mangrove forests at Gazi bay Kenya. For Ecol
Manag 256(6):1290–1297.
Temmerman S, Meire P, Bouma TJ, Herman P, Ysebaert T, De Vriend HJ (2013) Ecosystem-based
coastal defence in the face of global change. Nature 504:79.
Tomiczek T, Wargula A, Lomonaco P, Goodwin S, Cox D, Kennedy A, Lynett P (2020) Physical
model investigation of mid-scale mangrove effects on flow hydrodynamics and pressures and
loads in the built environment. Coast Eng 162:103791.
Tomlinson PB (2016) The botany of mangroves. 2nd ed. Cambridge University Press, Cambridge.
Van Vinh T, Marchand C, Linh TVK, Vinh DD, Allenbach M (2019) Allometric models to estimate
above-ground biomass and carbon stocks in Rhizophora apiculata tropical managed mangrove
forests (Southern Viet Nam). For Ecol Manag 434:131–141.
Vovides AG, Vogt J, Kollert A, Berger U, Grueters U, Peters R, Lara-Domínguez AL, López-
Portillo J (2014) Morphological plasticity in mangrove trees: salinity-related changes in the
allometry of Avicennia germinans. Trees 28:1413–1425.
Vovides AG, Marín-Castro B, Barradas G, Berger U, Lopez-Portillo J (2016) A simple and cost-
effective method for cable root detection and extension measurement in estuary wetland forests.
Estuar Coast Shelf Sci 183:117–122.
Willemsen PWJM, Horstman EM, Borsje BW, Friess DA, Dohmen-Janssen CM (2016) Sensitivity
of the sediment trapping capacity of an estuarine mangrove forest. Geomorphology
273:189–201.
Woodroffe CD, Rogers K, McKee KL, Lovelock CE, Mendelssohn IA, Saintilan N (2016)
Mangrove sedimentation and response to relative sea-level rise. Ann Rev Mar Sci 8:243–266.
Woodruff JD, Irish JL, Camargo SJ (2013) Coastal flooding by tropical cyclones and sea-level rise.
Nature 504:44.
Xiong Y, Liu X, Guan W, Liao B, Chen Y, Li M, Zhong C (2017) Fine root functional group based
estimates of fine root production and turnover rate in natural mangrove forests. Plant Soil
413:83–95.
12 Measurement and Modeling of Above-Ground Root Systems as Attributes of. . . 303

Xiong YM, Liao BW, Wang FM (2018) Mangrove vegetation enhances soil carbon storage
primarily through in situ inputs rather than increasing allochthonous sediments. Mar Pollut
Bull 131:378–385.
Yanagisawa H, Miyagi T (2020) A study on the structural characteristics of prop root of mangrove
trees using the 3-D laser scanner. Mangrove Sci 11:23–26. (in Japanese)
Yanagisawa H, Koshimura S, Miyagi T, Imamura F (2010) Tsunami damage reduction perfor-
mance of a mangrove forest in Banda Aceh, Indonesia inferred from field data and a numerical
model. J Geophys Res 115:C06032.
Yando ES (2018) Dispersal, establishment, and influence of black mangrove (Avicennia germinans)
at the salt marsh-mangrove ecotone (Doctoral dissertation, University of Louisiana at
Lafayette). Retrieved from ProQuest Dissertations and Theses database. (10814132)
Yaney-Keller A, Tomillo PS, Marshall J, Paladino F (2019) Using Unmanned Aerial Systems
(UAS) to assay mangrove estuaries on the Pacific coast of Costa Rica. PLoS ONE 14(6):
e0217310.
Yao T, Yang X, Zhao F, Wang Z, Zhang Q, Jupp D, Lovell J, Culvenor D, Newnham G,
Ni-Meister W, Schaaf C, Woodcock C, Wang J, Li X, Strahler A (2011) Measuring forest
structure and biomass in New England forest stands using Echidna ground-based lidar. Remote
Sens Environ 115:2965–2974.
Yoshikai M, Nakamura T, Suwa R, Argamosa R, Okamoto T, Rollon R, Basina R, Primavera-
Tirol Y, Blanco AC, Adi NS, Nadaoka K (2021) Scaling relations and substrate conditions
controlling the complexity of Rhizophora prop root system. Estuar Coast Shelf Sci 248:107014.
Zhang X, Chua VP, Cheong H-F (2015a) Geometrical and material properties of Sonneratia alba
mangrove roots. Trees 29:285–297.
Zhang X, Chua VP, Cheong H (2015b) Hydrodynamics in mangrove prop roots and their physical
properties. J Hydro-environ Res 9:281–294.
Zhang Y, Ding Y, Wang W, Li Y, Wang M (2019) Distribution of fish among Avicennia and
Sonneratia microhabitats in a tropical mangrove ecosystem in South China. Ecosphere 10(6):
e02759.
Zhu S, Huang C, Su Y, Sato M (2014) 3D Ground penetrating radar to detect tree roots and estimate
root biomass in the field. Remote Sens 6:5754–5773.
Mangrove as a Natural Barrier
to Environmental Risks and Coastal 13
Protection

Nazlin Asari, Mohd Nazip Suratman, Nurul Atiqah Mohd Ayob, and
Nur Hasmiza Abdul Hamid

Abstract

Mangroves has been widely acknowledged for its role as a natural barrier to
various environmental risks such as storms, tsunamis, waves, and coastal erosions
by becoming the first defense in protecting the coastlines. In addition, this natural
barrier also provides a wide range of ecosystem services including fisheries,
timber productions, provision of foods, tourism, climate regulation, carbon stor-
age that reduce the coastal communities’ vulnerability to hazards. Unfortunately,
in many parts of the world, mangroves have been lost due to the pressure of urban
developments, rapid expansion of aquaculture and agriculture, mining, overex-
ploitation of timbers, and an increasing environmental risk. Given the extensive
damages caused by the impact of environmental risks, for instance, the recent
tsunami occurred in 2018 at Sulawesi, mangroves have never been more impor-
tant to be conserved and restored due their crucial roles in providing natural
protection to the ecosystems. Therefore, mangroves need to be seen as valuable
resources to be managed and sustainably. This chapter aimed at reviewing the
roles of mangrove as coastal protection as well as a natural barrier to storm surges,
tsunami, wind, waves, and erosion, and the benefits of protecting mangroves.

N. Asari · N. H. Abdul Hamid


Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM), Shah Alam, Malaysia
Institute for Biodiversity and Sustainable Development, Universiti Teknologi MARA (UiTM),
Shah Alam, Malaysia
M. N. Suratman (*)
Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]
N. A. Mohd Ayob
Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM), Shah Alam, Malaysia

# The Author(s), under exclusive license to Springer Nature Singapore Pte 305
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_13
306 N. Asari et al.

Keywords

Mangrove forest · Coastal protection · Environmental risks · Natural barrier ·


Mangrove ecosystem

13.1 Introduction

Mangrove forests are a unique ecosystem that exist between the land and the sea at
low latitudes in the tropical or subtropical regions of the planet and able to withstand
tidal differences. It has been widely known to be the first defenses against the
environmental risks such as storms, tsunamis, waves, and coastal erosions that can
save lives and property. Despite its coastal protection importance, this ecosystem
also provides a variety of services to local community including as sources of
income, food, fisheries, home to fish breeding, timber provision, tourism, climate
regulation, and carbon storage.
However, it was estimated that around one quarter of the world’s mangroves have
been lost due to anthropogenic activities, mainly through conversion to aquaculture,
agriculture and urban land uses (Barbier and Cox 2003; Duke et al. 2007; Spalding
et al. 2010; Friess and Webb 2014; Barbier 2016). Asia has the largest mangrove
extent which is about 42% from 1% of all tropical forests in the world (Menéndez
et al. 2018); however, these mangrove forests continue to loss at an alarming rate of
2% a year (Spalding et al. 2010).
As the mangrove worldwide continues to disappear or degraded due to environ-
mental risks, human population, and development pressures, it becomes essential for
better understanding the roles of mangroves as a natural barrier to environmental
risks and to assess the benefit of having mangrove forests as our natural coastal and
livelihood protection. These existing protective values offered by mangrove forests
will be sacrificed and coastal risk will be increased if these habitats continue to loss
or degraded. Therefore, the fact that these mangrove forests are not just providing the
protective values to the coastal, but also have economic importance to humans. The
consideration for protection, restoration, and conservation of these habitats may be
important in future coastal management decision. Mangrove forests need to be seen
as a valuable resource to be managed and conserved on a sustainable basis. This
chapter aimed at reviewing the roles of mangroves as coastal protection as well as a
natural barrier to storm surges, tsunami, wind, waves, and erosion. Apart from that,
this paper will also highlight the benefits of protecting mangroves.

13.2 An Overview of Mangrove Forests

Mangrove forests occur along sheltered coastlines in the tropics and sub-tropics. The
first analysis of global mangrove area was conducted by the Food and Agriculture
Organization (FAO) in 1980 with 14,653,000 ha (FAO 2007). However, the recent
analysis performed by Hamilton and Casey (2016) indicated that the total mangrove
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 307

areas globally are estimated to be 8,349,500 ha. The estimated figure by Hamilton
and Casey (2016) was more accurate and reliable as compared to the compilation of
statistical reports because they incorporated remotely-sensed technology. Remote
sensing techniques are known to be essential for mangroves as it involves an
extensive area which is difficult to access and larger topographic maps are not
available (Suratman 2014). Various advancements in remote sensing technology
for mangrove ecosystems have been discussed by Suratman (2014). The change
detection analysis conducted by Hamilton and Casey (2016) has clearly proven that
the mangrove area worldwide has significantly reduced about 57% in the last four
decades from 1980.
Malaysia, which is one of the countries in South East Asia, has among the largest
extents of mangroves (Hamdan 2012) despite consisting of only less than 2% of the
total land area in Malaysia. According to a report by Hamdan et al. (2018) on the
characterizing and monitoring of mangroves in Malaysia using Landsat, in 2017,
there are a total of 627,567 ha of mangrove forest reserve in Malaysia, where mostly
found in Sabah (60%), followed by Sarawak (22%) and Peninsular Malaysia (18%).
This report showed the changes that occurred in mangrove areas over 27 years of
monitoring from 1990 in which the rate of mangroves deforestation was about 0.1%
per year between 1990 and 2017. Most of the changes occurred mainly outside the
Permanent Forest Reserve and according to the states’ structural planning. The major
factors that contributed to these changes have been identified as direct conversion to
other land uses, predominantly for aquaculture and agriculture, and coastal erosion
(Hamdan et al. 2018).
Mangrove forests are able to withstand harsh environment. They have dense and
massive root systems in which some of grow above the ground mainly to allow these
plants to absorb nutrients and water, to breathe and to take root in the muddy soil that
characterizes these regions. The roots and some parts of trunks of mangrove trees are
covered by water during high tide and will rise above the water level during low tide.
Some species of mangrove trees are resistant to saltwater or brackish water on tidal
flats. The specific characteristics of the mangrove ecosystem make it a good habitat
for a variety of terrestrial, aquatic, and amphibian animals (Rasmeemasmuang and
Sasaki 2015). According to Spalding et al. (2010), such forests are fertile and create a
balanced ecosystem. There are a total of 67 species of mangrove out of 27 genera
from 16 families (Field 1995). Among the 67 species, 29 are from the families of
Acanthaceae and Rhizophoraceae.

13.3 Mangroves as a Natural Barrier and Coastal Protection

In the past decades, various events occurred on the coastline that caused a huge
impact to the coastal communities, properties, and lives, such as the Indian Ocean
Tsunami in 2004 with a magnitude of 9.0 MW, which caused severe damages to
Sumatra, Nicobar and Andaman Islands, Bangladesh, Myanmar, Thailand,
Malaysia, and Singapore. Hurricane Katrina occurred in the United States in 2005,
Typhoon Haiyan had caused significant destruction in the Philippines and Southeast
308 N. Asari et al.

Asia in 2013, and the recent tsunami that occurred at Sulawesi Indonesia in 2018.
While there are growing concern on the possibility of more frequent and powerful
tropical cyclones due to climate change (Webster 2005; Knutson et al. 2010; Marois
and Mitsch 2015), the restoration, conservation, and creation of mangrove forests
can provide the solution in reducing these environmental risks damages to the areas
that are vulnerable to this natural disaster. Mangrove forests are highly tolerant in
harsh environment, being daily subject to salt exposure, temperature, water, and
varying degree of anoxia. Therefore, these forests exhibit a high degree of ecological
stability as compared to other forests (Alongi 2008). There is growing evidence that
mangrove served as a natural barrier to tsunami, wind, wave, storm, and coastal
erosion, which will be discussed further in the next sections.

13.3.1 Mangroves as a Tsunami Barrier

According to Ahmadun et al. (2020), tsunami refers to as the series of waves with an
extremely long wavelength that grow in height and velocity as they enter the port or
near coastal area where the depth is relatively shallow. It usually occurs due to
earthquakes, asteroid impact, or volcanic activities in the deep sea, which caused the
sudden movements of the seabed (Egorov 2007; Ahmadun et al. 2020). The evi-
dence from the Indian Ocean Tsunami 2004 indicated that mangrove can help in
reducing tsunami damage, but it is depending on the critical factors such as wave
depth (Cunningham 2019), width of forest, proportion of aboveground biomass
vested to roots, tree height, soil texture, size and speed of the tsunami, distance
from tectonic event, and angle of the tsunami incursion relative to the coastline
(Alongi 2008). According to Cunningham’s findings from Wood Environment &
Infrastructure Solutions UK Limited, it is estimated that a 500 meter wide mangrove
forest can significantly dissipate the force of tsunamis under 3 meters, but beyond
this the mangroves are likely to become damaged. This was supported by the finding
from modeling results of Yanagisawa et al. (2010) indicated that mangrove forests
could potentially reduce the hydrodynamic force of tsunami by 70% when incoming
waves remained under 3 meters.
Based on few initial post-impact studies in Southeastern India, Andaman Islands,
Sri Lanka, and Malaysia indicated that mangroves offered a positive and significant
defense against the full impact of the tsunami (Kathiresan and Rajendran 2005;
Chang et al. 2006; Alongi 2008; Ahmadun et al. 2020). Cunningham (2019) stated
that number of studies using hydraulic models and wave experiments that have been
validated using the data collected in the aftermath of the Indian Ocean Tsunami 2004
and it has been widely reported (FAO n.d.) that the extensiveness of mangrove
forests can help in reducing loss of life and property damage from tsunami by
absorbing the first brunt of impact and dissipating wave energy as it approaches
coastline. For instance, in the Penang State of Malaysia, large inland territories
where the mangrove forests were intact, only slightly damaged as reported by
Penang Inshore Fisherman Welfare Association. Similar examples for the positive
role of mangrove to help in reducing the damages caused by tsunami and act as
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 309

buffer zone during past natural hazards such as in Tamil Nadu, India, the well-
established mangrove forest of Pichavaram acted as a protective belt slowing down
the waves and protected around 1700 people living in hamlets build inland between
100 to 1000 m from the mangroves (FAO n.d.).
It is clear that the effectiveness of mangrove forests against tsunami depends on a
few critical factors such as depth of forest (100–500 m), wave depth, and speed.
However, the presence of mangrove forests could still help in protecting the
properties and people even though it does not offer full protection against tsunami
over certain wave depth.

13.3.2 Reduction of Wind and Swell Wave Damage

Wind and swell wave are formed by the action of the wind on the water surface in
areas of open water. The wind waves are generated near the coast, while swell waves
are generated away from the coast that have been agitated by the effects of wind
(Pugh 1987; Woodroffe 2002; McIvor et al. 2012a). Wind and swell waves are
commonly seen at the seashore, as they break on the beach or smash against rocks
(Spalding et al. 2014). Number of reports indicated that the mangrove forests have
the ability to reduce wave height and buffer wind speed (Mazda et al. 2006; Quartel
et al. 2007; Bao 2011). While mangrove forests are receiving little incoming waves
every day, it may receive larger waves during storms and periods of high winds
which can cause flooding and damage to coastal infrastructure. The value of these
mangrove forests in providing such protection for high speed and damaging winds
and wave are often overlooked (Barbier 2016).
McIvor et al. (2012b) reviewed available information about the capacity of
mangroves to reduce wind and swell waves, in order to inform decision makers,
planners, and coastal engineers about the role mangroves can play in coastal defense
against these hazards. According to their study, by reducing wave energy and height,
mangroves can potentially reduce the height of wind and swell waves over relatively
short distances in which wave height can be reduced by between 13 and 66% over
100 m of mangroves. The highest rate of wave height reduction per unit distance
occurs near the mangrove edge, as the waves begin to pass through the mangroves.
There are a few critical factors that are affecting the wave attenuation in
mangroves such as the depth of water in the mangrove forest, slope, topography,
the size and age of trees, the height of trees, the types of mangrove in the area,
density of mangrove trees as shown in Fig. 13.1 (McIvor et al. 2012b). Hence, the
amount of protection that is offered by the mangrove forest depends greatly on the
forest quality.

13.3.3 Protection against Storm Damage

Storm surges refer to as the abnormal rise of sea water level in coastal areas during a
short-lived atmospheric disturbance such as storm or hurricane, measured as the
310 N. Asari et al.

density and spacing of trees

height of
age and size
leaves
of trees

mangrove
species

wave
height
distance travelled aerial root
aphy water depth
Slope, topogr through mangrove structure and
ry (≈ tidal phase)
and bathymet height

Fig. 13.1 Factors affecting the wave attenuation in mangroves. (Adapted from McIavor et al.
2012b)

Surge Level:
With Mangroves
Surge Level:
Without Mangroves

Mangroves
increase
surge level
Mangroves in front
decrease
surge level
at back

Fig. 13.2 General schematic of mangrove effects on surge level. (Source: Narayan et al. 2019)

height of the water over and above the normal predicted astronomical tide (NOAA
2020). It is a phenomenon mainly linked to wind, but also depending on other
elements such as atmospheric pressure, swells, and tides as well as bathymetry and
topography (Bertin 2016). Storm surge reduction rate can be measured as a certain
number of centimeters of water level reduction per meter of inland distance (McIvor
et al. 2012a).
Mangrove forests have been shown as particularly important ecosystem-based
protection (first line defend) against storm surge (Krauss et al. 2009; Zhang et al.
2012; McIvor et al. 2015; Narayan et al. 2019). By acting as obstacles to storm
waves and surges, they reduce flood extents and therefore protecting people and
reduce property damages on coastlines (WAVES 2017; Menéndez et al. 2018;
Dasgupta et al. 2019). Figure 13.2 shows the general schematic of mangrove effects
on surge level. The reduction level of storm surge on the adjacent coastal to
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 311

mangrove resulting from the mangrove “blocking” (increasing surge level in sea-
ward and decreasing level on the back of mangrove) (Narayan et al. 2019).
The complex network of tree roots, trunks, and branches of the mangroves forest
obstructs the flow of water and can serve to trap debris movement, even large
moving objects which can reduce the damage to the adjacent coastal lands. Where
mangroves are extensive, they are able to reduce storm surge water depths as the
surge flows inland. Although the measured rates of storm surge reduction through
mangrove range from 5–50 cm km 1 width of mangroves, a small reduction in water
level can already greatly reduce the extent of flooding in adjacent coastal land
(Spalding et al. 2014). In addition, surface wind waves are expected to be reduced
by more than 75% over one kilometer of mangroves (McIvor et al. 2012b). The
extent of mangrove protection depends on a number of factors, including mangrove
width, density of mangrove forest, topography, storm characteristic such as the size
and forward speed of the storm and spectral features of waves and tidal stage at
which waves enter the forest (McIvor et al. 2012b; Dasgupta et al. 2019).
A study in Florida found that the surge height decreases at the rate of 23 cm km 1
through an area with a mixture of mangrove and open water, while in areas with less
open water recorded surge height reduction rates ranged from 40 to 48 km. In
contrast, the storm surges height at the front of the mangrove zone increases by
about 10% to 30% because of the “blockage” of mangroves to surge water (Zhang
et al. 2012). Studies have shown that Hurricane Wilma that hit South Western
Florida in 2005 had flooded the area and extended to 70% further inland without
the protection of the mangroves (Zhang et al. 2012; Liu et al. 2013).
Storm surge and associated flooding often lead to loss of human life, and
destruction of property and infrastructure in populated, low coastal areas. Between
2005 and 2015, WAVES (2017) recorded 56% of property damage from natural
hazards in the Philippines was due to typhoons and storms, and another 29% due to
floods. Mangroves not only provide the most protection for frequent lower intensity
storms, but for more catastrophic events, such as the 1-in-25-year storm, they
provide more than US $1.6 billion in averting damages throughout the Philippines
(WAVES 2017). The flood impact on Tanya Delta, Kenya experienced by villages
unprotected by mangroves was 2.4 times greater than that of villages partially
protected by mangroves and 14.7 times greater than that of villages completely
protected by mangroves (Karanja and Saito 2018). Das and Vincent (2009)
conducted a study following the cyclone with 9 m storm surge in 1999 in Orissa
India. They found that villages with wider mangroves between them and the coast
had significantly fewer deaths than villages with narrow mangroves or no mangroves
at all.

13.3.4 Reduction of Coastal Erosion

Erosion refers to as the land ward displacement of the shoreline resulting in the loss
of land and retreating shoreline. Erosion may be caused by storms, wave, tides, and
winds, or in response to large-scale events such as glaciation or orogenic cycles that
312 N. Asari et al.

may significantly alter sea levels and tectonic activities that cause coastal land
subsidence or emergence (Prasteya 2007). Wind, waves, and currents are natural
forces that easily move the unconsolidated soils in the coastal area, resulting in rapid
changes in the position of the shoreline. When the waves and current tap against the
shore, it creates change where it can sometimes bring the sediment to the coast but
sometimes causing erosion and the loss of land (Vo Quoc and Kuenzer 2012). Vo
Quoc and Kuenzer (2012) study found that there is a clear occurrence of erosion-
deposition processes but the erosion process is more dominant. Over the past
decades, the erosion is the most important issue in the coastal areas all over the
world which has threatened coastal community. Nearly 30% of the Malaysian
coastline is undergoing erosion (Othman 1994). In Vietnam, the coastline has been
eroded continuously at the rate of approximately 50 m per years since the early
twentieth century (Mazda et al. 1997). The erosion can cause saline intrusion,
affecting drinking water sources and agricultural production. They also can lead to
massive flooding during storm surge, high tides, or periods of excessive rainfall
(Winterwerp et al. 2014).
According to Danielsen et al. (2005), the establishment of mangroves serves as an
important ecological function in enhancing the sedimentation and mitigating coastal
erosion by reducing the energy of waves, tidal currents and storms that would
otherwise erode the coastline. A number of factors affect the ability of mangrove
forests to reduce erosion, which included forest width, degree of sediment compac-
tion, and tree morphology (height, root structure, ratio of above- to below-ground
biomass) (Alongi 2008). As discussed earlier, mangroves able to dissipate wind and
ease the wave energy, thus slow the flow of current which reducing the water
capacity to dislodge sediment and carry them out of the mangrove area. Slower
water flow allowed already suspended sediment to settle out from the water, hence
encourage sediment deposition (Spalding et al. 2014). Mangroves can trap more than
80% of incoming sediment brought by a wave due to the calmness of water
(Furukawa and Wolanski 1996) and contributed at 1–8 mm sediment per year
(Horstman et al. 2014).
Many species of mangrove trees have different types of aerial root such as stilt
root, pneumatophores, root knees, and plank roots. The dense network of fine root
systems helps consolidate the coastal soil (Mazda et al. 1997), binding the sediments
in places (Saenger 2002; Alongi 2008), and preventing the sediment from being
washed away by waves. Observation along the west coast of peninsular Malaysia
shows mangrove has an important role to play in the sediment deposition and erosion
cycle of the muddy coast (Othman 1994). Rhizophora and Bruguiera species are
found to form a denser mat of roots compared to Avicennia species, therefore more
consolidate soil and deep root system to reduce the erosion (Othman 1994). A
similar study by Azlan and Othman (2009) stated species such as Rhizophora
apiculata (Bakau Minyak), Rhizophora mucronata (Bakau Kurap), and Sonneratia
(Perepat) found along the shoreline are able to serve as a natural buffer against the
destruction of wave action and/or tide and Rhizophora sp. is one of the best
attenuators among other species. Mangrove soils are rich in organic matter produced
by the mangroves themselves, including living root, but also dead leaves and woody
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 313

materials. The organic matter accumulates to form a peat that increase in thickness
over time due to mangrove soils are often waterlogged and have very low oxygen
content due to anaerobic condition (Spalding et al. 2014).
Thampanya’s et al. (2006) study showed mangrove-covered coastlines are said to
be less likely to erode or will erode more slowly than non-vegetated coastlines or
former loss of mangroves contributed to the dramatically change in the sediment
movement pattern where the land begins to erode so that land disappears into the sea.
Conversion of mangrove area to agriculture or aquaculture purposes leads to the
rapid breakdown of organic matter as oxygen became available in the soil which in
turn causes subsidence. The conversion effect has been seen in coastal areas, such as
along the coast in Central Java (Winterwerp et al. 2014), Gulf of Thailand, and
Guyana where the coastal erosion and accretion occur irregularly along the coast, but
an intensification of erosion has been noticed every year (Spalding et al. 2014).

13.4 Importance of Mangrove Forests

13.4.1 Act as Carbon Stores

Mangrove forests are present in 105 countries, including Indonesia, Brazil,


Malaysia, and Papua New Guinea. It contains 50% of the global mangrove intensity
and grouped as among the most carbon rich ecosystems in the tropics. Mangroves
are important tropical carbon sinks, and their role in mitigating climate change is
well documented across the globe. Different from tropical forests, mangrove forests
are able to store more carbon per area as compared to other terrestrial ecosystem
(Shaltout et al. 2020) which makes them an important natural carbon sink. One of the
studies on mangrove carbon storage in North Sulawesi, Central Kalimantan, and
Central Java, Indonesia conducted by Murdiyarso et al. (2009) indicated that that
total carbon storage in mangrove ecosystems is exceptionally high compared with
most forest types, with a mean of 968 Mg C ha 1 and a range of 863–1073 Mg C
ha 1. These carbon stocks were a result from a combination of well-structured forest
with tree up to 2 m in diameter and organic rich peat soil to a depth of 5 m or more.
On other studies by Trumper et al. (2009) and Mcleod et al. (2011), it has been
demonstrated that these coastal forests can store up to five times the carbon present in
tropical forests per area, sequestering over 2 tons of carbon on average per ha/year.
Moreover, the quantification of carbon stocks is necessary to evaluate their contri-
bution to climate regulation, as it is one of the most important ecosystem services of
mangroves (Palacios Peñaranda et al. 2019).
Several studies have reported that most of the carbon stocks in the mangrove
forests are stored in below-ground biomass and sediment carbon reserves of over
80% from the total carbon stocks (Romañach et al. 2018; Palacios Peñaranda et al.
2019; Eid et al. 2020). Among the main sources of organic carbon for soil are litter
production and dead wood debris from the plant. The decomposition and decay of
organic material caused by bacteria increase the accumulation of organic matter in
the soil sediment. The amount of carbon storage in mangrove ecosystems varies
314 N. Asari et al.

geographically, which is determined by the surroundings (Eid et al. 2020) such as


structural forest properties (basal area and height), and water quality parameters such
as salinity, and dissolved oxygen (Palacios Peñaranda et al. 2019).
It is also important to highlight that mangrove forests can be carbon sinks, but, if
the ecosystem is disturbed, they can become a carbon source as well. The clearing of
mangrove forests causes the drying up of mangrove sediments, which increase the
microbial activity following the loss of anaerobic environment. This in turn causes
an oxidation for the soil and leads to the release of stored carbon into the atmosphere
which then contributes to the remission of atmospheric greenhouse gas emissions. A
study conducted by Shaltout et al. (2020) found that carbon loss from deforestation
of mangrove represents 0.6% of its global emission annually. It is also found that
mangrove conversion could release to the atmosphere by 84 to 159 million tonnes of
CO2 with the values up to 90% of the carbon stored in the soils are lost after
deforestation. For instance, by converting mangroves to pasture would release
three times more CO2 per hectare to the atmosphere than the conversion of Amazon
forests. Therefore, higher rates of mangrove loss in recent decades put their carbon
stocks at risk and reduce its ability to act as carbon stores. The impacts are obviously
on the loss of the carbon sink function as well as the 94% of the potential carbon
sequestration by mangrove forest worldwide (Friess 2016). Thus, it is essential to
acknowledge the importance of the mangrove forests and to value their conservation.
As mangrove forests can store sizeable volume of carbon, they need to be preserved
and managed sustainably, to retain along with the increase in carbon storage.

13.4.2 Rural Livelihoods and Mangroves

Ecologically, mangrove ecosystem serves as a defense of ecological balance


between life on the land and the sea. They form an important aspect of the
livelihoods of coastal communities in developing countries. Mangrove forests
have contributed significantly to the socio-economic lives of coastal resident and
benefit various human populations from local and regional scales (Orchard et al.
2015; Jakovac et al. 2020). With the existing mangrove forest, its benefit returns to
local people itself as one of main sources of income to the communities. It also
expected to increase the added value for the people who live around the mangrove
area, including fish farmers and fishermen (Saw and Kanzaki 2015). For example,
the high price of crab in the world market has helped the locals in gaining extra
income as reported by Saw and Kanzaki (2015) in the province of South Sulawesi.
With the development of fish pond around the mangrove forest area, it increases the
social income and at the same time creates jobs for locals.
Given that the mangrove forests influence marine fish production, this mangrove
area produces a wide range of import and export commodities with greater number
of fodders such as shrimp, crabs, shells, and fish (Jakovac et al. 2020). For instance,
in 2013–2014, there were 3443 thousand tonnes of marine fish production at 36% of
total fish production in the India. These marine fish species include cuttlefish, squid,
lobster, shrimp, and certain types of finfish. As reported by Jakovac et al. (2020),
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 315

marine fishes are mangrove-dependent and referring to as crustaceans such as


prawns and crabs, mollusks, demersal finfish, snappers, catfishes, pomfrets, and
croakers. According to Friess (2016), 1 km2 increase in mangrove areas leads to a
185.84 tonnes increase in total marine fish production yearly. The annual per hectare
contribution of mangroves to total marine fish production is therefore 1.86 tonnes.
Although mangrove forests become an energy source for many species of marine
biota, they also act as shelter and habitat for some livelihoods (Umilia and Asbar
2016). An eco-farming aquaculture system in Guangxi, for example, allows the
implementation of fishery systems without cutting mangrove trees. The system is
based on a network of underground tubes and pipes in between mangrove roots that
augments availability of habitat for fishes. The system generated between 27,000
and 45,000 US$ per hectare per year in fish production (Dat and Yoshino 2013).
Besides habitat for fishes, mangrove forest also increases the yield of shrimp
productivity as reported by Spalding and Parrett (2019). In reality, as in Dat and
Yoshino (2013) study, shrimp productivity was great with the existence of mangrove
forests as compared to without any mangrove forest (Jakovac et al. 2020). Hence, it
is obvious that mangrove forest has provided sources of income and food to local
communities as well as home for marine species breeding.
Apart from sources of income and food, mangrove ecosystem also serves as a
supplier of products that bring economic benefits to humans. It supplies fuel
(charcoal), construction materials, and also medicine (Anneboina and Kumar
2017). In addition to their contribution to human used, mangroves also provide
raw materials such as timber and wood (Friess 2016).

13.4.3 Biodiversity of Mangroves

Mangrove forests are typically composed of shrubs and trees and form extensive
forested wetlands along both muddy and carbonate coasts in tropical and subtropical
climates (Eid et al. 2020). Mangrove has contributed to the present diversity in the
mangrove ecosystems as old as 65 million years. According to Duncan et al. (2016),
with only 0.12% of the world’s total land area and 0.7% of the total global area of
tropical forest, mangrove forest gives an effect on the biodiversity. Rich in sediments
and nutrients in mangrove forest provides better nursery grounds and breeding sites
for birds, reptiles, and mammals in this environment (Duncan et al. 2016). There are
over 9000 ha of mangrove forests in Peninsular Malaysia which act as largest
estuarine mangrove system. They provide special habitat for animals and also for
marine life including seagrasses, algae, fungi, and fishes. The salinity changes with
the tides and season influences the distribution of organism in this ecosystem
(Pasquaud et al. 2015).
As a coastal intertidal wetland forest, mangroves composed of halophytic tree and
shrub species. According to Pasquaud et al. (2015), there are approximately 70 veg-
etation species in 40 genera. Associated with this vegetation are a plethora of coastal
and terrestrial fauna, including fish, crustaceans, snakes, and mammals. These flora
and fauna diversity in mangrove forests provide scientific study and also offer
316 N. Asari et al.

tourism opportunities such as in Malaysia, there is a well-known habitat for


twinkling fireflies (Photuris lucicrescens) that lives on mangrove trees along the
river bank located at Kampung Kuantan (Suratman 2008). Table 13.1 shows the list
of flora species found in mangrove forests in Malaysia.
At higher latitudes, mangroves intergrade into temperate intertidal habitats such
as salt marsh ecology. In this ecology, it is composed of fine silts and clays, mud flats
harbor burrowing creatures, including clams, mussels, oysters, fiddler crabs, sand
shrimp, and bloodworms (Adi and Sari 2016). As reported by Nusantara et al. (2015)
with the presence of mangrove, the shrimp productivity is at greatest point. It tends
to improve the production of shrimp in the mangrove forest compared to other area.
The increase density of fauna includes tiger prawn (Penaeus monodon), sea crab
(Scylla paramamosain), and brackish fishes including Latescal carifer and
Oreochromis niloticus as well as seagrasses (Gracilaria spp.).
Apart from that, mangroves also provide shelter for many types of birds
(Romañach et al. 2018). They are acting as critical habitat for native and migratory
bird species. It includes a wide variety of waterbirds and terrestrial birds. There are
150–250 bird species in each bio-geographical as reported by Hattam et al. (2021).
The most common species includes waders, herons, egrets, storks, and birds or prey
such as sea eagles, brahminy kites, ospreys, and fish eagles (Hattam et al. 2021).

13.5 Conclusions

The main objective of this chapter was to review and summarize the role of
mangrove in coastal protection as well as a natural barrier to storm surges, tsunami,
wind, waves, and erosion which continue to be debated globally. Apart from that,
this paper also reviews the benefits of protecting mangroves. Evidences from few
studies have shown that the mangrove forests provide the solution in reducing the
environmental risks of damages to the areas that are vulnerable to this natural
disaster. Mangrove forests act as a natural barrier to the tsunamis which help in
reducing the damages impact of the tsunami. The presence of mangrove forests
could still help in protecting the properties and people even though it does not offer
full protection against tsunami over certain wave depth. Mangrove forests can
potentially reduce the height of wind and swell waves over relatively short distances;
however, the amount of protection depends greatly on the forest structure, depth, and
composition. Mangrove forests not only provide protection from frequent lower
intensity storms, but from more catastrophic events, which depends on mangrove
width, density of mangrove forest, topography, size, and forward speed of the storm.
Apart from that, mangrove forests also offer the protection against coastal erosion by
reducing the wave energy and slow the flow of current which reducing the water
capacity to dislodge sediment out of mangrove areas. Therefore, the conservation of
healthy mangrove forests can be viewed as an adaptive measure that is important in
reducing the risk as a form of sustainable coastal protection. It is important to note
that forest quality influences the amount of protection that is offered by the man-
grove forests. Hence, the mangrove forest with good structure, composition, depth,
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 317

Table 13.1 A list of plant species found in the mangrove forests in Malaysia (Saenger et al. 1983;
Ashton and Macintosh 2002; MOSTI 2003 as cited by Suratman 2008)
Life- Common name
No. Family Species Categorya form (in Malay)
1 Acanthaceae Acanthus ilicifolius MA Shrub Jeruju puteh
2 Arecaceae Nypa fruticans MA Palm Nipah
3 Asteraceae Pluchea indica MA Shrub Beluntas
4 Avicenniaceae Avicennia alba M Tree Api-api puteh
5 Avicenniaceae A. lanata M Tree Api-api bulu
6 Avicenniaceae A. marina M Tree Api-api jambu
7 Avicenniaceae A. officinalis M Tree Api-api ludat
8 Combretaceae Lumnitzera littorea M Shrub/ Teruntum merah
tree
9 Combretaceae L. racemosa M Shrub/ Teruntum putih
tree
10 Euphorbiaceae Excoecaria M Tree Buta-buta
agallocha
11 Meliaceae Xylocarpus M Tree Nyireh bunga
granatum
12 Meliaceae X. meluccensis M Tree Nyireh batu
13 Myrsinaceae Aegiceras M Shrub Kachang-kachang
corniculatum
14 Myrsinaceae A. floridum M Shrub Kachang-kachang
15 Pteridaceae Acrostichum aureum M Fern Piai raya
16 Pteridaceae A. speciosum M Fern Piai lasa
17 Rhizophoraceae Bruguiera cylindrica M Tree Berus
18 Rhizophoraceae B. gymnorhiza M Tree Tumu merah
19 Rhizophoraceae B. parviflora M Tree Lenggadai
20 Rhizophoraceae B. sexangula M Tree Tumu putih
21 Rhizophoraceae Ceriops decandra M Tree Tengar
22 Rhizophoraceae C. tagal M Tree Tengar
23 Rhizophoraceae Rhizophora M Tree Bakau minyak
apiculata
24 Rhizophoraceae R. mucronata M Tree Bakau kurap
25 Rubiaceae Scyphiphora M Shrub Chigam
hydrophyllacea
26 Sapotaceae Planchonella MA Tree Menasi
obovata
27 Sonneratiaceae Sonneratia alba M Tree Perepat
28 Sonneratiaceae S. caseolaris M Tree Berembang
29 Sonneratiaceae S. ovata M Tree Gedabu
30 Sterculiaceae Heritiera littoralis MA Tree Dungun
31 Leguminosae Caesalpinia crista MA Tree Unak
32 Leguminosae Derris trifoliata MA Tree Tuba laut
33 Leguminosae D. uliginosa MA Tree Setui
34 Malvaceae Thespesia populnea MA Tree Bebaru
(continued)
318 N. Asari et al.

Table 13.1 (continued)


Life- Common name
No. Family Species Categorya form (in Malay)
35 Pandanaceae Pandanus MA Palm Pandan
odoratissimus
36 Tiliaceae Brownlowia MA Shrub/ Kiei
argentata tree
MA: mangrove associate
a
M: true mangrove

width, and high density could potentially become the protective layers in any
catastrophic events associated with environmental risks damages as compared to
small and scattered mangrove forests. In assessing the overall value of these
ecosystems, it is important not to focus just on their protective value, but at the
exclusion of the wide range of benefits and synergistic relationships that these vital
habitats provide which supports thousands of marine life species breeding, rich in
biodiversity and sources of livelihood for coastal communities. This habitat is also
important in carbon storing as compared to the other tropical forests in which
contribute to the reduction of CO2 and their conservation can be look upon as
investing in climate mitigation. Improving the value of the protective service of
mangroves, as well as the other benefits provided by these critical habitats, may
prove important in the future coastal management decisions.
Given the enormous benefits of mangrove forests, proper management and
conservation is therefore necessary to ensure the continued existence of mangrove
forests. The Conservation of mangroves can be enhanced by few initiatives such as
gazetting all remaining mangrove forests within forest reserves or protected areas.
Some mangrove forests are already gazetted but many other mangroves have yet to
be protected. Other measures are by retaining protective mangrove buffers along
coastlines and rivers to prevent erosion. Managing mangrove forests as fishery
reserves to encourage the environmentally sensitive commercial aquaculture
activities could help in conservation of mangrove. Public awareness and educating
the community by enlightenment campaigns should be conducted to sensitize the
local population on the effect of unsustainable removal of the mangrove vegetation
as well as on the sustainable management and utilization of the forest and forest
resources. The loss of mangrove forest will not only contribute to the rapid loss of
biodiversity and shoreline, but also negatively impact human livelihoods and eco-
system function. Mangroves conservation should not be overlooked, especially as
they are important for speciation and can be significant drivers of diversification
over time.

References
Adi W, Sari SP (2016). Detection of mangrove distribution in Pongok Island. Procedia Environ Sci
33: 253–257.
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 319

Ahmadun F, Wong RMM, Said MA (2020) Consequences of the 2004 Indian Ocean Tsunami in
Malaysia. Saf Sci 121: 619–631.
Alongi DM (2008) Mangrove forests: resilience, protection from tsunamis, and responses to global
climate change. Estuar Coast Shelf Sci 76: 1–13.
Anneboina LR, Kumar KV (2017) Economic analysis of mangrove and marine fishery linkages in
India. Ecosyst Serv 24: 114–123.
Azlan NI, Othman R (2009) Monitoring of mangroves area using remote sensing towards shoreline
protection. Proceedings of 16th International Symposium GIS, Ostrava, 85–95.
Bao TQ (2011) Effect of mangrove forest structures on wave attenuation in coastal Vietnam.
Oceanologia 53: 807–818.
Barbier EB (2016) The protective service of mangrove ecosystems: a review of valuation methods
analysis. Mar Pollut Bullet 109 (2): 676–681.
Barbier EB, Cox M (2003) Does economic development lead to mangrove loss: a cross country.
Contemp Econ Policy 21: 418–432.
Bertin X (2016) Storm surges and coastal flooding: status and challenges. La Houille Blanche 2:
64–70.
Chang SE, Adams BJ, Alder J, Berke PR, Chuenpagdee R, Ghosh S, Wabnitz C (2006) Coastal
ecosystems and tsunami protection after the December 2004 Indian Ocean tsunami. Earthquake
Spectra 22: 863–887.
Cunningham E (2019) The role of mangrove in coastal protection. Published by Wood Environ-
ment & Infrastructure Solutions UK Limited.
Danielsen F, Sorensen MK, Olwig MF, Selvam V, Parish F, Burgess ND, Hiralshi T, Karunagaran
VM, Rasmussen MS, Hansen LB, Quarto A, Suryadiputra N (2005) The Asian tsunami: a
protective role for coastal vegetation. Science 310: 643.
Das S, Vincent JR (2009) Mangroves protected villages and reduced death toll during Indian super
cyclone. Proc Natl Acad Sci USA 106: 7357–7360.
Dasgupta S, Islam MS, Huq M, Khan ZH, Hasib MR (2019) Quantifying the protective capacity of
mangroves from storm surges in coastal Bangladesh. PLoS One 14: e0214079.
Dat PT, Yoshino K (2013) Comparing mangrove forest management in Hai Phong City, Vietnam
towards sustainable aquaculture. Procedia Environ Sci 17: 109–118.
Duke NC, Meynecke JO, Dittmann S, Ellison AM, Anger K, Berger U, Cannicci S, Diele K, Ewel,
KC, Field CD, Koedam N, Lee SY, Marchand C, Nordhaus I, Dahdouh-Guebas F (2007) A
world without mangroves? Science 317: 41–42.
Duncan C, Jurgenne HP, Nathalie P, Julian RT, Rona J, Loma A, Heather JK (2016) Rehabilitating
mangrove ecosystem services: a case study on the relative benefits of abandoned pond reversion
from Panay Island, Philippines. Mar Pollut Bullet 109: 772–782.
Egorov Y (2007) Tsunami wave generation by the eruption of underwater volcano. Nat Hazards
Earth Syst Sci 7 (1): 65–69.
Eid ME, Khaled MK, Hamdi A, Muhammad A, Adel M, Abir Mc (2020) Evaluation of carbon
stock in the sediment of two mangrove species, Avicennia marina and Rhizophora mucronata,
growing in the Farasan Islands, Saudi Arabia. Oceanologia 62: 200–213.
Field C (1995) Journey amongst mangroves. International Society for Mangrove Ecosystem,
Okinawa, Japan.
Food and Agriculture Organization (FAO) (2007) The world’s mangroves 1980–2005, FAO
Forestry Paper 153: 77.
Food and Agriculture Organization (FAO) (n.d.) Tsunami mitigation by mangroves and coastal
forests. http://www.fao.org/forestry/tsunami/27285@69434/en/
Friess DA (2016) Mangrove forests. Magazine, National University of Singapore. Curr Biol 26:
748.
Friess DA, Webb EL (2014) Variability in mangrove change estimates and implications dominated
by Sonneratia sp. Wet Ecol Manag 14 (4): 365–378.
Furukawa K, Wolanski E (1996) Sedimentation in mangrove forest. Mangroves Salt Marshes 1:
3–10.
320 N. Asari et al.

Hamdan O, Misman MA, Linggok V (2018) Characterizing and monitoring of mangroves in


Malaysia using Landsat-based spatial-spectral variability. IOP Conf Ser: Earth Environ Sci
169: 10.
Hamdan O (2012) Status of Mangroves in Peninsular Malaysia. Hamdan, O (Ed.), Forest Research
Institute Malaysia, Gemilang Press Sdn Bhd.
Hamilton SE, Casey D (2016) Creation of a high spatio-temporal resolution global database of
continuous mangrove forest cover for the 21st century (CGMFC-21). Global Ecology and
Biogeography 25(6) 729–738. doi:10.1111/geb.12449.
Hattam C, Goh HC, Then AY, Edwards-Jones A, Ruslan NF, Yap JS, Moh HH (2021) Using nexus
thinking to identify opportunities for mangrove management in the Klang Islands, Malaysia.
Estuar Coast Shelf Sci 248: 107157.
Horstman EM, Dohmen-Janssen CM, Narra PMF, van den Berg NJF, Siemerink M, Hulscher
SJMH (2014) Wave attenuation in mangroves: a quantitative approach to field observations.
Coast Eng 94: 47–62.
Jakovac CC, Latawiec AE, Lacerda E, Lucas IL, Korys K , Iribarrem A, Malaguti GA, Turner RK,
Luisetti T, Strassburg BBN (2020) Costs and carbon benefits of mangrove conservation and
restoration: A global analysis. Ecol Econ 176 (2020) p 106758 10.1016/j.ecolecon.2020.106758
Karanja JM, Saito O (2018) Cost–benefit analysis of mangrove ecosystems in flood risk reduction: a
case study of the Tana Delta, Kenya. Sustain Sci 13: 503–516.
Kathiresan K, Rajendran N (2005) Coastal mangrove forests mitigated tsunami. Estuar Coast Shelf
Sci 65: 601–606.
Knutson TR, McBride JL, Chan J, Emanuel K, Holland G, Landsea C, Held I, Kossin JP, Srivastava
AK, Sugi M (2010) Tropical cyclones and climate change. Nat Geosci 3: 157–163.
Krauss KW, Doyle TW, Doyle TJ, Swarzenski CM, From AS, Day RH, Conner WH (2009) Water
level observations in mangrove swamps during two hurricanes in Florida. Wetlands 29 (1):
142–149.
Liu H, Zhang K, Li Y, Xie L (2013) Numerical study of the sensitivity of mangroves in reducing
storm surge and flooding to hurricane characteristics in southern Florida. Cont Shelf Res 64:
51–65.
Marois DE, Mitsch W (2015) Coastal Protection from tsunamis and cyclone provided by mangrove
wetlands-A review. Int J Biodivers Sci Ecosyst Serv Manag 11 (1): 71–83.
Mazda Y, Magi M, Ikeda Y, Kurokawa T, Asano T (2006) Wave reduction in a mangrove forest for
the assessment of ecosystem service provision. Glob Ecol Biogeogr 23: 715–725.
Mazda Y, Magi M, Kogo M, Hong PN (1997) Mangroves as a coastal protection from waves in the
Tong King Delta, Vietnam. Mang. Salt Marsh 1(2): 127–135.
McIvor A, Spencer T, Möller I, Spalding M (2012a) Storm surge reduction by mangroves. Natural
Coastal Protection Series: Report 2. Cambridge Coastal Research Unit Working Paper 41. The
Nature Conservancy and Wetlands International.
McIvor AL, Möller I, Spencer T, Spalding M (2012b) Reduction of wind and swell waves by
mangroves. Natural Coastal Protection Series: Report 1. Cambridge Coastal Research Unit
Working Paper 40. Published by The Nature Conservancy and Wetlands International. p 27.
McIvor A, Spencer T, Spalding M, Lacambra C, Moller I (2015) Mangroves, tropical cyclones, and
coastal hazard risk reduction. In: Coastal and Marine Hazards, Risks, and Disasters,
J.F. Shroder, J.T. Ellis, and D.J. Sherman, eds. Elsevier. pp 403–429.
Mcleod E, Chmura GL, Bouillon S, Salm R, Björk M, Duarte CM, Lovelock CE, Schlesinger WH,
Silliman BR (2011) A blueprint for blue carbon: toward an improved understanding of the role
of vegetated coastal habitats in sequestering CO2. Front Ecol Environ 9: 552–560.
Menéndez P, Losada IJ, Beckb MW, Torres-Ortegaa S, Espejoa A, Narayanc S, Díaz-Simala P,
Langed G (2018) Valuing the protection services of mangroves at national scale: The
Philippines. Ecosys Serv 34: 24–36.
Murdiyarso D, Donato D, Kauffman JB, Kurnianto S, Stidham M, Kanninen M (2009) Carbon
storage in mangrove and peatland ecosystems. A preliminary account from plots in Indonesia. In
working paper 48 Center for International Forest Research (CIFOR): p. 39.
13 Mangrove as a Natural Barrier to Environmental Risks and Coastal Protection 321

Narayan S, Thomas C, Matthewman J, Shepard CC, Geselbracht L, Nzerem K, Beck MW (2019)


Valuing the Flood Risk Reduction Benefits of Florida’s Mangroves (The Nature Conservancy).
National Oceanic and Atmospheric Administration, NOAA (2020) From https://oceanservice.noaa.
gov/facts/stormsurge-stormtide.html
Nusantara MA, Malikusworo H, Helmi P (2015) Evaluation and planning of mangrove restoration
programs in Sedari Village of Kerawang District, West Java: Coastal Development Programs.
Procedia Environ Sci 23: 207–214.
Orchard SE, Stringer CL, Quinn CH (2015) Mangrove system dynamics in Southeast Asia: Linking
livelihoods and ecosystem services in Vietnam. Reg.Environ. Change Open access (Available
online from: http://link.springer.com/article/10.1007%2Fs10113-015-0802-5 http://dx.doi.org/
10.1007/s10113-015-0802-5
Othman MA (1994) Value of mangroves in coastal protection. In: Sasekumar A., Marshall N.,
Macintosh D.J. (eds) Ecology and conservation of Southeast Asian Marine and freshwater
environments including Wetlands. Developments in Hydrobiology, vol 98. Springer, Dordrecht.
Palacios Peñaranda, ML, Kintzb JRC, Salamancac NJP (2019) Carbon stocks in mangrove forests
of the Colombian Pacific. Estuar Coast Shelf Sci 227: 106299.
Pasquaud S, Vasconcelos RP, França S, Henriques S, Costa MJ, Cabral H 2015 Worldwide patterns
of fish biodiversity in estuaries: Effect of global vs. local factors Estuarine. Coastal and Shelf
Science 154: 122–128.
Prasteya G (2007) Protection from coastal erosion. In: Braatz, S., Fortuna, S., Broadhead, J., &
Leslie, R. (Eds.), Coastal protection in the aftermath of the Indian Ocean tsunami: What role for
forests and trees? Bangkok: Food and Agriculture Organization.
Pugh D (1987) Tides, surges and mean sea level. Published by John Wiley and Sons Ltd.
Quartel S, Kroon A, Augustinus P, Van Santen P, Tri NH (2007) Wave attenuation in coastal
mangroves in the Red River Delta, Vietnam. J Asian Earth Sci 29 (4): 576–584.
Rasmeemasmuang T, Sasaki J (2015) Wave reduction in mangrove forests: general information and
case study in Thailand. In: Chapter 24, Handbook of Coastal Disaster Mitigation for Engineers
and Planners. Elsevier: p 511–535.
Romañach SS, Donald L De A, Hock LH, Yuhong L, Su YT, Raja SRB, Lu Z (2018) Conservation
and restoration of mangroves: Global status, perspectives, and prognosis. Ocean Coast Manag
154: 72–82.
Saenger P (2002) Mangrove ecology, silviculture and conservation. Kluwer Academic Publishers,
Netherlands: p 360.
Saw AA, Kanzakia M (2015) Local livelihoods and encroachment into a mangrove forest reserve: a
case study of the Wunbaik reserved mangrove forest, Myanmar. Procedia Environ Sci 28:
483–492.
Shaltout KH, Ahmed MY, Alrumman Sulaiman A, Ahmed Dalia A, Eid Ebrahem M (2020)
Evaluation of the carbon sequestration capacity of arid mangroves along nutrient availability
and salinity gradients along the Red Sea coastline of Saudi Arabia. Oceanologia 62 (1): 56–69.
Spalding M, McIvor A, Tonneijck FH, Tol S, van Eijk P (2014) Mangroves for coastal defence.
Guidelines for coastal managers & policy makers. Published by Wetlands International and The
Nature Conservancy: p 42.
Spalding M, Kainuma M, Collins L (2010) World Atlas of mangroves. A collaborative project of
ITTO, ISME, FAO, UNEP-WCMC, UNESCO-MAB, UNU-INWEH and TNC. Earthscan,
London, UK: p 319.
Spalding M, Parrett CL (2019) Global patterns in mangrove recreation and tourism. Mar Pol 110:
103540.
Suratman MN (2008) Carbon sequestration potential of mangroves in southeast Asia. In: Bravo F,
LeMay V, Jandl R, von Gadow K (eds) Managing forest ecosystems. The challenge of climate
change. Springer Science, The Netherlands: p 297–315.
Suratman MN (2014) Remote sensing technology: recent advancements for mangrove ecosystem.
In Faridah Hanum et al. (Eds.), Mangrove ecosystems of Asia, Springer Science and Business:
p 295–317.
322 N. Asari et al.

Thampanya U, Vermaat JE, Sinsakul S, Panapitukkul N (2006) Coastal erosion and mangrove
progradation of Southern Thailand. Estuar Coast Shelf Sci 68: 75–85.
Trumper K, Bertzky M, Dickson B, van der Heijden G, Jenkins M, Manning P (2009) The natural
fix? The role of ecosystems in climate mitigation. A UNEP Rapid Response Assessment. United
Nations Environment Programme, UNEPWCMC, Cambridge, UK: p 65.
Umilia E, Asbar (2016) Formulation of mangrove ecosystem management model based on eco-
minawisata in the Coastal Sinjai, South Sulawesi. Social Behav Sci 227: 704–711.
Vo Quoc T, Kuenzer C (2012) Can Gio mangrove biosphere reserve evaluation 2012: current status,
dynamics and ecosystem services. IUCN, Ha Noi, Vietnam: p 100.
Wealth Accounting and the Valuation of Ecosystem Services, WAVES (2017) Valuing the
protection services of mangroves in the Philippines. WAVES Technical Report 2017.
Webster PJ (2005) Changes in tropical cyclone number, duration, and intensity in a warming
environment. Science 309: 1844–1846.
Winterwerp H, Wesenbeeck BV, Dalfsen JV, Tonneijck F, Astra A, Verschure S, Eijk PV (2014) A
sustainable solution for massive coastal erosion in Central Java. Towards Regional Scale
Application of Hybrid Engineering ~Discussion paper.
Woodroffe CD (2002) Coasts: form, process and evolution. Cambridge University Press,
Cambridge: p 623.
Yanagisawa H, Koshimura S, Miyag T, Imamura F (2010) Tsunami damage reduction performance
of a mangrove forest in Banda Aceh, Indonesia inferred from field data and a numerical model. J
Geophys Res-Oceans 115.
Zhang KQ, Liu H, Li Y, Hongzhou X, Jian S, Rhome J, Smith III TJ (2012) The role of mangroves
in attenuating storm surges. Estuar Coast Shelf Sci 102: 11–23.
Diversity and Community Structure
of Polychaetes in Mangroves of Indian 14
Coast

P. Murugesan and T. Balasubramanian

Abstract

This review article provides a comprehensive account of the diversity of


polychaetes associated to Indian mangrove ecosystems. Polychaetes constitute
an important component in marine benthic communities and play a major eco-
logical role in mangrove ecosystem. Information on polychaete diversity is
available only from 12 mangroves out of 19 on the east coast and only from
11 out of 21 those of west coast of India. Altogether 385 species of polychaetes
belonging to 176 genera under 49 families, representing 35.22% of the total
polychaetes in India, have been listed from mangroves skirting east and west
coasts of India. Of this total, as many as 231 species of polychaetes in east coast
and 230 species in mangroves of west coast of India were reported to occur. That
way, species of the following families such as Nereididae, Spionidae,
Terebellidae Sabellidae, Phyllodocidae, Eunicidae, Serpulidae, Lumbrinereidae,
Onuphidae, Syllidae, Maldanidae, Capitellidae, and Glyceridae were found to be
dominant in the mangrove environment. Among these, Genera Nereis, Glycera,
Lumbriconereis, Prionospio were found to be dominant followed by Phyllodoce,
Nephtys, Onuphis, Polydora, and Syllis in the next level of abundance. Regretta-
bly, despite its immense role, studies related to polychaete diversity have been
made only in 23 mangroves out of 40 mangroves in India. Compared to the
studies carried out in other parts of the world, especially Western hemisphere,
continuous survey is to be made along the mangroves of both the coasts of India
focusing polychaetes, which attach quite a lot of ecological significance and
economic importance.

P. Murugesan (*) · T. Balasubramanian


Faculty of Marine Sciences, Centre of Advanced Study in Marine Biology, Annamalai University,
Parangipettai, India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 323
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_14
324 P. Murugesan and T. Balasubramanian

Keywords

Polychaetes · Mangroves · West coast · East coast · Diversity

14.1 Introduction

Mangroves form a habitat for a wide variety of species, some occurring in high
densities and form food and shelter for a plethora of commercially important fin and
shell fishes. According to Forest Survey of India, mangroves occupy around
4921 km2 of the Indian coast. Of this total, 71.5% of the mangrove cover is found
in east coast and the remaining 28.5% in the west coast (FSI 2017). India has three
types of mangrove habitats, namely deltaic, backwater-estuarine, and insular. Com-
paring both the coasts of India, the east coast is shallow and the shore is quietly
shelving with several back water region and thus the mangroves development is
quite high in the East coast compared to West coast.

14.1.1 Importance of Mangroves and its Associated Faunal Groups

Mangroves are the distinctive source of ecology and are extremely important coastal
resources, which are vital for socio-economic development of the region
(Murugesan et al. 2018). They are productive habitats and support coastal fisheries.
As a detritus-based ecosystem, leaf litter from the mangroves provides the base for
adjacent aquatic and terrestrial food webs. Mangroves afford provisions and shelter
for a large number of commercially important fin and shellfishes. True to its sense,
the mangrove ecosystem serves as breeding, feeding, and nursery grounds for most
of the commercially important fin and shellfishes, on which thousands of coastal folk
depend for their livelihood (Manson et al. 2005). It is considered to have physical,
chemical, and biological processes which promote the adaptation of inhabiting
organisms to tolerate both the extremes of environmental variables both morpholog-
ically and physiologically.
Krom and Berner (1980) have reported that the decomposition of organic matter
consists of nutrients such as nitrogen and phosphorus, which play a vital role in the
establishment of healthy mangroves. However, sediment where the animals inhabit
often acts as buffer either as a source or sink of nutrients especially phosphorus by
adsorption–desorption reactions. Hence, the sediment plays a crucial role on benthic
faunal diversity in the mangrove ecosystem. In recent years, the mangrove
ecosystems have been the cynosure of Benthic Ecologists world over as many
researches is being focused towards carbon sequestration, one of the vogue words
of the day. Understandably, this ecosystem is known to sequester around 22.8 million
metric tons of carbon annually, covering 0.1% of the earth’s forests, which is equal
to 11% of terrestrial carbon into the ocean (Jennerjahn and Ittekkot 2002).
Under this circumstance, a detailed and complete knowledge of the bottom fauna
is not only important for the determination of productivity (Raveenthiranath Nehru
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 325

Jaw
Palp
Head/ Antenna
Prostomium Prostomium
Eye
Peristomium
Peristomium Cirrus
Segment
Anterior end of dorsal view
Antenna

Parapodia Peristomium Palp


Trunk/
Cirrus Prostomium
Metastomium

Seta Mouth

Peristomium

Parapodium
Anal
Cirrus

Anus Anterior end of Ventral view


Tail/
Pygidium Notopodium
Dorsal Cirrus
Aciculum Notochaetae

Neurochaetae

Vental Cirrus
Neuropodium
General Morphology of Polychaete
Parapodia of Polychaete

Fig. 14.1 Diagnostic characters of a typical polychaete worm

1990) but also helpful in understanding the diversity of the habitat. That way, the
macrofauna are the most widely studied group which is retained on 500 micron
sieve. They reside beneath the sediment surface in burrows and tubes. Thus,
seemingly, the bottom of the mangrove substratum habitats forms for an array of
macro-benthic organisms of various size and taxonomic categories. Unsurprisingly,
Indian mangrove ecosystems are known to have a total of 3985 biological species
that include 919 (23%) floral species and 3066 (77%) faunal species. Thus, the
faunal species component is about three times greater than the floral component of
the mangrove ecosystem (Kathiresan 2000; Kathiresan and Quasim 2005).
Mangroves are inhabited by a variety of macrobenthic invertebrates, such as
polychaetes, brachyuran crabs, gastropods, bivalves, hermit crabs, barnacles,
sponges, tunicates, and sipunculids. Mangrove invertebrates often exhibit marked
zonation patterns and colonize a variety of specific micro-environments. While some
species dwell on the sediment surface or reside in burrows, others live on
pneumatophores and lower tree trunks or prop-roots, burrow in decaying wood, or
can even be found in the tree canopies (Sasekumar 1974; Smith et al. 1991; Ashton
1999). Of these, polychaetes are found to be the major macro-benthic organisms in
mangrove environment. They are segmented worms belonging to the phylum
Annelida and Class Polychaeta. The characteristic features of typical polychaete
worm are shown in Fig. 14.1. They are habitually found to be the predominant taxa
of macro-benthic communities, which they typically constitute a major proportion of
the total macrofaunal diversity owing to their high adaptability (Fauchald 1977;
326 P. Murugesan and T. Balasubramanian

Ward and Hutchings 1996; Brunel 2005; Prabakaran et al. 2019) and their diet
consist of microbial, meiobial, and organic substances, in mangrove soil in which
they have certain special adaptations for survival, such as mucus secreting devices
which is used to protect themselves in unfavorable conditions in the estuarine
environment (Bandekar et al. 2017).

14.1.2 Role of Polychaetes in Mangrove Environment

Polychaetes form an important links between the primary detritus at the base of the
food web and consumers of higher trophic levels (Macintosh 1984). In view of their
greater abundance and biomass (secondary production), the energy assimilated by
the polychaetes play a significant role in nutrient recycling in the mangrove ecosys-
tem. They also serve as the food for a variety of demersal organisms of higher
trophic level. Most of the macrobenthos assist in the breakdown of particulate
organic material by exposing them to microbes and their waste materials contain
rich nutrients forming the food for other consumers. Thus, the macrobenthos in
general plays a major ecological role in the mangrove ecosystem (Warren and
Underwood 1986). Further, they are also secondary producers of mangroves subsoil
habitat production, which is crucial for tracing the biotic stability of the area from the
fisheries point of view as has been described by Murugesan et al. (2018). Justifiably,
they play a fundamental role in recycling of organic materials formed during the
decomposition of mangrove litter and subsequent detritus formation and nutrient
release, thereby improving the soil structure and its productivity through nutrient
cycling, which eventually exposed to microbes and their wastes forming food for
other consumers (Kumar 2003). Added to this, polychaetes are also used as veritable
pollution indicator groups for the detection of disturbance in an environment. Based
on their adaptability, certain groups of polychaetes, namely Spionids, Capitellids,
Maldanids are reported to occur exceedingly large in number in polluted
environments (Pearson and Rosenberg 1978; Prabakaran et al. 2019) and those of
Hesionids, Terebellids, Syllids reported to occur more in relatively pristine
environments (Khan and Murugesan 2005). Rightly, Samidurai et al. (2012)
reported that progression in macro-benthic communities is highly related to the
organic fortification and pollution in any given ecosystem. In addition to ecological
role, commercial prospects also exist for wealth creation as these worms are used
widely as feed for fin and shell fish brooders (Wang et al. 2018). True to this, their
lipid content may provide a source of essential PUFA (poly unsaturated fatty acids),
which are fundamental for the production of high quality seedlings of shell fishes.
On the whole, these worms have a pivotal role in the mangrove environments
(Muniasamy et al. 2013; Parvez et al. 2018).
Taking into account the facts stated supra, a ken on benthic faunal diversity in
general and polychaete diversity, in particular occurring in Indian mangroves will be
an immediate concern to understand the status of mangrove ecosystems. Accord-
ingly, an extensive literature survey on polychaete diversity in Indian mangroves has
been made and based on the available information and intensive Google search, this
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 327

review account pertinent to polychaetes diversity along the Indian mangroves has
been given.

14.2 Method of Collection of Polychaetes

There have been various kinds of broad relative studies regarding sampling strategy
and their capabilities (Ankar 1977; Eleftheriou and Holme 1984; Gage 1975;
Rosenberg 1978; Rumohr 1990). For professionally collecting the marine benthic
macro-faunal samples, the sampling instrument must be capable of penetrating to an
enough depth to collect the organisms present in the desired area of sampling. Most
of the earlier reports state that the greater part of species and individuals are to occur
in the upper 5–10 cm, even though large burrowing molluscs and crustaceans may be
found deeper and the pelagic polychaetes even collected from the plankton samples.
Accordingly the protocol is formulated and samples are collected in intertidal,
sub-tidal and shelf and slope regions. In the intertidal region, 625 cm2 wooden
quadrate is kept and the surface is dug out for 10 cm. Sub-tidal polychaetes can be
quantitatively collected using Peterson grab covering an area of about 1 m2. The
Peterson grab is found to be an efficient gear for sampling in shallow water
environments, both the mangrove and estuarine environments. In addition to sedi-
ment Grabs, Dredges, trawls and traps are also used for collecting benthos for
qualitative purpose. Pelagic forms will be collected by filtering around 200 l of
surface water and the sample will be calculated under a microscope using Sedgwick
rafter and the number of animals per cubic meter of water will be calculated. In
addition to these, a few groups of polychaetes can also be collected by hand-picking
method from intertidal region, floating wood, bottom mud, gastropod and bivalve
shells and various forms of hard substrates such as stones, cement boulders, wooden
piers, and boat hulls in the coastal zone.

14.2.1 Sieving, Treatment, Sorting, and Identification


of Polychaetes

After collection of sediment samples, the faunal components are to be segregated


from the sediment using set of sieves. The cost-effective and accurate size of sieve
has to be used. Once the mesh size has been set, the sediment samples collected will
be sieved through 0.5 mm aperture size mesh and the number of polychaetes
obtained in the sieve is counted. If there is any damage in fauna, the head alone
will be taken into account. After sieving, the sieve retains are to be collected in a
marked container and preserved in 5–7% (neutral) formaldehyde. Prior to extraction
of polychaetes, selective staining of the sample is to be done with a few drops of
Rose Bengal dye for the improved visibility and recognition of the species at the time
of identification (Pfannkuche and Thiel 1988). Although the use of rose Bengal is
not recommended by veteran benthic ecologists, it is undoubtedly helpful in sorting
and identification of the faunal communities. After adding the stain, the container is
328 P. Murugesan and T. Balasubramanian

to be turned upside down gently for complete mixing. As it is acidic, leaving the
fauna in it for too long may decay the faunal samples. To avoid this, the animals are
washed carefully with fresh water and then are preserved in 75% alcohol. The
concentration of alcohol should be gradually increased, i.e. 30%, 50%, 75%. Mag-
nesium chloride or propylene phenoxetol or menthol is also used as narcotics prior to
alcohol so as to prevent the distortion of the body parts.
After a day or two, the preserved faunal samples are smoothly but thoroughly
washed in fresh water to remove the formalin and salt concentration, preventing the
former from dissolving the shells of weak molluscs. For initial sorting, white
background trays with bench lights, needles, and pen brushes will be enough.
Dissection microscope (40 x) can also be used for sorting purposes. The key factor
for the achievement of sorting is not to add more samples in the sorting tray at a
given point of time as more the materials, the greater is the chance of overlooking
specimens. Therefore, adequate care is to be taken in every step since there is every
possibility of missing of many species.
For the sake of convenience, the Indian mangroves have been categorized into
(i) East coast and (ii) West coast mangroves and accordingly based on the available
published reports and intensive Google search, information on polychaete diversity,
state-wise and coast-wise is provided here.

14.3 Diversity of Mangrove Associated Polychaetes along the


Indian Coast

Globally, ~ 11,500 species of polychaetes belonging to 1417 Genera and 85 families


have been reported to occur, which account for >60% of the total macro-benthic
community. Of these species, 6033 species belong to the group Errantia, whereas
5085 and 158 species belong to the groups Sedentaria and Echiura, respectively.
Additionally, 180 species were from families currently outside of or as yet unas-
signed to, groupings in WoRMS, and referred to as Polychaeta incertae sedis, newly
created group (Pamungkas et al. 2019). In them, as many as 1093 species,
representing 8.66% of the total number of polychaete species, are known in India
(Achary et al. 2005). Among the twelve different mangroves habituated States or
Union Territories, which comprises of ~40 different mangrove ecosystems, infor-
mation on polychaete diversity is available only from 12 mangroves out of 19 on the
east coast and only from 11 out of 21 those of the west coast of India. Among the
mangroves of both the coasts of India, the maximum number of polychaetes was
recorded from the mangroves of Tamil Nadu (213) followed by those of Maharashtra
(189), West Bengal (85), Karnataka (64), Odisha (50), Andhra Pradesh (46), Kerala
(34), Gujarat (8), Pondicherry (5) in that order (Fig. 14.2). Regrettably, there is no
detailed study on the polychaete taxonomy from mangrove habituated states such as
Gujarat, Goa, Andaman and Nicobar, and Daman and Diu. Altogether 385 poly-
chaete species were recorded from the mangroves of east and west coasts of India. Of
these, 213 species belonged to group Errantia and the remaining 172 species to
group Sedentaria as the species are categorized into groups Errantia and Sedentaria
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 329

Fig. 14.2 State-wise contribution of polychaetes reported in Indian mangroves

Fig. 14.3 Family-wise contribution of polychaetes recorded in Indian mangroves

based on their mobility. Going through the available published information, it is


revealed that representatives of the following families’, viz., Phyllodocidae,
Syllidae, Nereidae, Eunicidae, Onuphidae, Glyceridae, Spionidae, Cirratulidae,
Terebellidae, Sabellidae, and Serpulidae are reported to occur very commonly in
the mangrove ecosystems (Fig. 14.3). Similarly, the following species were found to
be common in mangrove which includes Lepidonotus kumara, Dendronereis
aestuarina, Dendronereis arborifera, Diopatra neapolitana, Nereis
chingringhattensis, Malacoceros indicus, Marphysa mossambica, Heteromastus
similis, Paraheteromastus tenuis, Prionospio cirrifera, Prionospio pinnata, and
Euclymene annandalei.
330 P. Murugesan and T. Balasubramanian

14.4 Diversity of Polychaetes in Mangroves of East Coast of


India

In respect of mangroves of the east coast, only a few studies are available which
include the monumental work of Nandi and Choudhury (1983) who have made
pioneering attempt of quantitative assessment of benthic macrofauna of Sagar Island
skirting Sundarbans mangroves. They have reported 17 species of macrofauna of
which the polychaetes were one of the best represented taxa with the following
species, namely Lumbrinereis sp., L. polydesma, L. notocirrata, Diopatra
neapolitana, and Talehsapia annandalei. After them, Misra and Choudhury
(1985) studied the benthic fauna in selected locations of Sundarbans mangroves
and reported 30 species of polychaetes belonging to 24 genera under 13 families.
Kasinathan and Shanmugam (1986) made preliminary studies on the benthic
macrofaunal communities of Pichavaram mangroves, Tamil Nadu. Subsequently
Kathiresan (2000) recorded four different dominant polychaete species such as
Heteromastus similis, Euclymene annandale, Perinereis sp., Mercierella enigmatica
from Pichavaram mangroves, Tamil Nadu. Kumar (2001) reported 25 species of
polychaetes belonging to 20 genera from the selected locations of Sundarbans
mangroves, West Bengal. Samidurai et al. (2012) studied on the spatial and temporal
distribution of macrobenthos in different mangrove ecosystems, namely
(i) artificially developed mangroves along the stretch of Vellar estuary, (ii) riverine
(true) mangroves, and (iii) island mangrove ecosystems of Tamil Nadu, India.
Among the three ecosystems, a total of 46 macro-benthic species comprising four
groups of which polychaetes were found to be the most dominant group with
27 species followed by gastropods, bivalves, and crustaceans. Kumar and Khan
(2013) reported the following 6 species of polychaetes, namely Capitella capitata,
Marphysa sp., M. macintoshi, Namalycastis indica, Nereis sp., and Pseudonereis
variegata from artificially developed mangroves in Pondicherry. In their study,
Thilagavathi et al. (2013) listed as many as 188 species of polychaetes from three
different kinds of mangrove ecosystems of Tamil Nadu stretching ~650 km long
coast, namely (i) developing mangrove ecosystems (from Chennai to Pondicherry),
(ii) Riverine mangrove ecosystems (Parangipettai to Pichavaram and partly in
Muthupettai), and (iii) Island mangrove ecosystems (Palk bay and Gulf of Mannar
region). Contrastingly, Sekar et al. (2013) reported the following polychaete species
such as Chaetopterus sp. Pista sp. Nephtys sp. exceedingly large in number in
selected places along the southeast coast of India. Pravinkumar et al. (2013) reported
16 species of polychaetes from three different mangrove zones of Pichavaram, Tamil
Nadu. Among the polychaetes, Polydora sp. Exogone clavator, Pygospio elegans,
and Eyclymene sp., were found to be common throughout the study. Rao et al. (2015)
studied the macrofaunal diversity in the mangroves of the Port Blair Bay, South
Andaman Islands and of the various faunal taxa reported; polychaetes were one of
the best represented taxa with 13 species. Murugesan et al. (2016) made a compara-
tive study on benthic biodiversity of natural vis-à-vis artificially developed
mangroves of south east coast of India. The study revealed as many as 23 species
in natural (Pichavaram mangroves) mangroves and 19 species in artificially
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 331

developed mangroves (developed in Vellar estuary). Of both the regions,


polychaetes emerged as the dominant group with 14 species in Pichavaram followed
by other taxa. Very recently, Murugesan et al. (2018) studied the polychaete
taxonomy along the mangroves of the east coast of India and reported as many as
58 species of polychaetes, in which Sundarbans mangroves stood at top with
58 species followed by Pichavaram with 51, Bhitarkanika with 42 and Muthupettai
mangroves came last in the list with 34 species of polychaetes.

14.5 Diversity of Polychaetes in Mangroves of West Coast of


India

With respect to the mangroves skirting west coast, as early as during 1970’s,
Parulekar (1971) carried out pioneering studies on marine benthos of mangroves
of Maharashtra and reported as many as 46 species of polychaetes. After him,
Padmakumar (1984) studied on the ecology of Mangrove swamp near Juhu beach
of Bombay and reported 49 species of macrofauna and of which the annelids were
the dominant group with the following polychaetes as Nereis sp., Lumbrinereis
polydesma, Polydora sp., Goniodopsis incerta, Glycera sp., and Scolelepis
squamata. Subsequently, Kumar and Antony (1994) and Kumar (1995) recorded
33 polychaete species belonging to 20 genera under 10 families from mangrove
Swamps of Cochin, west coast of India. The errant polychaetes were found to be
more common than sedentaria group. Of these 33 species, Marphysa gravely,
Paraheteromastus tenuis, Dendronereis aestuarina, and D. heteropoda were the
most dominant species that occurred throughout the study. Further, Kumar (1999)
reported five different species of polychaetes from the Cochin mangrove ecosystem.
Subsequently, Kumar (2003) recorded as many as 43 species of polychaetes along
the major mangrove ecosystems such as Cochin—Bombay with 33 species in
Cochin mangroves and 10 in Bombay. Among the 15 families identified, the species
of the following families, viz., Nereidae (24 species), Eunicidae (20 species),
Glyeridae (8 species), Capitellidae (7 species), and Spionidae (7 species) were
found to be dominant. Saravanakumar et al. (2007) studied the benthic macrofaunal
assemblages of mangroves of Gulf of Kachchh and reported 62 species comprising
5 taxa. Of which, the polychaetes consisted of the following nine species, namely
Diopatra neapolitana, Eunice sp., Glycera alba, Lumbriconereis latreilli, Marphysa
stragulum, Nereis sp. Perinereis sp., Pulliella armata, Thalehasapia tenuis. Pati
et al. (2015) recorded a total of 180 species of polychaetes belonging to 113 genera
under 41 families and six orders from Maharashtra coast covering 6 coastal districts
(Thane, Mumbai-suburban, Mumbai, Raigad, Ratnagiri, and Sindhudurg) stretching
over 653 km long coastline. There are about 18 prominent creeks/back waters, many
of which having strong presence of mangrove cover and thus the compilation of Patil
et al. revealed a surge in the species richness to considerable species diversity from
46 species of Parulekar’s study to 180 species by Patil et al. Moving a little towards
north, in Karnataka, Bandekar et al. (2017) studied on the macro-benthic polychaetes
in mangroves regions along Kali estuary and reported as many as 61 species of
332 P. Murugesan and T. Balasubramanian

Fig. 14.4 Common polychaete species recorded in mangroves of east and west coast of India

polychaetes under 13 families. Of this, members of Nereidae, Spionidae, Eunicidae,


Glyceridae, Sabellaridae, and Terebellidae were found to be higher in number.
Similarly, species of genera Nereis, Lumbrinereis, and Glycera outnumbered the
other genera. Similarly, Satish et al. (2018) studied temporal variations in
polychaetes of the mangrove region of Shirgaon, Ratnagiri and reported a total of
six polychaetes species such as Nereis sp. Perinereis sp. Lumbrinereis sp. Marphysa
sp. - I, Marphysa sp. - II and a few un-identified polychaete sp. from the intertidal
mangrove area of Shirgaon, Ratnagiri, Maharashtra. Of this species, Nereis sp. and
Marphysa sp. showed their consistency in their occurrence throughout the study.
Looking at the commonness of the species occurrence, among the families, the
representative of the following families, viz., Nereididae (50 species), Spionidae
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 333

Fig. 14.5 Common polychaete species reported from east and west coast mangroves of India

(28 species), Lumbrineridae (21 species), Terebellidae (20 species), Sabellidae


(19 species), Syllidae (17 species), Capitellidae (17 species), Eunicidae (16 species),
Serpulidae (16 species), and Cirratulidae (12 species) were found to be dominant in
that order (Figs. 14.4 and 14.5).
Based on the foregoing account and available published reports, a total of
385 species of polychaetes belonging to 176 genera under 49 families are known
to occur in mangroves of east and west coasts of India (Table 14.1). This total covers
35.22% of the total polychaete species recorded in India. Comparing mangroves, as
many as 231 species of polychaetes were reported in various mangrove of east coast
and 230 in west coast mangroves.
Table 14.1 List of polychaetes reported in mangroves of east and west coasts of India
334

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
1 Abarenicola Arenicolidae Sedentaria +
gilchristi
2 Acoetes melanonota Acoetidae Errantia +
3 Aglaophamus Nephtyidae Errantia +
dibranchis
4 Ammotrypane Opheliidae Sedentaria
aulogaster
5 Ampharete capensis Ampharetidae Sedentaria +
6 Amphicteis gunneri Ampharetidae Sedentaria + + +
7 Amphicteis sp. Ampharetidae Sedentaria +
8 Amphinome Amphinomidae Errantia +
rostrata
9 Amphinome sp. Amphinomidae Errantia + + +
10 Ancistrosyllis Pilargidae Errantia + +
constricta
11 Ancistrosyllis Pilargidae Errantia +
groenlandica
12 Ancistrosyllis parva Pilargidae Errantia +
13 Ancistrosyllis sp. Pilargidae Errantia + + + +
14 Ancistrosyllis Pilargidae Errantia
constricta
15 Aphrodita alta Aphroditidae Errantia + +
16 Arabella iricolor Oenonidae Errantia + +
P. Murugesan and T. Balasubramanian
14

17 Arenicola Arenicolidae Sedentaria +


bombayensis
18 Arenicola loveni Arenicolidae Sedentaria +
19 Armandia Opheliidae Sedentaria + +
lanceolata
20 Armandia Opheliidae Sedentaria +
longicaudata
21 Axiothella Maldanidae Sedentaria + +
obockenisis
22 Bhawania Chrysopetalidae Errantia + +
cryptocephala
23 Bhawania goodei Chrysopetalidae Errantia +
24 Boccardia Spionidae Sedentaria + + + +
polybranchia
25 Brada villosa Flabelligeridae Sedentaria + + + +
26 Branchiocapitella Capitellidae Sedentaria + + +
singularis
27 Capitella capitata Capitellidae Sedentaria + + + + + + +
28 Capitella sp. Capitellidae Sedentaria + +
29 Ceratonereis Nereididae Errantia +
burmensis
30 Ceratonereis costae Nereididae Errantia + + + +
31 Ceratonereis Nereididae Errantia +
keiskama
32 Ceratonereis Nereididae Errantia + +
mirabilis
33 Chaetopterus Chaetopteridae Sedentaria +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

variopedatus
34 Chaetozone setosa Cirratulidae Sedentaria +
35 Chloeia flava Amphinomidae Errantia +
335

(continued)
Table 14.1 (continued)
336

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
36 Chloeia parva Amphinomidae Errantia +
37 Chloeia rosea Amphinomidae Errantia +
38 Chone collaris Sabellidae Sedentaria + + +
39 Chone filicaudata Sabellidae Sedentaria +
40 Chone letterstedti Sabellidae Sedentaria + + + +
41 Cirratulus Cirratulidae Sedentaria +
chrysoderma
42 Cirratulus cirratus Cirratulidae Sedentaria +
43 Cirratulus Cirratulidae Sedentaria + +
concinnus
44 Cirratulus gilchristi Cirratulidae Sedentaria +
45 Cirratulus sp. Cirratulidae Sedentaria + +
46 Cirriformia afer Cirratulidae Sedentaria +
47 Cirriformia Cirratulidae Sedentaria +
chrysoderma
48 Cirriformia filigera Cirratulidae Sedentaria +
49 Cirriformia Cirratulidae Sedentaria +
limnoricola
50 Cirriformia Cirratulidae Sedentaria + +
tentaculata
51 Cirrophorus Paraonidae Sedentaria + + + +
branchiatus
52 Cossura coasta Cossuridae Sedentaria + + + + +
53 Cossura delta Cossuridae Sedentaria +
P. Murugesan and T. Balasubramanian
14

54 Cossura Cossuridae Sedentaria +


longocirrata
55 Dasychone Sabellidae Sedentaria + +
cingulata
56 Dasychone Sabellidae Sedentaria + +
serratibranchis
57 Dendronereides Nereididae Errantia + + + +
heteropoda
58 Dendronereis Nereididae Errantia + + + + +
aestuarina
59 Dendronereis Nereididae Errantia + + + + + + +
arborifera
60 Diopatra claparedii Onuphidae Errantia + +
61 Diopatra cuprea Onuphidae Errantia + + +
62 Diopatra Onuphidae Errantia + + + + + +
neapolitana
63 Diopatra sp. Onuphidae Errantia +
64 Dipolydora coeca Spionidae Sedentaria + +
65 Dipolydora giardi Spionidae Sedentaria + +
66 Disoma orissae Trochochaetidae Sedentaria +
67 Dorvillea incertus Dorvilleidae Errantia +
68 Dorvillea neglecta Dorvilleidae Errantia +
69 Enice pennata Eunicidae Errantia +
70 Enice tentaculata Eunicidae Errantia +
71 Eteone barantollae Phyllodocidae Errantia +
72 Eteone ornata Phyllodocidae Errantia + +
73 Eteone siphodonta Phyllodocidae Errantia +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

74 Eteone sp. Phyllodocidae Errantia +


75 Eteone syphodonta Phyllodocidae Errantia +
337

(continued)
Table 14.1 (continued)
338

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
76 Euchone rosea Sabellidae Sedentaria +
77 Euclymene Maldanidae Sedentaria + +
annandalei
78 Euclymene Maldanidae Sedentaria
oerstedii
79 Euclymene sp. Maldanidae Sedentaria + + +
80 Eulalia macroceros Phyllodocidae Errantia +
81 Eulalia sanguinea Phyllodocidae Errantia +
82 Eunice antennata Eunicidae Errantia +
83 Eunice australis Eunicidae Errantia +
84 Eunice guttata Eunicidae Errantia +
85 Eunice indica Eunicidae Errantia +
86 Eunice laticeps Eunicidae Errantia +
87 Eunice savignyi Eunicidae Errantia +
88 Eunice sp. Eunicidae Errantia + + + + + + +
89 Eunice tubifex Eunicidae Errantia + + +
90 Euphrosine Euphrosinidae Errantia + +
capensis
91 Eurythoe Amphinomidae Errantia + + + + +
complanata
92 Eurythoe Amphinomidae Errantia + + +
parvecarunculata
93 Exoene sp. Syllidae Errantia +
94 Exogone clavator Syllidae Errantia + + + +
P. Murugesan and T. Balasubramanian
14

95 Exogone verugera Syllidae Errantia +


96 Fabrica filamentosa Fabriciidae Sedentaria + + +
97 Fabricia bansei Fabriciidae Sedentaria +
98 Fabriciola Fabriciidae Sedentaria +
mossambica
99 Ficopomatus Serpulidae Sedentaria +
uschakovi
100 Galathowenia Oweniidae Sedentaria +
oculata
101 Gattyana deludens Polynoidae Errantia +
102 Gaudichaudius Polynoidae Errantia +
cimex
103 Glycera africana Glyceridae Errantia +
104 Glycera alba Glyceridae Errantia + + + + + +
105 Glycera Glyceridae Errantia + + + +
benguellana
106 Glycera convoluta Glyceridae Errantia +
107 Glycera emerita Glyceridae Errantia
108 Glycera incerta Glyceridae Errantia +
109 Glycera longipinnis Glyceridae Errantia + + + + + +
110 Glycera onicornis Glyceridae Errantia +
111 Glycera rouxii Glyceridae Errantia +
112 Glycera sp. Glyceridae Errantia + + +
113 Glycera unicornis Glyceridae Errantia + + + +
114 Glycinde capensis Goniadidae Errantia +
115 Glycinde multidens Goniadidae Errantia +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

116 Glycinde oligodon Goniadidae Errantia + +


117 Goniada emerita Goniadidae Errantia + + + + +
118 Goniada goniada Goniadidae Errantia +
339

(continued)
Table 14.1 (continued)
340

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
119 Goniada sp. Goniadidae Errantia + +
120 Goniadella gracilis Goniadidae Errantia + +
121 Goniadopsis Goniadidae Errantia +
longicirrata
122 Goniadopsis Goniadidae Errantia +
maskallensis
123 Haploscoloplos Orbiniidae Sedentaria +
kerguelensis
124 Haplosyllis Syllidae Errantia +
spongicola
125 Harmothoe Polynoidae Errantia +
africana
126 Hermundura Pilargidae Errantia +
annandalei
127 Hesione intertexta Hesionidae Errantia +
128 Hesione pantherina Hesionidae Errantia +
129 Heteromastides Capitellidae Sedentaria +
bifidus
130 Heteromastus Capitellidae Sedentaria +
filiformis
131 Heteromastus Capitellidae Sedentaria + +
similis
132 Heteromastus sp. Capitellidae Sedentaria +
133 Hololepidella Polynoidae Errantia +
maculata
P. Murugesan and T. Balasubramanian
14

134 Hyalinoecia Onuphidae Errantia + + +


tubicola
135 Hyboscolex Scalibregmatidae Sedentaria + + +
longiseta
136 Hydroides albiceps Serpulidae Sedentaria + +
137 Hydroides Serpulidae Sedentaria +
diramphus
138 Hydroides Serpulidae Sedentaria + +
heteroceros
139 Hydroides Serpulidae Sedentaria +
homoceros
140 Hydroides Serpulidae Sedentaria +
norvegicus
141 Hydroides Serpulidae Sedentaria +
operculatus
142 Inermonephtys Nephtyidae Errantia +
inermis
143 Irmula spissipes Syllidae Errantia +
144 Isolda pulchella Melinnidae Sedentaria + +
145 Lanice socialis Terebellidae Sedentaria +
146 Laonice cirrata Spionidae Sedentaria + + + +
147 Laonome indica Sabellidae Sedentaria + +
148 Lavensia sp. Nereididae Errantia +
149 Leanira hystricis Sigalionidae Errantia +
150 Leocrates Hesionidae Errantia + +
claparedii
151 Leocratides ehlersi Hesionidae Errantia +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

152 Lepidonotus Polynoidae Errantia +


carinulatus
(continued)
341
Table 14.1 (continued)
342

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
153 Lepidonotus Polynoidae Errantia + + + +
tenuisetosus
154 Loimia medusa Terebellidae Sedentaria +
155 Lopadorhynchus Lopadorrhynchidae Errantia +
henseni
156 Lopadorhynchus Lopadorrhynchidae Errantia +
nationalis
157 Lumbriconereis Lumbrineridae Errantia + +
heteropoda
158 Lumbriconereis Lumbrineridae Errantia +
impatiens
159 Lumbriconereis Lumbrineridae Errantia + + + +
latreilli
160 Lumbriconereis Lumbrineridae Errantia + + +
polydesma
161 Lumbriconereis Lumbrineridae Errantia + + +
pseudobifilaris
162 Lumbriconereis Lumbrineridae Errantia + + +
simplex
163 Lumbriconereis sp. Lumbrineridae Errantia + + +
164 Lumbriconeries Lumbrineridae Errantia +
aberrans
165 Lumbrinereis Lumbrineridae Errantia +
brevicirra
P. Murugesan and T. Balasubramanian
14

166 Lumbrinereis Lumbrineridae Errantia +


magalhaensis
167 Lumbrineris Lumbrineridae Errantia + + +
albidentata
168 Lumbrineris Lumbrineridae Errantia +
bifilaris
169 Lumbrineris Lumbrineridae Errantia +
brevicirra
170 Lumbrineris Lumbrineridae Errantia +
hartmani
171 Lumbrineris Lumbrineridae Errantia +
heteropoda
172 Lumbrineris Lumbrineridae Errantia +
japonica
173 Lumbrineris Lumbrineridae Errantia +
polydesma
174 Lumbrineris Lumbrineridae Errantia +
pseudobifilaris
175 Lumbrineris Lumbrineridae Errantia +
simplex
176 Lycastis indica Nereididae Errantia +
177 Lycastonereis Nereididae Errantia +
indica
178 Magelona cincta Magelonidae Sedentaria + + + +
179 Magelona Magelonidae Sedentaria + +
papillicornis
180 Magelona rosea Magelonidae Sedentaria +
181 Malacoceros Spionidae Sedentaria + + + + + +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

indicus
182 Maldane sarsi Maldanidae Sedentaria + + + +
343

(continued)
Table 14.1 (continued)
344

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
183 Maldanella Maldanidae Sedentaria +
capensis
184 Maldanella grossa Maldanidae Sedentaria +
185 Marphysa gravelyi Eunicidae Errantia + +
186 Marphysa Eunicidae Errantia +
macintoshi
187 Marphysa Eunicidae Errantia +
mossambica
188 Marphysa Eunicidae Errantia +
sanguinea
189 Marphysa sp. Eunicidae Errantia + + +
190 Marphysa Eunicidae Errantia + +
stragulum
191 Mediomastus Capitellidae Sedentaria +
capensis
192 Mediomastus sp. Capitellidae Sedentaria +
193 Megaloma sp. Magelonidae Sedentaria
194 Megalomma Magelonidae Sedentaria + + +
quadrioculatum
195 Mercierella Serpulidae Sedentaria + +
enigmatica
196 Mesochaetopterus Chaetopteridae Sedentaria +
197 Micronephthys Nephtyidae Errantia +
sphaerocirrata
P. Murugesan and T. Balasubramanian
14

198 Micronereides Nereididae Errantia +


capensis
199 Minuspio cirrifera Spionidae Sedentaria + + +
200 Namalycastis Nereididae Errantia +
indica
201 Neanthes Nereididae Errantia +
chilkaensis
202 Neanthes Nereididae Errantia +
chingrighattensis
203 Neanthes Nereididae Errantia +
cricognatha
204 Neanthes sp. Nereididae Errantia + + + +
205 Neanthes Nereididae Errantia +
unifasciata
206 Neosabellaria Sabellariidae Sedentaria +
cementarium
207 Nephtys bucera Nephtyidae Errantia +
208 Nephtys capensis Nephtyidae Errantia +
209 Nephtys dibranchis Nephtyidae Errantia + + +
210 Nephtys lyrochaeta Nephtyidae Errantia +
211 Nephtys Nephtyidae Errantia +
oligobranchia
212 Nephtys Nephtyidae Errantia + + +
polybranchia
213 Nephtys sp. Nephtyidae Errantia +
214 Nephtys Nephtyidae Errantia +
sphaerocirrata
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

215 Nereis abbreviata Nereididae Errantia +


216 Nereis aibuhitensis Nereididae Errantia +
217 Nereis chilkaensis Nereididae Errantia +
345

(continued)
Table 14.1 (continued)
346

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
218 Nereis Nereididae Errantia +
chingrighattensis
219 Nereis cricognatha Nereididae Errantia +
220 Nereis diversicolor Nereididae Errantia + + +
221 Nereis falcaria Nereididae Errantia +
222 Nereis glandicincta Nereididae Errantia
223 Nereis granulata Nereididae Errantia +
224 Nereis jacksoni Nereididae Errantia +
225 Nereis kauderni Nereididae Errantia +
226 Nereis sp. Nereididae Errantia + + + + + + + + +
227 Nereis Nereididae Errantia +
talehsapensis
228 Nereis virens Nereididae Errantia +
229 Neries Nereididae Errantia + +
oligobranchia
230 Ninoe lagosiana Lumbrineridae Errantia +
231 Ninoe notocirrata Lumbrineridae Errantia +
232 Notocirrus Oenonidae Errantia +
brevicirrus
233 Notomastus Capitellidae Sedentaria + + + + +
aberans
234 Notomastus fauveli Capitellidae Sedentaria + +
235 Notomastus Capitellidae Sedentaria +
giganteus
P. Murugesan and T. Balasubramanian
14

236 Notomastus Capitellidae Sedentaria + + +


latericeus
237 Notoproctus Maldanidae Sedentaria + + + +
pacificus
238 Onuphis eremita Onuphidae Errantia + +
239 Onuphis eremite Onuphidae Errantia +
240 Onuphis Onuphidae Errantia + +
geophiliformis
241 Onuphis Onuphidae Errantia + +
holobranchiata
242 Onuphis sp. Onuphidae Errantia + + +
243 Ophelia africana Opheliidae Sedentaria +
244 Ophelia capensis Opheliidae Sedentaria +
245 Ophelid sp. Opheliidae Sedentaria +
246 Ophelina Opheliidae Sedentaria +
acuminata
247 Ophiodromous sp. Hesionidae Errantia +
248 Opisthosyllis sp. Syllidae Errantia +
249 Orbinia Orbiniidae Sedentaria + + + +
angrapequensis
250 Oriopsis eimeri Sabellidae Sedentaria +
251 Oriopsis neglecta Sabellidae Sedentaria +
252 Owenia fusiformis Oweniidae Sedentaria + +
253 Owenia sp. Oweniidae Sedentaria +
254 Panthalis oerstedi Acoetidae Errantia +
255 Paraheteromasus Capitellidae Sedentaria +
tenuis
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

256 Paralacydonia Paralacydoniidae Sedentaria +


weberi
347

(continued)
Table 14.1 (continued)
348

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
257 Paralepidonotus Polynoidae Errantia +
ampulliferus
258 Paraonidea sp. Paraonidae Sedentaria + + +
259 Paraonis sp. Paraonidae Sedentaria + + +
260 Paraprionospio Spionidae Sedentaria +
cristata
261 Paraprionospio Spionidae Sedentaria +
pinnata
262 Parheteromastus Capitellidae Sedentaria +
tenuis
263 Pectinaria crassa Pectinariidae Sedentaria +
264 Pelagobia Lopadorrhynchidae Errantia +
longicirrata
265 Perinereis Nereididae Errantia + + + +
cavifrons
266 Perinereis Nereididae Errantia + +
cultrifera
267 Perinereis Nereididae Errantia + + +
falsovariegata
268 Perinereis helleri Nereididae Errantia +
269 Perinereis Nereididae Errantia + +
nigropunctata
270 Perinereis nuntia Nereididae Errantia +
271 Perinereis nuntia Nereididae Errantia +
bombayensis
P. Murugesan and T. Balasubramanian
14

272 Perinereis nuntia Nereididae Errantia +


brevicirris
273 Perinereis nuntia Nereididae Errantia +
typica
274 Perinereis nuntia Nereididae Errantia +
vallata
275 Perinereis sp. Nereididae Errantia + + + + + + +
276 Perinereis Nereididae Errantia +
vancaurica
277 Petaloproctus Maldanidae Sedentaria +
terricolus
278 Pherusa monroi Flabelligeridae Sedentaria +
279 Phyllochaetopterus Chaetopteridae Sedentaria +
socialis
280 Phyllochaetopterus Chaetopteridae Sedentaria
sp
281 Phyllodoce Phyllodocidae Errantia +
longipes
282 Phyllodoce Phyllodocidae Errantia + + +
madeirensis
283 Phyllodoce Phyllodocidae Errantia +
malmgreni
284 Phyllodoce sp. Phyllodocidae Errantia + +
285 Phyllodoce tenuiss Phyllodocidae Errantia +
286 Phyllodoce tubicola Phyllodocidae Errantia +
287 Pisione africana Sigalionidae Errantia +
288 Pisionidens indica Sigalionidae Errantia + +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

289 Pista cristata Terebellidae Sedentaria + + + +


290 Pista herpini Terebellidae Sedentaria +
291 Pista indica Terebellidae Sedentaria + + +
349

(continued)
Table 14.1 (continued)
350

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
292 Pista Terebellidae Sedentaria +
pachybranchiata
293 Pista quadrilobata Terebellidae Sedentaria +
294 Pista typha Terebellidae Sedentaria +
295 Platynereis Nereididae Errantia +
calodonta
296 Platynereis Nereididae Errantia + + + +
dumerilii
297 Podarke Hesionidae Errantia +
angustifrons
298 Poecilochaetus Poecilochaetidae Sedentaria + +
serpens
299 Polycirrus Terebellidae Sedentaria +
plumosus
300 Polycirrus Terebellidae Sedentaria +
tribullata
301 Polydora capensis Spionidae Sedentaria +
302 Polydora ciliata Terebellidae Sedentaria + +
303 Polydora hophura Terebellidae Sedentaria +
304 Polydora kempi Terebellidae Sedentaria +
305 Polydora sp. Terebellidae Sedentaria + + + + + +
306 Polydra ciliata Terebellidae Sedentaria +
307 Polyphysia crassa Scalibregmatidae Sedentaria + + +
P. Murugesan and T. Balasubramanian
14

308 Pomatoceros Serpulidae Sedentaria +


triqueter
309 Pomatoleios Serpulidae Sedentaria + + + +
kraussii
310 Potamilla Sabellidae Sedentaria +
leptochaeta
311 Praxillella affinis Maldanidae Sedentaria +
pacifica
312 Prionospio Spionidae Sedentaria +
aucklandica
313 Prionospio cirrifera Spionidae Sedentaria + + + + +
314 Prionospio Spionidae Sedentaria + + + +
cirrobranchiata
315 Prionospio Spionidae Sedentaria +
malmgreni
316 Prionospio pinnata Spionidae Sedentaria + + + + + +
317 Prionospio Spionidae Sedentaria + + +
polybranchiata
318 Prionospio Spionidae Sedentaria + + + +
saldanha
319 Prionospio Spionidae Sedentaria + + + +
sexoculata
320 Prionospio sp. Spionidae Sedentaria + +
321 Protodorvillea Dorvilleidae Errantia +
biarticulata
322 Protodorvillea Dorvilleidae Errantia +
egena
323 Protula tubularia Serpulidae Sedentaria +
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

324 Pseudonereis Nereididae Errantia + +


variegata
351

(continued)
Table 14.1 (continued)
352

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
325 Pseudopolydora Nereididae Errantia +
kempi
326 Pulliella armata Capitellidae Sedentaria + + +
327 Rhodine gracilior Maldanidae Sedentaria +
328 Sabella sp. Sabellidae Sedentaria + + + +
329 Sabella Sabellidae Sedentaria +
spallanzanii
330 Sabellaria Sabellidae Sedentaria +
cementarium
331 Sabellaria intoshi Sabellidae Sedentaria + + + +
332 Sabellaria Sabellidae Sedentaria +
pectinata
333 Sabellaria sp. Sabellidae Sedentaria +
334 Sabellaria Sabellidae Sedentaria +
spimnulosa
335 Sabellastarte longa Sabellidae Sedentaria +
336 Salmacina dysteri Serpulidae Sedentaria +
337 Scolelepis Spionidae Sedentaria + + + +
squamata
338 Scoloplella Spionidae Sedentaria + +
capensis
339 Scoloplos armiger Orbiniidae Sedentaria +
340 Scoloplos Spionidae Sedentaria + +
marsupialis
P. Murugesan and T. Balasubramanian
14

341 Scoloplos sp. Spionidae Sedentaria +


342 Scyphoproctus Capitellidae Sedentaria +
djiboutiensis
343 Serpula Serpulidae Sedentaria + + +
vermicularis
344 Sigalion Sigalionidae Errantia +
squamatum
345 Sigambra Pilargidae Errantia +
constricta
346 Sigambra parva Pilargidae Errantia +
347 Sphaerosyllis Syllidae Errantia +
erinaceu
348 Sphaerosyllis Syllidae Errantia +
sublaevis
349 Spio bengalensis Spionidae Sedentaria +
350 Spio filicornis Spionidae Sedentaria + + + +
351 Spiochaetopterus Spionidae Sedentaria +
costarum
352 Spiophanes bombyx Spionidae Sedentaria +
353 Spiophomianes Spionidae Sedentaria +
soderstromi
354 Spirobrachus Serpulidae Sedentaria
teraceros
355 Spirobranchus Serpulidae Sedentaria +
kraussii
356 Spirorbis Serpulidae Sedentaria + +
foraminosus
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast

357 Sternaspis scutata Sternaspidae Sedentaria + + + + + +


358 Sthenelais boa Sigalionidae Errantia + + +
359 Sthenelais japonica Sigalionidae Errantia + +
353

(continued)
Table 14.1 (continued)
354

East coast West coast


Andaman
and
West Nicobar Andra Tamil Daman
S. No Polychaete species Family Group Bengal Islands Pradesh Odisha Nadu Pondicherry Gujarat Maharashtra Goa Kerala Karnataka and Diu
360 Streblosoma Persia Terebellidae Sedentaria + + +
361 Streblospio Spionidae Sedentaria +
benedicti
362 Stylarioides Terebellidae Sedentaria +
stylarioides
363 Syllidia armata Hesionidae Errantia +
364 Syllis armillaris Syllidae Errantia +
365 Syllis benguellana Syllidae Errantia + + + +
366 Syllis cornuta Syllidae Errantia +
367 Syllis gracilis Syllidae Errantia + + + + +
368 Syllis hyalina Syllidae Errantia +
369 Syllis longocirrata Syllidae Errantia +
370 Syllis sp. Syllidae Errantia + + + +
371 Syllis trifalcata Syllidae Errantia +
372 Syllis variegata Syllidae Errantia +
373 Talehsapia Pilargidae Errantia + +
annandalei
374 Tambalagamia Nereididae Errantia +
orientalis
375 Terebella Terebellidae Sedentaria +
pterochaeta
376 Terebellid sp. Terebellidae Sedentaria +
377 Terebellides Terebellidae Sedentaria + + + + + +
stroemi
P. Murugesan and T. Balasubramanian
14

378 Tharyx filibranchia Cirratulidae Sedentaria +


379 Thelepus setosus Terebellidae Sedentaria +
380 Tylonereis fauveli Nereididae Errantia +
381 Tomopteris Tomopteridae Errantia +
helgolandica
382 Travisiopsis Typhloscolecidae Errantia +
lobifera
383 Typhloscolex Typhloscolecidae Errantia +
muelleri
384 Vanadis Formosa Phyllodocidae Errantia +
385 Vermiliopsis Sabellidae Sedentaria +
glandigerus
Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast
355
356 P. Murugesan and T. Balasubramanian

14.6 Unknown/Way Forward

Going through the information stated above, it is quite apparent that detailed studies
related to the macrobenthos in general and polychaete diversity, in particular, have
been made only less than 60% of the mangroves skirting both the coasts of India.
Added to this woe, there is a great shortage of updated taxonomical information and
recent literature barring the classical works of Fauvel (Fauvel, 1953) and Day (Day,
1967) to carry out benthic investigations. Advantageously, this group occurs almost
in all the mangrove environments. Nevertheless, the only sketchy information is
available in many mangroves. To tide over this problem, the following aspects on
biodiversity of polychaetes of mangroves need to be given focus:

1. Continuous survey on cataloguing and spatial variability of polychaetes along the


mangroves of Indian coast.
2. The extent of their utility in assessing habitat deterioration, especially in a few
mangroves like Pulicat stretch, Pichavaram, and Muthupettai mangroves (shrimp
farm discharges), Godavari belt in Andhra, etc., is to be studied.
3. Comparison of polychaete diversity in relation to various mangrove zones like
Avicennia, Rhizophora, etc., so as to understand the harboring nature of the
mangroves.
4. Information on pollution indicator species (r-selected species/opportunistic spe-
cies and t-selected species from intensively exploited mangrove environs).
5. Genetic diversity of polychaetes among the various mangroves.
6. Co-variations in polychaete diversity with that of other taxa like nematodes and
amphipods and their utility in ecological health assessment.

14.7 Conclusion

Based on the forgoing account, it is concluded that the present review yielded a fairly
good amount of information on the benthic biodiversity in general and polychaete
diversity in particular in the mangroves of the Indian coast. Further, as there is no
comprehensive account of the diversity of polychaetes from the mangroves of the
Indian coast, a comparison was done only based on the existing sporadic information
and therefore a definite conclusion could not be drawn. On the other hand, reports
related to the taxonomy of benthic faunal community are limited as the researchers
worldwide do not evince interest in this line besides the enrolment of a young
generation of benthic taxonomists has also been poor in the recent past. There are
quite a lot of reasons for this: (i) unconcerned perspectives, both in society and
educational systems, and (ii) organisms that are “unseen” from the outlook of
instant, cost-effective and medical interest to man and more significantly poor
financial support from the Government. To accomplish this, there is a dire need to
establish a strong collaboration of benthic researchers among the Asian countries is
the need of the hour. Once this is done and studies are initiated, certainly this will
throw an important beam of light on the Polychaete taxonomy in the mangroves with
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 357

a scope to develop management strategies as well as to attain momentous


conclusions for the policy makers. The conservation troubles pertaining to this
community can also be addressed when biosystematics studies are well recognized.

Acknowledgments The authors are thankful to the Director and Dean, Centre of Advanced Study
in Marine Biology, Annamalai University, for the encouragement and facilities provided. We
gratefully acknowledge the Ministry of Environment, Forest and Climate Change, Govt. of India,
New Delhi, for the financial support (MoEn&F No:22-11/2009-CS.I).

References
Achary GP, Chakravarty G, Chakraborty SK, Jayasurya PK, Sarala Devi K (2005) Benthos-
polychaetes. CMFRI Special Publication Mangrove ecosystems: a manual for the assessment
of biodiversity. 83: 25–145.
Ankar S (1977) Digging profile and penetration of the Van Veen grab in different sediment types.
University of Stockholm, Department of Zoology and Askö Laboratory. 16: 3–22.
Ashton EC (1999) Biodiversity and community ecology of mangrove plants, molluscs and
crustaceans in two mangrove forests in Peninsular Malaysia in relation to management
practices. Ph.D. Thesis. University of York, UK.
Bandekar PD, Naik UG, Haragi SB (2017). Diversity status of benthic macro polychaetes species in
estuarine region of Karwar, West Coast of India. Int J Fish Aqua Stud 5(1): 216–219.
Brunel P (2005). Visages de la Biodiversite marine. Vertigo 6: 1–13.
Eleftheriou A, Holme NA (1984) Macro fauna techniques. Methods for the study of marine benthos.
p 140–216.
Fauchald K (1977) The polychaete worms. Definitions and keys to the orders, families and genera.
Nat Hist Mus LA County Mus Sci Ser 28: 1–190.
FSI (2017) India state of forest report. Forest Survey of India, Dehradun, p 55–61.
Gage JD (1975) A comparison of the deep-sea epibenthic sledge and anchor-box dredge samplers
with the van Veen grab and hand coring by diver. Deep-Sea Res Oceanogr Abstr 22(10):
693–702.
Jennerjahn TC, Ittekkot V (2002) Relevance of mangroves for the production and deposition of
organic matter along tropical continental margins. Naturwissenschaften 89(1): 23–30.
Kathiresan K, Quasim SZ (2005) Biodiversity of mangrove ecosystems (p. 251). New Delhi:
Hindustan Publishing Corporation.
Kasinathan R, Shanmugam A (1986) Benthic macrofauna of the Pichavaram mangroves, South
India. J Annamalai University Sci 34: 109–119.
Kathiresan K (2000) A review of studies on Pichavaram mangrove, southeast coast of India.
Hydrobiol 430(1–3): 185–205.
Khan SA, Murugesan P (2005) Polychaete diversity in Indian estuaries. Indian J Geo-Mar Sci 34
(1): 114–119.
Krom MD, Berner RA (1980) Adsorption of phosphate in anoxic marine sediments. Limnol
Oceanogr 25(5): 797–806.
Kumar, RS (1995) Macrobenthos in the mangrove ecosystem of Cochin backwaters, Kerala
(southwest coast of India). Indian J Geo-Mar Sci 24(2): 56–61.
Kumar R (1999) New record of five annelids (Class: Polychaeta) from the mangrove habitat of the
south west coast of India. J Mar Biol Assoc India 41(1&2): 116–118.
Kumar R (2001) Biodiversity and affinity of polychaetous annelids within the mangrove ecosystem
of Indo-pacific region. J Mar Biol Ass India 43(1&2): 206–213.
Kumar RS (2003) A checklist of polychaete species some mangroves of Asia. Zoos’ Print J 18:
1017–1020.
358 P. Murugesan and T. Balasubramanian

Kumar RS, Antony A (1994) Impact of environmental parameters on polychaetous annelids in the
mangrove swamps of Cochin, South West coast of India. Indian J Geo-Mar Sci 23(3): 137–142
Kumar PS, Khan AB (2013). The distribution and diversity of benthic macro invertebrate fauna in
Pondicherry mangroves, India. Aqua Biosyst 9(1): 15.
Macintosh DJ (1984) Ecology and productivity of Malaysian mangrove crab populations
(Decapoda: Brachyura). Proceedings of the Asian Symposium on Mangrove Environmental
Research and Management 354–377.
Manson FJ, Loneragan NR, Skilleter GA, Phinn SR (2005) An evaluation of the evidence for
linkages between mangroves and fisheries: a synthesis of the literature and identification of
research directions. Oceanogr Mar Biol 43: 483.
Misra A, Choudhury A (1985) Polychaetous annelids from the mangrove swamps of Sunderbans,
India. In Pro Nat Symp Biol Util Cons Mangroves (Ed. LJ Bhoslae), p 448–452.
Muniasamy M, Muthuvelu S, Balachandar K, Murugesan P (2013) Diversity of benthic fauna in
Coleroon estuary, south east coast of India. Inter J Recent Sci Res 4(10): 1617–1621.
Murugesan P, Pravinkumar M, Muthuvelu S, Ravichandran S, Vijayalakshmi S, Balasubramanian
T (2016) Benthic biodiversity in natural vis-a-vis artificially developed mangroves of south east
coast of India. Indian J Geo-Mar Sci 45(8): 1049–1058.
Murugesan P, Sarathy PP, Muthuvelu S, Mahadevan G (2018) Diversity and distribution of
polychaetes in mangroves of east coast of India. In: Sahadev Sharma (ed) Mangrove ecosystem
ecology and function, IntechOpen, UK, p 107–130.
Nandi S, Choudhury A (1983) Quantitative studies on the benthic macrofauna of Sagar Island,
intertidal zones, Sunderbans, India. Mahasagar 16(3): 409–414.
Padmakumar KG (1984) Ecology of a mangrove swamp near Juhu Beach, Bombay with reference
to sewage pollution. Unpublished PhD thesis submitted to the University of Bombay.
Pamungkas J, Glasby CJ, Read GB, Wilson SP, Costello MJ (2019) Progress and perspectives in the
discovery of polychaete worms (Annelida) of the World. Helgol Mar Res 73(1): 4.
Parulekar AH (1971) Polychaetes from Maharashtra and Goa. J Bombay Nat History Soc 68:
726–749.
Parvez S, Long MJ, Poganik JR, Aye Y (2018) Redox signaling by reactive electrophiles and
oxidants. Chem Rev 118(18): 8798–8888.
Pati SK, Swain D, Sahu KC, Sharma RM (2015) Diversity and distribution of polychaetes
(Annelida: Polychaeta) along Maharashtra Coast, India. In: Rawat M, Dookia S, Sivaperuman
C (eds) Aquatic ecosystem: biodiversity, ecology and conservation, Springer, New Delhi,
p 53–65.
Pearson TH, Rosenberg R (1978) Macrobenthic succession in relation to organic enrichment and
pollution of the marine environment. Oceanogr Mar Biol 16: 229–311.
Pfannkuche O, Thiel H (1988) Sample processing. Introduction to the study of meiofauna, 9:
134–145.
Prabakaran JR, Vibin A, Deivakumari M, Muruganantham M, Ramasubburayan R, Palavesam A,
Immanuel G (2019) Comparison of polychaete diversity and distribution along the south Tamil
Nadu coast (Lat. 8.08∘ to 10.79∘ N), India. Regional Stud Mar Sci 28: 100564.
Pravinkumar M, Murugesan P, Prakash RK, Elumalai V, Viswanathan C, Raffi SM (2013) Benthic
biodiversity in the Pichavaram mangroves, Southeast Coast of India. J Oceanog Mar Sci 4(1):
1–11.
Rao VM, Kumar AT, Ghosh S (2015) Studies on the mangrove macro faunal diversity and
assessment among different sites in port Blair Bay, South Andaman Islands. Fisher Aqua J 6
(2): 1.
Raveenthiranath N (1990) Ecology of macrobenthos in and around Mahandrapalli region of
Coleroon estuary, Southeast coast of India. Ph.D. thesis, Annamalai University, India: 231.
Rosenberg DM (1978) Practical sampling of freshwater macro zoobenthos: a bibliography of useful
texts, reviews, and recent papers. Canadian Fisheries and Marine Service Technical Report
No. 790, p 15.
14 Diversity and Community Structure of Polychaetes in Mangroves of Indian Coast 359

Rumohr H (1990) Soft bottom macrofauna: Collection and treatment of samples. ICES Tech Mar
Environ Sci p 19.
Samidurai K, Saravanakumar A, Kathiresan K (2012) Spatial and temporal distribution of
macrobenthos in different mangrove ecosystems of Tamil Nadu Coast, India. Environ Monit
Assess 184(7): 4079–4096.
Saravanakumar A, Serebiah JS, Thivakaran GA, Rajkumar M (2007) Benthic macrofaunal assem-
blage in the arid zone mangroves of Gulf of Kachchh-Gujarat. J Ocean U China 6(3): 303–309.
Satish KM, Adsul AD, Indulkar ST, Pai R (2018) Seasonal variation in polychaetes of intertidal
mangrove area of Shirgaon, Ratnagiri, Maharashtra, India. Int J Life Sci 6(1): 156–160.
Sasekumar A (1974) Distribution of macrofauna on a Malayan mangrove shore. J Anim Ecol 43:
51–69.
Sekar V, Prithiviraj N, Savarimuthu A, Rajasekaran R (2013) Macrofaunal assemblage on two
mangrove ecosystems, southeast coast of India. Inter J Recent Sci Res 4(5): 530–535.
Smith TJ, Boto KG, Frusher SD, Giddins R (1991) Keystone species and mangrove forest
dynamics: the influence of burrowing by crabs on soil nutrient status and forest productivity.
Estuar Coast Shelf Sci 33: 19–32.
Thilagavathi B, Varadharajan D, Babu A, Manoharan J, Vijayalakshmi S, Balasubramanian T
(2013) Distribution and diversity of macrobenthos in different mangrove ecosystems of Tamil
Nadu Coast, India. J Aqua Res Develo 4(6): 1–12.
Wang CA, Hu G, Sun P, Gu W, Wang B, Xu Q, Liu H (2018) Effects of dietary protein at two lipid
levels on growth, gonadal development, body composition and liver metabolic enzymes of
brown trout (Salmo trutta fario) broodstock. Aqua Nutr 24(5): 1587–1598.
Ward TJ, Hutchings PA (1996) Effects of trace metals on infaunal species composition in polluted
intertidal and sub-tidal marine sediments near a lead smelter, Spencer Gulf, South Australia.
Mar Ecol Prog Ser 135: 123–135.
Warren JH, Underwood AJ (1986) Effects of burrowing crabs on the topography of mangroves
swamps in New South Wales. J Exp Mar Biol Eco 102: 223–235.
Structure and Diversity of Plants
in Mangrove Ecosystems 15
Nurun Nadhirah Md Isa and Mohd Nazip Suratman

Abstract

Mangroves are among the most productive, prominent and complex ecosystems
which comprised of salt-tolerant trees and shrubs. These complex ecosystems are
found between the latitudes of 32 north and 38 south, along the tropical coast of
Africa, Australia, Asia and Americas. Mangrove forests are rich with diversity of
flora and fauna. However, these unique ecosystems also challenged with destruc-
tion as a result from a variety of human activities such as aquaculture and effects
from global climate change. This review paper highlights the distribution of
world mangroves, including their structure and diversity as well as the threats
facing by them. It is recommended that the mangrove ecosystems should always
be protected and restored in order to sustain these ecosystems for biodiversity
conservation and other ecosystem services.

Keywords
Conservation, Diversity · Mangrove · Structure · Threat

N. N. Md Isa
Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM) Pahang Branch, Jengka,
Pahang, Malaysia
M. N. Suratman (*)
Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 361
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_15
362 N. N. Md Isa and M. N. Suratman

15.1 Introduction

Mangrove forest is one of the primary features of coastal ecosystems throughout the
tropical and subtropical regions of the world. It has been previously described as
“Coastal Woodland” and “Intertidal forest” (Aksornkoae 1995). Joshi and Ghose
(2014) referred mangrove as a group of taxonomically heterogeneous woody shrubs
and trees growing in the intertidal zone of tropical and subtropical coasts. Usually,
mangrove vegetations are dicotyledonous woody shrubs or trees that are virtually
confined to the tropics. They often form dense forests that dominate intertidal muddy
shores, frequently consisting of virtually monospecific patches (Hogarth 2007).
The vascular plants in mangrove forests have special morphological, physiologi-
cal and other non-visible adaptations to live in a saline intertidal environment
dominated by low dissolved oxygen or sometimes anoxic fine sediments. These
plants, composed with their complement of microorganisms and animals, form the
mangrove ecosystem. The term mangrove thus refers both to the plants themselves
as well as to the ecosystem. Regularly, plants which occur in the non-mangrove
ecosystem and with none or only a few of these morphological adaptations are also
found in the mangrove forests (Ong and Gong 2013). This chapter reviews the
structure and composition of mangrove ecosystem along with the species diversity
and distribution in different locations in the world.

15.2 Distribution of World Mangroves

Mangroves are located in a highly productive intertidal zone in more than


120 countries in the tropics and subtropics (Gandhi and Jones 2019). Worldwide,
there are a total of 114 species of true mangroves belonging to 66 genera with
species richness being greatest in the Indo-Pacific region (Tomlinson 1986).
Figure 15.1 shows mangrove distribution zones and the number of mangrove species
within each region. The most diverse mangrove species and largest mangrove areas
in the world is in Southeast Asia with 6.8 million hectares. The largest areas of
mangrove are found in Indonesia, Malaysia, Myanmar, Papua New Guinea and

Fig. 15.1 World map of the mangrove distribution zones and the number of mangrove species
along each region (Deltares 2014)
15 Structure and Diversity of Plants in Mangrove Ecosystems 363

Thailand (Faridah-Hanum et al. 2012). Most of the largest mangrove areas, particu-
larly in Asia, are in the deltas of major rivers. Examples include the Indus delta of
Pakistan, the Sundarbans where the Ganges and Brahmaputra Rivers flow into the
Bay of Bengal, the Merbok and Matang deltas of Peninsular Malaysia, the Fly River
of Papua New Guinea and, in South America, the mangroves of the Amazon delta.
Thirty-five percent of the total 18 million ha of global mangrove forests are found
in the Southeast Asian countries of Brunei Darussalam, Cambodia, Indonesia,
Malaysia, Myanmar, the Philippines, Thailand and Vietnam. Indonesia alone has
4.5 million ha of mangroves (Primavera et al. 2000). Characteristically, they have a
low tidal range and strong freshwater flow carrying a considerable load of sediment,
much of which is deposited in the mangroves (Hogarth 2007).

15.3 Structure of Mangrove Ecosystems

Generally, mangroves can be divided into two groups which are exclusive and
non-exclusive. Exclusive mangroves are restricted to intertidal areas, while
non-exclusive mangroves are distributed in terrestrial or aquatic but also occur in
the typical mangrove’s environment. True or exclusive mangrove only can be found
in certain habitats of mangroves (Hogarth 2007). While, non-exclusive mangrove
also recognized as semi-mangrove, back mangrove or mangrove associates (Wang
et al. 2011). Mangrove forests literally live in two worlds at once, acting as the
interface between land and sea.
Adaptation to salt tolerance is of three types which are salt excluders, salt
secretors and salt accumulators. The salt excluders are from the members of the
genera like Rhizophora, Bruguiera and Ceriops from the family of Rhizophoraceae.
The species in the genera Acanthus, Aegialitis, Aegiceras and Avicennia have salt-
secreting glands on the leaf surface. The species like Sonneratia, Xylocarpus and
Excoecaria are the salt accumulators. Various types of root adaptations of
mangroves in the habitat are lenticels (Bruguiera spp.), pneumatophores (Sonneratia
spp. and Avicennia spp.), knee roots (Bruguiera spp.), cable roots (Avicennia spp.),
and stilt roots (Rhizophora spp.) (Ong and Gong 2013).
Zonation in mangrove forests often has been attributed to the responses of
individual species to variation in degree of tidal inundation, salinity or other mea-
surable edaphic gradients that vary across the intertidal (Ellison et al. 2000). Man-
grove species zonation can be considered at different scales. On a tide-dominated
shore, a clear vertical sequence of species often appears. An example of a general
mangrove profile is shown in Fig. 15.2. One of mangrove main species, Avicennia
often has a bimodal distribution, being abundant near the seaward margin and some
way up to the shore areas. Rhizophora is distributed next to the seaward zone of the
profile, followed by Bruguiera and Ceriops.
Mangrove vegetation in Southeast Asia may range from 1 to 2 m tall. Avicennia
alba and A. marina stands on the seaward side of accreting shores can achieve up to
30 40 m tall stands along with the mixed species of Bruguiera-Rhizophora stands.
On more exposed areas, but not eroding coastlines, one may find Sonneratia alba
364 N. N. Md Isa and M. N. Suratman

HWS

Rhizophora Bruguiera Ceriops


Seaward Landward
zone Mesozore zone Terrestrial zone

Fig. 15.2 Profile of a shore in north-eastern Australia (an area of high rainfall) that shows
mangrove zonation (Tomlinson 1986)

and A. alba. While, along waters of lower salinity Nypa fruticans, Cerbera odollam
and S. caseolaris are common. Apart from saplings, undergrowth has been often
scarce, but certainly not absent, and species such as sea holly Acanthus ilicifolius and
mangrove fern, Acrostichum aureum may be common along the banks of streams
and in disturbed area. However, mangrove vegetation structure may vary depending
on the location of the mangrove ecosystem, management and disturbance (Sarno
et al. 2015). Moreover, climatic factors are the key in controlling the worldwide
distribution of mangroves including topography, soil properties and tide fluctuations
(Abou Seedo et al. 2017).

15.4 Diversity of Mangrove Plants

Mangroves have a diverse group of trees and shrubs that flourish in flooded and
saline habitats (Hogarth 2007). Diversity is a community that has different
characteristics with other communities. The characteristics of the community in an
environment are biodiversity, the more diverse the biotic component (biodiversity),
the higher the diversity. The greatest diversity of mangrove plant species exists in
Southeast Asia. According to Giesen et al. (2006), the largest plant families recorded
in Southeast Asia mangroves are Rhizophoraceae and followed by Orchidaceae and
Asclepiadaceae. The smallest number of species was from Sonneratiaceae family
with five species. According to Tomlinson (1986), Rhizophoraceae is the most
dominant family because it has widespread distribution globally. Moreover, the
family has also the adaptability to both extreme as well as non-extreme conditions.
Therefore, in terms of floristic composition, species in this family is common and
easily found in mangrove habitats.
From the total of 52 mangrove species recorded globally, 48 of the species occur
in Indonesia, which is considered to be the most diverse of the Southeast Asia
15 Structure and Diversity of Plants in Mangrove Ecosystems 365

Table 15.1 The true mangroves and mangrove associates endemic species to Southeast Asia
No. Botanical Name Family Groups
1. Aegiceras floridum Primulaceae True mangroves
2. Avicennia eucaltyptifolia Acanthaceae True mangroves
3. Avicennia lanata Acanthaceae True mangroves
4. Azima sarmentosa Salvadoraceae Mangrove associates
5. Barringtonia conoidea Lecythidaceae Mangrove associates
6. Blumeodendron tokbrae Euphorbiaceae Mangrove associates
7. Camptostemon philippinense Malvaceae Mangrove associates
8. Croton heterocarpus Euphorbiaceae Mangrove associates
9. Fagraea crenulata Gentianaceae Mangrove associates
10. Gluta velutina Anacardiaceae Mangrove associates
11. Heritiera globosa Malvaceae Mangrove associates
12. Ilex cymosa Aquifoliaceae Mangrove associates
13. Ilex maingayi Aquifoliaceae Mangrove associates
14. Ixora timoriensis Rubiaceae Mangrove associates
15. Podocarpus polystachyus Podocarpaceae Mangrove associates
16. Ochthocharis borneensis Melastomataceae Mangrove associates
17. Quassia harmandiana Simaroubaceae Mangrove associates
18. Rapanea porteriana Myrsinaceae Mangrove associates
19. Scolopia macrophylla Flacourtiaceae Mangrove associates
20. Serianthes grandiflora Leguminosae Mangrove associates
21. Sindora siamensis Leguminosae Mangrove associates
22. Symplocos celastrifolia Symplocaceae Mangrove associates
Source: Suratman (2008)

countries, followed by Malaysia, with 42 species. The 35 mangrove species occur-


ring in Southeast Asia are reported as uncommon or rare. Around 18 percent
(51 species) of mangrove flora of Southeast Asia are endemic to the region, and
includes 22 trees and shrubs, 13 epiphytes, eight ferns, four palms and climbers,
respectively. Table 15.1 shows the list of true mangroves and mangrove associates
species that are endemic in Southeast Asia. According to Tomlinson (1986), true
mangroves refer to as plants species that occur only in mangrove forests, play a
major role of the mangrove community, are not found in the terrestrial communities
and have morphological specializations to the mangrove environment (i.e. aerial
roots, vivipary) and some mechanism for salt exclusion. In contrast, the group of
species that belong to the mangrove associates do not the criteria of true mangroves
specified by Tomlinson above. These plants are mostly representing
non-herbaceous, sub-woody and climbers that are found mostly in the regions
adjoining the tidal periphery of mangrove forests. Most of them are either naturally
or accidentally dispersed from next the adjacent forest types such as beach forests
and lowland forests.
Malaysia has approximately 645,852 ha of mangroves which is the third largest in
the Asia Pacific region. Peninsular Malaysia consists of 38 exclusive and
366 N. N. Md Isa and M. N. Suratman

57 non-exclusive and associate mangrove species (Rozainah and Mohamad 2006).


Mangrove species from Rhizophora’s genus is abundantly found throughout the
Peninsula Malaysia and the Borneo island of Malaysia (Mojiol et al. 2019).
Rapid development in Singapore has affected the adjacent wetland area of the
island. The areas have been destroyed and degraded due to mangrove area conver-
sion into land development. The mangroves have long been reported to be
undervalued for their utilitarian and intrinsic values. For example, only about 1%
(6 km2 out of 647.8 km2) mangrove area remains from the original area of 13%
(Hsiang 2000). Findings from the various studies on the present extent of mangrove
forests show inconsistencies. For example, Hsiang (2000) reported that seven
mangrove species were extinct in Singapore in the last half-century. In another
study, Hsiang (2000) through his personal communication with Hugh Tan of
National University of Singapore reported that only four mangrove species, i.e.,
Barringtonia conoidea, Ochthocharis borneensis, Bruguiera sexangular and
Brownlowia argentata, were extinct. However, Shufen et al. (2012) reported only
one species extinct, which is B. argentata whereas a total 35 true mangrove species
can still be found in Singapore.
In Brunei, plant species composition in the mangroves was monitored using the
Advanced Land Observation Satellite (ALOS). From the observation, it was reported
that the total extent of mangrove cover was found to be 35,183.74 ha, of which
Weston and Menumbok areas occupied more than two-folds (58%), followed by
Sundar (27%) and Limbang (15%). The medium resolution of ALOS data was
efficient for mapping dominant mangrove species such as Nypa fruticans,
R. apiculata, S. caseolaris, S. alba and Xylocarpus granatum in the vicinity with
classification accuracy of 80% (Satyanarayana et al. 2018).
In Thailand, mangrove forests occur on the muddy tidal flats at the river mouths
and along the coast of southern and eastern parts of the country, both on the
extensive coasts along the Gulf of Thailand and on the Andaman Sea one, where
they are most heavily concentrated. Large mangrove stands are found along the
Chao Phraya delta. Mangroves form a two-storeyed forest, with an upper layer
generally growing up to 20 m in height, dominated by Rhizophora apiculata and,
to a lesser extent, R. mucronata, Heritiera littoralis and Xylocarpus mekongensis.
Common species of the lower layer are Bruguiera cylindrica, B. parviflora,
B. sexangula, Ceriops decandra and C. tagal. B. gymnorrhiza is a common emer-
gent of up to 40 m in height and 2 m in girth (FAO 2005). Landwards, where mud
has accumulated, dryer soils are overgrown with ferns and herbs and can give way to
evergreen forest. On the edge of stream N. fruticans is common.
Indonesia is one of the countries with an extensive mangrove forests in the world.
In addition, Indonesia has a very high level of mangrove diversity. In Indonesia,
N. fruticans also displays high density structure in Sembilang and Bungin Rivers in
Sumatra. Twelve species of mangroves also found in both areas (Sarno et al. 2015).
Indonesia comprised largest mangrove areas with approximately 31,890 km2 in
Southeast Asia. In Kumbewaha, Buton Island in Indonesia, 20 species and 17 man-
grove tribes are commonly found. In terms of plant diversity, R. mucronata recorded
the highest diversity value while the lowest is the S. ovata (Iksan et al. 2019).
15 Structure and Diversity of Plants in Mangrove Ecosystems 367

Panda et al. (2017) reported that in Bhitarkanika National Parks, India a total
29 true mangrove species and 72 associate species were found. The recorded true
mangroves belong to 11 families and 15 genera, and the associates were recorded
from 39 families and 56 genera. Among the true mangrove families, Rhizophoraceae
showed maximum richness both at species and generic levels with 10 true mangrove
species.

15.5 Threat to Mangrove Forests

Despite the important socio-economic and environmental roles offered by mangrove


forests, they are facing the same threats as terrestrial forests. Recently, there have
been an increased demand of land for development, settlement and aquaculture
activities by the surrounding communities that have resulted in overharvesting
mangrove poles for piling that has made mangroves becoming plant in peril and
degrading at an alarming rate. The conversion of mangroves has led to the destruc-
tion which is believed to be exceeding the terrestrial forest such as lowland and hill
forests. The destruction of mangrove forests could be a catastrophic disaster espe-
cially to the local community livelihoods that rely on mangrove resources. If not
handled wisely, the mangrove destruction can create a new source of carbon in the
atmospheres.
From the past decade, mangroves received constant threats from the global
climate change factors. The rapid anthropogenic activities coupled with sea level
rises, storm and tsunami seem to be the new menace that could be the catalyst for the
destruction of mangroves worldwide. The ad-hoc changes due to anthropogenic
disturbance have altered the mangrove habitats entirely. The relevant stakeholders
and whole communities need to make sure that these phenomena need to address and
given serious attention to ensure these majestic ecosystems survive.
Changes in mangroves have been reported in many studies throughout the world.
For instance, in Malaysia, Khuzaimah et al. (2013) in their study at Merbok
mangrove forests indicated that during the period of 2000–2010 found that
mangroves have experienced loss about 1246 ha due to an intense aquaculture
farming, land development and agriculture sector. Another mangroves study
conducted by Suratman and Ahmad (2012) in Pulau Indah, Selangor using Landsat
images from 1995–1999 and 1999–2005 found that the net reduction during the two
intervals were 2.73% and 0.24%, per annum, respectively. The study concluded that
the main reasons for the decreased of were caused by rapid land development (port
and ship terminals) and land settlement. Meanwhile, a study that was conducted in
Johor, Malaysia using the remote sensing approach found out that between 1989 and
2014, the estimated loss of mangroves in the area was about 6030 ha (Kanniah et al.
2015). The underlying causes reported were uncontrollable land development,
aquaculture activities, intensified erosion and urbanization.
Mangrove is the host with a wide variety of biodiversity, providing habitats for
flora and fauna, including aquatic and terrestrial insects, fish, crustacean, mamma-
lian, amphibian, reptilian and avian species (Thomas et al. 2017). Although
368 N. N. Md Isa and M. N. Suratman

mangroves contribute many returns to other living creatures, this ecosystem, the
ecosystems also face their own threats and degradation that affect the socio-
economic livelihood fulfillment. Therefore, understanding and awareness on the
mangrove structure and diversity may provide insights for this precious ecosystem
to be protected.

15.6 Conclusion

The rich mangrove ecosystems provide support to the people for their livelihoods
and planet against global warming in the unique ways by preserving habitats for flora
and fauna and providing protection against storm and flooding. Protecting mangrove
forests not only help local communities that are depending on their resources, but
also helps in providing the breeding ground for marine life, carbon storage and
preserving mangrove plant biodiversity. Understanding about structure and species
diversity in mangroves is crucial for management and conservation purposes. The
mangrove plant structure has a significant effect on the ecosystem functions. In
addition, understanding structural attributes of mangrove vegetation is important
with respect to their productivity. Moreover, planning any species conservation
initiatives and scientific management of mangroves requires an understanding of
their community structure, species, diversity and composition. Therefore, the need to
emphasize the documentation of this information is fundamental for long-term
management and sustainability of these magnificent yet degraded ecosystems.
Considering mangroves to be one of the rich habitats with an endemic species,
while offering so many benefits, both ecological and socio-economic, there is an
urgent need to implement conservation efforts to control their losses. Uncontrolled
anthropogenic activities in the mangrove forests will make them as a threatened
ecosystem in the world.

References
Abou Seedo K, Abido MS, Salih A, Abahussain A (2017) Structure and Composition of Mangrove
Associations in Tubli Bay of Bahrain as Affected by Municipal Wastewater Discharge and
Anthropogenic Sedimentation. Int J Biodivers Conserv 2017:1–9.
Aksornkoae S (1995) Ecology and management of mangrove restoration and regeneration in East
and Southeast Asia. Proceedings of the Ecotone IV, 18, 22.
Deltares (2014) Habitat requirements for mangroves. Available at: https://publicwiki.deltares.nl/
display/BWN/Building+Block+Habitat+requirements+for+mangroves
Ellison AM, Mukherjee BB, Karim A (2000) Testing patterns of zonation in mangroves: Scale
dependence and environmental correlates in the Sundarbans of Bangladesh. J Ecol 5: 813–824.
FAO (2005) Global Forest Resources Assessment 2005: Thematic Study on Mangroves. Forest
Resources Development Service Forest Resources Division Forestry, Forestry Department
FAO, Rome (Italy), p 10.
Faridah-Hanum I, Kudus KA, Saari NS (2012) Plant diversity and biomass of marudu bay
mangroves in Malaysia. Pak J Bot 44: 151–156.
15 Structure and Diversity of Plants in Mangrove Ecosystems 369

Gandhi S, Jones TG (2019) Identifying mangrove deforestation hotspots in South Asia, Southeast
Asia and Asia-Pacific. Remote Sens 728: 1–27.
Giesen W, Wulffraat S, Zieren M, Scholten L (2006) Mangrove Guidebook for South East Asia.
FAO and Wetlands International.
Hogarth PJ (2007) The Biology of Mangroves and Seagrasses. Oxford University Press, London.
Hsiang LL (2000) Mangrove conservation in Singapore: A physical or a psychological impossibil-
ity? Biodivers Conserv 3: 309–332.
Iksan M, Aba L, Taharu FI, Alfian A, Ardyati DPI, Jumiati, Alzarliani WOD, Hardin, Larekeng SH
(2019) The diversity of mangrove forests in Kumbewaha, Buton Island, Indonesia. IOP Con-
ference Series: Earth and Environmental Science 1:1–6.
Joshi HG, Ghose M (2014) Community structure, species diversity, and aboveground biomass of
the Sundarbans mangrove swamps. Tropic Ecol 3: 283–303.
Kanniah KD, Sheikhi A, Cracknell AP, Goh HC, Tan KP, Ho CS, Rasli FN (2015) Satellite images
for monitoring mangrove cover changes in a fast growing economic region in southern
Peninsular Malaysia. Remote Sens 11: 14360–14385.
Khuzaimah Z, Ismail MH, Mansor S (2013) Mangrove changes analysis by remote sensing and
evaluation of ecosystem service value in Sungai Merbok’s mangrove forest reserve, Peninsular
Malaysia. In Lecture Notes in Computer Science, Proceedings of the 13th International Confer-
ence in Computational Science and Its Applications–ICCSA. Springer, Berlin, Germany,
p 611–622.
Mojiol AR, Wan Maisyahirah MMN, Muhammad Ali SH (2019). Tree Species Diversity of
Mangrove at Tunku Abdul Rahman Park, Sabah Malaysia. Acta sci Agric 9: 181–187.
Ong JE, Gong WK (eds) (2013) Structure, Function and Management of Mangrove Ecosystems.
International Society for Mangrove Ecosystems (ISME) Book Series No. 2, Okinawa, Japan,
and International Tropical Timber Organization (ITTO), Yokohama, Japan.
Panda M, Murthy T, Samal R, Lele N, Patnaik A, Chand P (2017) Diversity of True and Mangrove
Associates of Bhitarkanika National Park (Odisha), India. Int J Adv Res 1: 1784–1798.
Primavera JH, Garcia LMB, Castanos MT, Surtida MB (eds) (2000) Mangroves of Southeast Asia.
Mangrove-Friendly Aquaculture: Proceedings of the Workshop on Mangrove-Friendly Aqua-
culture. 12–14 October 1999, Iliolo City, Philippines, p 231–239.
Rozainah MZ, Mohamad MR (2006) Mangrove Forest Species Composition and Density in Balok
River, Pahang, Malaysia. Ecoprint: An International Journal of Ecology 13:23–28.
Sarno, Suwignyo RA, Zulkifli D, Munandar, Ridho MR (2015) Primary Mangrove Forest Structure
and Biodiversity. Int J Agri Sys 2:135–141.
Satyanarayana B, Muslim AM, Horsali NAI, Zauki NAM, Otero V, Nadzri MI, Ibrahim S, Husain
ML, Dahdouh-Guebas F (2018) Status of the undisturbed mangroves at Brunei Bay, East
Malaysia: A preliminary assessment based on remote sensing and ground-truth observations.
PeerJ 6:e4397 https://doi.org/10.7717/peerj.4397.
Shufen Y, Lim RLF, Chiou-rong S, Yong JWH (2012) The Current Status of Mangrove Forests in
Singapore. Proceedings of Nature Society, Singapore’s Conference on ‘Nature Conservation for
a Sustainable Singapore,’ October 2011, p 99–120.
Suratman MN, Ahmad S (2012) Multi temporal Landsat TM for monitoring mangrove changes in
Pulau Indah, Malaysia. In Proceedings of the 2012 IEEE Business, Engineering and Industrial
Applications (ISBEIA) Symposium, Bandung, Indonesia, 23–26 September 2012, p 163–168.
Suratman MN (2008) Carbon sequestration potential of mangroves in southeast Asia Managing
forest ecosystems: The challenge of climate change. Springer, p 297–315.
Thomas N, Lucas R, Bunting P, Hardy A, Rosenqvist A, Simard M (2017) Distribution and drivers
of global mangrove forest change, 1996–2010. PLoS ONE 6: 1–14.
Tomlinson PB (1986) The botany of mangroves. J Tropic Ecol 2: 188–189.
Wang L, Mu M, Li X, Lin P, Wang W (2011) Differentiation between true mangroves and
mangrove associates based on leaf traits and salt contents. J Plant Ecol 4: 292–301.
Livelihood of Forest Dependent Dwellers
in Relation to the Exploitation of Resources 16
at the Fringe of Indian Sundarban

Chandan Surabhi Das

Abstract

Sundarban, a single largest mangroves block in the world, shared by India and
Bangladesh, has a rich biodiversity that provides staggering ecosystem services to
local inhabitants for livelihood. Poor agriculture in one way and rich
bio-resources in other have compelled local inhabitants to depend upon forest
resources from time immortal. Mangroves are diverse and highly productive
ecological communities, but the pristine Sundarban forest had been dwindled
during the last two centuries because of committed destruction. The present study
assesses the status of coastal communities who exclusively depend on the
Sundarban forest for subsistence and livelihoods. For identifying the livelihood
pattern of these communities a household survey (n ¼ 1079) and in-depth
interviewing (n ¼ 157) were done in three villages of the Sundarban, such as
Rajatjubille in Gosaba, Samsernagar in Hingalganj, and Kisorimohanpur in
Kultali block. A Relative Livelihood Index (RLI) was developed among major
occupant groups (n ¼ 14), for assessing the degree of strength among primary
occupations. The strength of fishing and crab collection as a major primary
occupation is observed in Samsernagar and Rajatjubille, whereas in
Kisorimohanpur agricultural labour dominates others. According to RLI, the
primary occupation like honey collection, fishing, and carpentry scores negatives
which justify these occupations were alone robust to support the occupants
continuing livelihood. So the category of forest users and the type of resource
collected from the forest should be considered as the main criteria for designing
any conservation plan as well as a sustainable livelihood programme in
Sundarban.

C. S. Das (*)
Department of Geography, Barasat Government College, Kolkata, India

# The Author(s), under exclusive license to Springer Nature Singapore Pte 371
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_16
372 C. S. Das

Keywords

Mangroves forest · Coastal communities · Primary occupation · Relative


livelihood index · Sustainable livelihood

16.1 Introduction

According to the British Department for International Development (DFID 2000)


livelihood comprises the capabilities, assets (including both material and social
resources), and activities required for a means of living. A livelihood is sustainable
when it can cope with and recover from stresses and shocks, and maintains or
enhances its capacities and assets both now and in the future, while not undermining
the natural resource base (Ashley and Carney 1999). A sustainable and vibrant
livelihood system helps to build up ‘layers of resilience’ to overcome ‘waves of
adversity’, and enables people to transform multiple adversities into opportunities
(Glavovic et al. 2003; Chambers and Conway 1992). In the mangrove forest, the
term ‘livelihood’ refers to the capabilities and assets that the households have at their
disposal to cope with the change caused by the conservation policy and strategy
within the framework (Darwin 2014). Mangrove is a shrub or small tree that grows in
tidal coastal or brackish saline water of the tropical world. Chapman (1976) used the
term ‘mangrove’ as intertidal forests or plant communities or ‘mangal’. The term
‘mangal’ was also commonly used in French and in Portuguese to refer to both forest
communities and individual plants. They, sometimes, have been described as
‘coastal woodlands’, ‘mangals’, ‘tidal forests’, or ‘mangroves’ or ‘mangrove forests’
(Saenger 2002). So, in general, mangroves are salt-tolerant plant species that strug-
gle every moment against tides as they straddle at the interface zone of the land and
the sea. Mangroves are uniquely adapted to the intertidal zone in the tropics and
sub-tropics ranging from mean sea level to high tide level, or in zones where
moderate salinity, average temperature, and high rainfall are experienced (Alongi
2002). Under this environment, mangroves extend in more than 118 countries in the
tropics, sub-tropics, and temperate regions covering, respectively, 24 million
hectares (Twilley et al. 1992), 10 million hectares (Bunt et al. 1992), and 14–15
million hectares (Schwamborn and Saint-Paul 1996). However, the mangroves are
confined to tropical or sub-tropical climates where the average monthly minimum
temperature is 20  C (Chapman 1976).
Sundarban is the largest patch of mangrove forest found anywhere in the world.
The vegetation of the Sundarban mangrove forest appeared during
31,750  2030 years before the past (BP) (Chaudhuri and Choudhury 1994). But
the origin of Bengal basin wherein present-day Sundarban mangrove forests exist
started as early as 126 million years BP (Naskar and Mandal 1999). Sundarban
mangrove forests have been painstaking as one of the seven most significant
wetlands globally, based on biological diversity (Junk et al. 2006). Importantly,
Indian Sundarban alone represents 30 true mangroves out of 40 true mangroves
reported in the Old World Tropics. In National perspective, Indian Sundarban
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 373

mangrove covers 62% area (about 2400 km2), which maintains 90% species diver-
sity, followed by Andaman and Nicobar Island with 76.5% and Bhitarkanika, Orissa
with 72.3% in respect of existing major mangroves diversity (Mandal and Naskar
2008). This densely populated reclaimed region, on the other hand, is one of the
immature low lands of the world with a multitude of socio-economic problems (Das
and Bandyopadhyay 2012). About 95% of the population in Sundarban is directly or
indirectly dependent on water bodies for their daily activities (agriculture, aquacul-
ture, and fishery) (Chowdhury et al. 2017). As a result, the virgin mangrove forest of
Sundarban was destructed for meeting the needs of the human population for the last
two hundred years (Mandal et al. 2010). Claude Russel, the British administrator in
the Bengal province of India, initiated to reduce mangrove forest areas for the
purpose of reclamation for human inhabitation since 1770 (Chaudhuri and
Choudhury 1994). Most of the labours who were engaged in the reclamation
activities were not sent back to their earlier homeland and rather rehabilitated in
the forest cleared areas. The reclaimed forest areas were utilized for agriculture,
fishing, and allied activities pertaining to livelihood purposes. Until 1971, the
reclamation of the forest continued (Chaudhuri and Choudhury 1994). Since then
the need for conservation of mangroves awoke, albeit lately. Nonetheless, degrada-
tion of the forest was not stopped, rather perpetuated slowly by illegal encroachment
because of the dense population (Mandal et al. 2010).
People in the fringe zone of Sundarban are basically forest dwellers who tradi-
tionally could spend their livelihood through a collection of resources emanating
from the forest. Since the Sundarban mangrove forest is full with varieties of
resources which are available easily from land and water through minimal efforts
spent. These resources comprise three items: (i) food, fodder, honey, leafy vegetable
(ii) Tannin, wax, wood, thatching materials, timber for construction of the house,
boat, fence, etc., and (iii) fuel wood. Most of the habitants in this region belong to SC
(Scheduled Caste), ST (Scheduled Tribe), and OBC (Other Backward Class)
communities accounting about 89% of the total population of Sundarban (Das and
Mandal 2016). They are, by and large, poor, down trodden, diffident, and have a
relatively least amount of inherent properties and wealth of economic benefit. Most
of them are landless daily wage labours or marginal labours. In this perspective, the
present study was intended to assess the patterns of livelihood of dwellers in relation
with exploitation of the forest’s resources as well as to analyse the present scenario
of the collection of NTFPs for sustainable livelihood.

16.2 Study Area

The Sundarban, the world’s largest continuous patch of mangrove forest, is formed
at the delta of the Ganges, Brahmaputra, and Meghna rivers on the Bay of Bengal.
The total area of the entire Sundarban is about 10,000 sq. km, 62% of which is found
in Bangladesh and the rest, 38%, lies in the southeast of West Bengal, India
(Spalding et al. 2010; Das 2012). The UNESCO Man and the Biosphere programme
in 1989 declared 9630 km2 of the western part of the Sundarban region as the
374 C. S. Das

Fig. 16.1 Phase wise reclamation and loss of forest in Sundarban

Sundarban Biosphere Reserve (SBR), out of which about 5364 km2 (c.56 percent) of
the northwestern part have been cleared or reclaimed for human habitation, devel-
opment of agriculture fields, and brackish water fisheries (Naskar and Guha Bakshi
1987) (Fig. 16.1). The northern boundary of the Sundarban is demarcated by the
Dampier Hodges line of 1829–1830 (Danda et al. 2011; World Bank 2014). The
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 375

Table 16.1 Legal and Institutional overview of Indian Sundarban


Area
Legal designation Year (sq. km) Activities prohibited/regulated
Sundarban Tiger 1973 2584.89 Divided into two parts—Core (1699.62 km2) and
Reserve (STR) buffer area (885.27 km2). The core area is prohibited
from human interference, fishing and tourism
activities allowed in buffer areas
Sajnakhali wildlife 1976 362.40 Part of the buffer area of the STR but fishing and
sanctuary hunting prohibited
Lothian wildlife 1976 38 Part of the buffer area of SBR. Fishing and hunting
sanctuary prohibited
Haliday wildlife 1976 6 Part of the buffer area of SBR. Fishing and hunting
sanctuary prohibited
Sundarban National 1984 1330 Part of the buffer area of STR. Fishing and hunting
Park activities prohibited
Sundarban 1989 9630 Total Sundarban (STR + SBR) divided into 3 parts—
Biosphere Reserve Core area (prohibited: 1692 km2), buffer area
(SBR) (restricted: 2233 km2), and transition zone
(inhabitated: 5705 km2).
Critical tiger habitat 2007 1699.62 Rename of the STR Core area where all
anthropogenic activities prohibited
Source: Ghosh (2014), Sahana and Sajjad (2019), STR (2014)

reclaimed Sundarban now supports a fast rising population of 4.42 million with an
average density of 1074 person km2 and a growth rate of 18.23 persons yr1
(Census 2011).
4226 km2 of the southeastern part of the Sundarban was included under the
Sundarban Reserve Forests in 1911. Out of these, 2585.89 km2 of forests and tidal
waterways of the extreme southeastern part of the Sundarban were declared as the
Sundarban Tiger Reserve (STR) in December 1973 to promote a major national
project for tiger conservation, the Project Tiger. The core area of the STR, measuring
1330 km2 was declared as National Park on 4 May 1984. This core area (1700 km2)
has been renamed to Critical Tiger habitat in 2007 (Table 16.1). The STR was
included in the list of UNESCO World Heritage Site in 1987.

16.3 Productivity of the Mangroves

Mangroves are diverse and highly productive ecological communities (Mandal et al.
2009; McDowell 1995) having important ecosystem services in the tropical coastal
land of the world (Day et al. 1987). Mangroves are also a significant supplier of
nutrients in the wetland ecosystems (Kamruzzaman et al. 2017). Biomass and net
primary productivity of mangrove forest have been studied previously in different
mangrove forests across the world (Putz and Chan 1986; Day et al. 1987, 1996;
Saintilan 1997; Komiyama et al. 2000; Kamruzzaman et al. 2017). Measuring of
these components indicates the ecosystem regulation and estimate of carbon stock
376 C. S. Das

Table 16.2 Net primary production and related characteristics of major ecosystems
Net primary production
Area Normal range Mean Total (109
Ecosystem type (106 km2) (g m2 yr1) (g m2 yr1) t yr1)
Tropical rainforest 17.0 1000–3500 2200 37.4
Tropical deciduous forest 7.5 1000–2500 1600 12.0
Temperate evergreen 5.0 600–2500 1300 6.5
forest
Temperate deciduous 7.0 600–2500 1200 8.4
forest
Boreal forest 12.0 400–2000 800 9.6
Woodland and scrubland 8.5 250–1200 700 6.0
Savanna 15.0 200–2000 900 13.5
Temperate grassland 9.0 200–1500 600 5.4
Tundra and alpine 8.0 10–400 140 1.1
Desert and semi-desert 18.0 10–250 90 1.6
scrub
Extreme desert: rock, 24.0 0–10 3 0.07
sand, and ice
Cultivated land 14.0 100–4000 650 9.1
Swamp and marsh 2.0 800–6000 3000 6.0
Lake and stream 2.0 100–1500 400 0.8
Open ocean 332.0 2–400 125 41.5
Upwelling zones 0.4 400–1000 500 0.2
Continental shelf 26.6 200–600 360 9.6
Algal beds and reefs 0.6 500–4000 2500 1.6
Estuaries (excluding 1.4 200–4000 1500 2.1
marsh)
Source: Whittaker and Likens (1975)

(Tamai et al. 1986; Komiyama et al. 1987, 2000) is significant in mangroves


compared to other varieties. Naskar and Ghosh (1989) opined that the mangrove
ecosystem is the most productive ecosystem of the world. The biological productiv-
ity of an ecosystem may be defined as the rate of appearance of matter as living tissue
(Simmons 1981). As shown by Whittaker and Likens (1975), swamps and marshes
occupy the top position among the world’s major ecosystem types in terms of mean
net primary production of dry matter. This amounts to 3000 g m–2 yr–1 in swamp and
marsh ecosystems compared to 2200 g m–2 yr–1 of tropical rain forests, 1600 g m–
2
yr–1 of tropical seasonal forests, and 1300 g m–2 yr–1 of temperate evergreen forests
(Table 16.2). It was estimated in the mangroves of Southeast Asia that rates of
biomass accumulation range between 6.3 and 45.4 t ha–1 yr1 (Hogarth 1999). The
available literature on the productivity of mangroves also indicate a wide range of
values between 2 and 16 m3 ha1 yr1 of mean wood increments (Tomlinson 1986).
The mean biomass increment and mean litterfall were 7.1 and 10.1 Mg ha1 yr1,
respectively, for the mangrove community in Bangladesh (Kamruzzaman et al.
2017).
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 377

Litter fall from mangrove vegetation, decomposition of litter and thus nutrient
release into soil and water, all are cumulative effects of ecosystem services, which
steadily reduce with the gradual loss of vegetation. The mangrove ecosystem alone is
rather productive as measured to yielding as 350–500 g C/m2/yr (Mann 1982) that
provide a considerable contribution to the food chain that leads to keep the
sustainability of coastal fisheries (Upadhyay et al. 2002). The Avicennia spp. and
Sonneratia spp. dominated forest area of Indian Sundarban has been assessed to
produce about 212 tons/ha of biomass (Chakraborty 1985). It is estimated that a full-
grown mangrove stand of 10 years old may add to soil the following amounts of
nutrients through litter decomposition: N, 46.6 kg/ha/yr; K, 25.6 kg/ha/yr; Ca,
99.3 kg/ha/yr; Mg, 34.1 kg/ha/yr; and Na, 31.8 kg/ha/yr (Gong et al. 1984).

16.4 Reclamation and Livelihood in Sundarban

The pristine Sundarban forest had been dwindled during the last two centuries
because of committed destruction. Degradation of Sundarban mangrove forests
started since 1770 during British India. Claude Russel and later Tilmen Henckell
started to reclaim Sundarban forests, which were then partly exploited for rehabili-
tation of the human population and remaining for rice cultivation and brackish water
fisheries (Mandal et al. 2010). Settlement in the Sundarban included migrant
populations of adjoining Midnapore district and from the district of Jharkhand
(then Bihar) and Chhattisgarh (then Madhya Pradesh), who came in search of
work and land (Danda 2007). Reclamation as well as settlements occurred in five
phases (Fig. 16.1). At the end of the eighteenth century, the mangrove forest
extended up to Kolkata (Ghosh et al. 2015); and at the time of independence in
1947 the forest was only 50% of its pre-colonial size (Giri et al. 2007). In 1947 India
witnessed the destiny of her partition. As a result, an enormous incursion of refugees
was settled in reclaimed Sundarban (Naskar and Guha Bakshi 1987) until 1971.
Since then the need for conservation of mangrove awoke, albeit late. Nonetheless,
degradation of the forest was not stopped, rather perpetuated slowly by illegal
encroachment because of the dense population (Mandal et al. 2010).
The human population in the Sundarban increased rapidly in the post-colonial
era, especially following the partition of India and Bangladesh. Since the 1970s, the
Indian Sundarban mangroves have been protected under various legal measures
which were established primarily to protect and help increase the threatened tiger
population. After independence, the population of Indian Sundarban grew from 1.15
million in 1951 to 4.42 million in 2011, which led to an increase of 343% in last
60 years. The population was increased almost 15 times from 1872 to 2011. The
decadal population growth rate of Sundarban is 18.23% (2001 to 2011). From 1901
to 1951 the population was increased almost double. After independence from 1951
to 2011, the population increased almost 4 times due to political phenomena. A mass
of the population has migrated from Bangladesh to the Sundarban. As a result,
mangrove forests had been decreased by 20.58% on account of the conversion of
378 C. S. Das

forest to agricultural land and settlements in the last 70 years (1968 to 2014) (Ghosh
et al. 2015).
Salinity of the Sundarban is another concern. Embankment construction together
with forest clearing made extensive human habitation possible in the Sundarban
(Sánchez-Triana et al. 2018). The embankment construction process started in the
late nineteenth century and continued through to the twentieth century. It made
reclamation possible by preventing saline water from inundating land that was
otherwise suitable for cultivation (Sánchez-Triana et al. 2014).

16.5 Poor Land Utilization

Sundarban, having 56% landless family, is considered to be one of the most


backward regions in India (Singh et al. 2010). Historically, the region was
dominated by a single crop farming system with the local Amon rice which was
cropped in the cropping pattern: ‘fallow-rice-fallow’ (BARC 1998). But in the early
1980s brackish water shrimp farming came out as a vital land use leading to the
emergence of a ‘fallow-shrimp-rice’ (Miah et al. 2003) farming. But all the rivers,
creeks, and canals of the region have saline water, and therefore unsuitable for
agriculture. Developing tube well irrigation proved difficult as the underground
water up to 20–25 ft. from the surface is saline (German Agro Action Plan 2009).
Besides, regular occurrence of high tides twice a day induces salinity in the water of
the canals. In the Indian Sundarban, around 56% of land mass lies within the coastal
low-lying ecosystems with an elevation of <5 m above mean sea level (Mandal et al.
2019b). Some parts are even below the mean sea level. All these factors force the
region into a monoculture as well as poor agricultural productivity. As such, every
year more and more agricultural lands were converted into more profitable shrimp
cultivation resulting into gradual declining rice cultivation. This situation urges the
landless and marginal people of the Sundarban to move frequently into the forests in
search of livelihood.

16.5.1 Agriculture

The main economic activity in the Sundarban, rain-fed paddy agriculture, is made
possible by the construction of earthen embankments to keep brackish tidal water at
bay. Limited irrigation facilities lead to mono-cropping agriculture in Sundarban
which are also constrained by the salinity of both ground water and surface water.
Historically, in the Sundarban a variety of salt-tolerant paddy were cultivated on
raised sections of the islands without embankments. However, according to the
National Bureau of Plant Genetic Resources (NBPGR) only two varieties of
them—Matla and Hamilton are cultivated now; others are believed to have been
lost under the onslaught of green revolution (Ghosh 2010).
Agriculture in the reclaimed areas is also constrained by excess water during the
rainy monsoon period (kharif) accompanied by two other phenomena like high
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 379

humidity and less sunshine hours. High humidity lures pests and diseases, whereas
low sunshine hours limit plant growth. The main kharif crop is the rice of a
traditional variety, which is able to survive deep water-logging. Basically, farmers
have no choice for alternate crops other than tall traditional rice varieties in kharif
(Sarangi et al. 2015). During the dry winter months acute shortage of irrigation water
along with an increase in soil and water salinity limits the agricultural production
(Mandal et al. 2019a). The dominant soluble salts identified in the study region were
NaCl and MgSO4 (Mandal et al. 2019b). As a result, during Rabi season, most of the
land is left fallow and is used to graze cattle, but the little amount of areas are
irrigated from village ponds stored with monsoon water. Paucity of sweet water in
the region poses a major problem for irrigation (Pakrashi 2016). Only a limited net
cropped area is irrigated in Sundarban. Sources of irrigation in the Sundarban are
primarily tanks (71%) and shallow tube-wells (13%), with the former method being
the preferred choice (World Bank 2014). The cropping intensity was low (114%)
and more than 80 per cent of the farm lands remained fallow during the Rabi season
(Mandal et al. 2017).
Because of non-remunerative agriculture and steady growth in the fishery sector,
many agricultural lands are getting converted to high output-intensive brackish water
farming (Mandal et al. 2019b) during last three or four decades, which poses severe
environmental threats to the region. According to Hazra and Samanta (2016)
because of increasing demands of shrimp cultivation, agricultural land of Indian
Sundarban has been reduced from 2149 km2 in 2001 to 1691 km2 in 2008.

16.5.2 Non-Timber Forest Products/Non-Wood Forest Products

Because of the extensive and diverse resources available in Sundarban, the forest
generates large-scale livelihood opportunities. Forest resources are christened as
NTFP (non-timber forest product) and NWFP (Non-wood forest products). A
biological product collected from forested area, including fish, crab, prawn seed,
tiger prawn, firewood, construction wood, thatching leaves, all these constitute a
major share of NTFP, with honey and bees wax being part of NWFP (Shackleton and
Shackleton 2004). In general, the term Non-Timber Forest Products (NTFPs) covers
all tangible products of forest origin, except wood (Ros-Tonen et al. 1995). Some-
times, these resources include parts of plants, fungi, and other biological materials or
natural resources collected from forests apart from sawn timber. NTFPs may provide
local job opportunities and contribute significantly to rural people as a source of
income and facilitate the subsistence living in the fringe areas of Sundarban (Peters
et al. 1989; Hegde et al. 1996).

16.5.2.1 Collection of Plant Resources


Utilization and exploitation of mangroves can take many forms (Tomlinson 1986)
and most are present in the Indian Sundarban. The mangrove trees are used as swan
timber, poles, fuel, and pulp wood. Tannins and dyes are also extracted from them.
Some species are used as thatching material. Leaves of Nypa fruticans (‘Golpata’) is
380 C. S. Das

a major source of thatching material, which is extensively used by the poorer section
of the rural population in Sundarban. Resources like tannin, wood, timber, etc., will
amount to a staggering figure in terms of pecuniary benefit, but local people use
these resources as their integral part of daily life without economic botheration.
Mangrove flora also provides important medicinal benefits: leaves of Bruguiera
gymnorrhiza are used for remedy of diarrhoea and blood pressure, Rhizophora
mucronata for angina, Acanthus ilicifolius for asthma and rheumatism, Lumnitzera
racemosa for herpes and itches, Cynometra ramiflora and Excoecaria agallocha for
leprosy (Kothari and Rao 1995; Untawale 2006; Rodrigues 2006; Mandal and Bar
2018). Nypa fruticans, commonly known as the nipa palm (or simply nipa) or
mangrove palm (Nishat 2019). The fruits of Sonneratia apetala are widely used as
food and in treating various diseases like asthma, febrifuge, ulcers, swellings,
sprains, bleeding, haemorrhages, and piles (Bandaranayake 1998; Hossain et al.
2016). The fruits of Sonneratia apetala are marketed now and also preferable food
items to the Rhesus monkey (Sanyal 1992). Frond of Acrostichum aureum is used as
leafy vegetables (Pillai and Ong 1999; Badhsheeba and Vadivel 2020). Most of the
mangrove leaves are suitable fodder for the domestic livestock, apart from wild
animals like deer, monkey, and wild boar. Heritiera fomes is the principal timber
species in the Sundarban (Dasgupta et al. 2017; Khan et al. 2020). However,
collection of plant resources are now restricted in Sundarban. Earlier, Golpata
(Nypa sp.) and Hental (Phoenix sp.) which were collected by the fringe villagers
were discontinued in 1978 and 1991, respectively. The coupe operation has been
discontinued since the year 2001 (STR 2018).

16.5.2.2 Fuelwood Collection


After food, fuel occupied an important place for daily livelihood. The two major fuel
wood species in the Sundarban are Heritiera fomes and Ceriops decandra. However,
there are a number of other species, having good quality fuelwood include Amoora
cucullata, Aegiceras majus, Rhizophora mucronata, Hibiscus tiliaceus, Ceriops
candolleana, and Cynometra ramiflora (Islam et al. 2019a, 2019b). 75% of the
total households of Sundarban were dependent on the fuel wood exclusively for
energy source or in combination with other forms of biomass (Das and Mandal
2016). As a result, a large number of people are engaged in collecting firewood from
the forest almost on a daily basis. Fuel wood collection is done by both solitary and
group activity. The groups usually invade the forest with small traditional boats, but
in solitary activity, people explore the riverbank of SRF and collect fuel wood from
the vicinity (Chowdhury et al. 2020). Nearly 70% households’ treat fuel wood, cow
dung, and rice straw as the major fuel sources for cooking and related activities
(Situmorang et al. 2020). Fuel wood was an energy source covering 41% of
households, accompanied by 25% of households, depending upon cow dung and
rice straw. Rural women of Sundarban generally collect shrimp fry in the river or
creeks, and at the same time they also collect fuel wood from the forest edge (Kibria
et al. 2018).
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 381

16.5.2.3 Apiculture
Apiculture is the science and practice of bee keeping which plays a valuable part in
rural livelihoods of Sundarban. Forests with a variety of flowering plants provide a
centre for beekeeping and honeybees (Hill and Webster 1995). Beekeeping
generates much more than just honey. The maintenance of biodiversity and pollina-
tion of crops are the most valuable services provided by bees. Honey is just one of
several different products that can be harvested: others are beeswax, pollen and
propolis, royal jelly and venom, and the use of bees in apitherapy, which are
medicine using bee products (Bradbear 2003, 2005). During the flower season of
mangroves the swarms of honeybee colonies with honey are developed. Honey and
the pollen in it are used as medicines, high energy food, and as a source of vitamins
and minerals.
Sundarban mangrove areas have high potential for honey production, but honey
collectors often use destructive methods of harvesting (Burgett 2000). Sundarban
harbours numerous hives of rock bees (Apis dorsata). The floristic composition of
the Sundarban is very favourable for honey production. The best quality unifloral
honey of the Sundarban is kulshi-type followed by amur, goran, and keora
(Table 16.3). Honey and Bee-wax in the Sundarban Reserve Forest are collected
by a traditional occupational group called mawallis or moulis during the months of
April–May every year. Honey collection from the Sundarban forest is a seasonal
activity occurring usually in the period of 15th April to 31st May every year. In an
average year, nearly 1000 honey collectors with a population of more than 8000
enter the forests with permits from forest department (Vyas 2012; Das and Mandal
2016; Sen and Pattanaik 2017). Generally a group comprising 10 members in one
venture may collect about 10 to 12 quintals of honey (Mandal et al. 2010) costing
Rs. 120,000 @ Rs.100.00/kg as per Govt. rate; so that one can earn Rs. 12,000 per
trip; whereas in the local market one kg of honey cost ranging between Rs. 500 and
600.
Collecting honey from the deep forest is always a risky job from man-eaters
(Kothari 2015) which occasionally result in fatalities (Table 16.4). The West Bengal
Government has taken a unique initiative of providing beekeeping with apiary boxes

Table 16.3 Important Sundarban plants, flowering times, and quality of honey
Local name Scientific name Peak season (flowering time) Quality of honey
Kulshi Aegiceras majus Feb to Mar Very high
Amur Amoora cucullata Feb to Mar High
Goran Ceriops decandra Mar to Apr Very high
Keora Sonneratia apetala Mar to Apr High
Passur Xylocarpus mekongensis Mar to Apr Very high
Sundri Heritiera fomes Apr to May Low
Gewa Excoecaria agallocha Apr to May Low
Kakra Bruguiera gymnorrhiza Apr to May Low
Baen Avicennia officinalis May to June Low
Source: Field Study and Zohora (2011)
382 C. S. Das

Table 16.4 Honey collection from STR areas of Sundarban


Honey
Collection 2010– 2011– 2012– 2013– 2014– 2015– 2016– 2017–
(STR) 2011 2012 2013 2014 2015 2016 2017 2018
Crude 14,300 18,025 24,750 20,950 47,412 33,515 19,050 15,000
honey (kg)
Persons 765 879 715 735 1155 979 604 486
involved
with permits
Injury due 5 1 5 4 0 1 5 3
to tiger
attack
Death due to 3 2 5 2 4 3 2 2
tiger attack
Source: STR (2018), GoWB (2018)

to the local people inside the protected forest camps in several places of Sundarban
(Thakur et al. 2016). It is not only safer for the villagers, but also the quality of the
collected honey is better. While collecting honey from the forest, the local squeeze
the entire beehives and in process several bees and maggots also get squished in the
honey resulting into low-quality impure honey (Ray 2000). This initiative is highly
acceptable by the villages as the pure honey collected by this new method is sold to
the forest department at a rate of Rs 600/per kg which is roughly five times higher
than the earlier attempt. It is 100% pure natural honey, often recommended by the
medical practitioners on account of its high therapeutic value (Kumar et al. 2010;
Ajibola et al. 2012).

16.5.3 Fishery and Aquaculture

Fishing has been perhaps the most common and staple livelihood options for the
people of Sundarban since their habitation of the area (Ghosh et al. 2018). Next to
agriculture, fisheries provide a distinct source of employment and income for small
and marginal farmers of Sundarban. Sundarban is considered as a gold mine for the
fishery (Ray 2000). The Sundarban water supports 208 species of fish and
crustaceans belonging to 84 families (IUCN 1994), a higher total than that of other
tropical mangroves (Robertson et al. 1992). According to a survey, 312 bony fishes
restricted to 214 Genera, 71 Families, and 18 Orders especially are found in marine
and estuarine waters of the Sundarban mangroves (Kar et al. 2017). During the
agricultural lean season, people resort to fishing and the collection of prawn seeds,
even risking their lives from man-eating tigers and crocodiles. The Sundarban is a
rich fishing ground. The fishing nets are still often knotted by hand, the weirs plaited
manually. Fishing is still handwork, yet overfishing increasingly becomes a chal-
lenge. A large section of people in and around Sundarban earn their livelihood from
fishing and pisciculture. But fish cultivation is not carried on the scientific method.
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 383

The shallow water, creeks, small and big rivers crossing mangrove forests support
many species of fish. Over 120 species of fish are caught routinely by commercial
fishermen (Hasan and Naser 2016; Chatterjee 2017). Some species such as Hilsa
ilisha are exclusively marine in nature, but travel through estuaries to the upstream
areas for breeding and then return to the sea. Shrimps and prawns constitute the most
important fishery of the zone. The most important crustaceans‘ species are Penaeus
monodon and Macrobrachium rosenbergii. Mud crab (Scylla serrata) is the largest
edible crab found in the forest area and has a high economic value because of its very
tasty meat and very high nutrient content.
At present, there are 5550 registered boats to fish within the forest area of
Sundarban (GoWB 2019). Each of these boats, which are quite small in size, has a
capacity of carrying about four persons and make several trips to the forests. Most of
the fishing takes place from November to January; a lesser amount of catch is
generally made between March and June. Due to the rough condition of the weather,
fishing activity does not take place during other times of the year.
Sundarban is the top producer of fish and prawn, producing roughly 31% of the
total inland fish/prawn production of West Bengal (Das et al. 2016). Brackish water
prawn aquaculture started along the creeks of reclaimed Sundarban from the late
1980s and developed quite rapidly. Prawn farming is basically a capital-intensive
mechanized undertaking that does not require large pools of labours. The people of
Sundarban, however, got involved in the activity in a different way, by collecting
seeds of tiger prawns from the coastal water during high tides. These then are sold to
the prawn farms through brokers for nourishing. Mostly carried out by the women
and children, the prawn seed collection became an economic sidekick for many
families. In the fringe areas of the forest, close to 6000 people sieve the shallow
intertidal water for prawn seeds regularly. In a study of Bangladesh, Angell (1994)
estimates that one small-scale hatchery with three cycles of production per year can
produce about 2 million post larvae, with fixed costs of about US$23,000 and
operating costs of about US$9000. West Bengal being the largest producer of tiger
shrimp among Indian maritime states and the majority of this was produced in Indian
Sundarban covering 0.21 million ha potential brackish water areas in the state
(Ghoshal et al. 2019). The areas where fishing is permitted within the buffer zones
(522.85 km2) are congested and overfished. The fishermen have to enter the core
areas secretly, risking a fine of Rs 500 (US$7.76) for the first offence, Rs 1000
(US$15.52) for the second, and Rs 1100 (US$17.08) for the third if caught by a
forest guard during patrolling (Sen and Pattanaik 2017).
Fishing activities in Sundarban have increased from 19% of the total workforce in
2000–2001 to 33.0% in 2010–2011 (Mistri and Das 2020). After Aila in 2009,
fishing increases due to the loss of huge agricultural land by sudden increasing
salinity. Approximately 11% of households in the Sundarban listed ‘fishing’ as one
of the family occupations (Sanchez-Triana et al. 2014). This percentage goes up to
60–70% in areas with easy access to rivers (Sen 2019). However, according to the
government report, more than one lakh families are engaged in inland fishing
(GoWB 2005).
384 C. S. Das

Steady declination of mangrove fisheries resources either due to overfishing or to


habitat degradation or both is observed in Sundarban with other parts of Asia (Hasan
and Naser 2016). Many residents of the Sundarban are dependent on income from
gathering of natural Penaeus monodon post larvae wild-caught brood stock and seed
for aquaculture, despite the formal ban on natural fry collection. This is one of the
main sources of earning for the small and landless fisherman and women of this area.
7–99 mm post larvae and juveniles were available throughout the year with peaks in
June, July, and December (De et al. 1978). The prawn seed collection is a highly
destructive practice with a high by-catch rate that results in the capture and discard of
non-target species and exerts a heavy toll on the sustainability of marine, estuarine,
and freshwater fish species. Various studies in Sundarban suggest that an alarming
declining rate in various fish species was observed in Sundarban (Hoq 2007). For
every tiger prawn seed collection, about 400 of others species (fishes, crabs, other
prawn, mollusc, etc.) are destroyed (Santhakumar et al. 2005; Mahapatra et al.
2014). Tiger shrimp seeds are fast dwindling away from the natural waters of
Sundarban due to overfishing at various stages of its life cycle. As its post larval
stage in estuaries, it is trapped by fine push and drag nets and fine meshed bag nets
(meen jal); the juveniles are trapped by bag nets (behundi jal) in estuaries; the
juveniles and pre adults are caught in marine waters by large bag nets; the pre adults
and adults by trammel nets. Even the spawns are not spared and are caught from the
open seas by trawl nets. (Mahapatra et al. 1999). Overexploitation and unscientific
practices of aquatic resources were found to be the key factors for the declination of
fishery resources in Sundarban.

16.6 Conservation and Livelihood Conflicts

Human survival and economic well-being are fully dependent upon biological
diversity that encompasses all life forms, ecosystems, and ecological processes,
acknowledging the hierarchy at genetic, taxa, and ecosystem levels (Wackernagel
et al. 1997). The more is the biodiversity the greater is to access available resources,
along with increased net primary production and decreased nutrient loss (Singh
2002). Globally top-down approaches suffer from the ailment of failing to factor
in the participation of local people as part of the conservation process (Hazarika and
Kalita 2019).
In the year 1973, the Government of India created the Sundarban Tiger Reserve,
and subsequently the creation of buffer and core area (Nishat 2019) which restrict the
forest users leading to deprived livelihood. It gave rise to conflict between forest
managing authorities and local habitants. Excluding the livelihood issues of fringe
dwellers from the conservation, the process makes it difficult to enforce any conser-
vation strategies successfully. However, fringe dwellers of Sundarban had invariably
participated in the conservation process long before it had started. Any type of
disruption of traditional ways of living can trigger unwanted adverse impacts on
local communities (García-Frapolli et al. 2009) which are tantamount to the
poaching of wildlife. Fishing in Sundarban has a long history and is a generational
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 385

Table 16.5 Forest entrants in STR area in a typical year between 2013 and 2017
Legal entrants (permit Illegal entrants (approximate
holdersa: authorized for estimation by Forest Department Total
Category multiple entries) staff/local survey) (approximate)
Fishermen 3840 c. 7000 10,840
Honey 750 250 1000
collectors
Nypa palm 1500 c. 600 2100
collector
Tiger prawn None c. 12,000 (during high tide) 12,000
collectors
Crab None c.18000 18,000
collectors
Forest 237 None 237
Department
staff
Poachers – c. 100 100
Total 6267 c. 37,900 45,167
Source: Field Study
a
No fresh permits are being issued by the Forest Department since 1992. It at present only renews
the permits granted before 1992

occupation. The fishers have engaged in forest-based economic activities since


civilization began in Sundarban. During the formation of the STR, its ‘Management
Plan’ mentioned that only non-motorized boats would be allowed in the buffer areas
for fishing. The people, especially communities, who depend on the natural environ-
ment of their locality to meet most of their material needs are known as ‘ecosystem
people’ (Kothari and Patel 2006). Any type of human interference or boatings is
strictly prohibited in the core areas. In the buffer that surrounds the core, access is
allowed for livelihood purposes with a proper license and permit. But the major issue
is that most of the forest entrants in Sundarban are illegal as they have no permits of
resource collection issued by the forest department because of livelihood demands.
According to a field survey estimate in 2018 near about 45 thousand forest users
enter into the forest (both in restricted and non-restricted areas) for livelihood issues
and surprisingly, only 14 percent are legal entrants (Table 16.5) As a result, success
of different conservation attempts is very limited and there is an imbalance between
resource collection and regeneration as well.

16.7 Livelihood vs Animal Attacks

Animal attacks on humans are common in Sundarban. Among the attacks from
animals of the protected areas, tigers top the list followed by attacks by crocodiles
and sharks. Attacks from crocodiles and tigers often prove fatal. The straying of
tigers from reserve forests into the human habitations also poses a major problem for
the residents living along the forest boundary. Conflicts between wildlife and
386 C. S. Das

humans in Sundarban are evidently owing to an increase in human population,


extensive loss of natural habitats, and increase in dependency of forest resources.
Conflicts are most acute when a species involved is critically imperiled while its
presence in an area poses a significant threat to human welfare (Saberwal et al.
1994). Human-wildlife conflict is potentially any situation where: (1) the behaviour
of people negatively impacts wildlife (this includes human impacts on habitat);
(2) the behaviour of wildlife creates a negative impact for some stakeholders or is
perceived by some stakeholders to impact themselves or others adversely, or (3) the
wildlife focused behaviour of some people creates a negative interaction with other
people, often in the form of a values clash. Thus, a people-wildlife problem can
involve a people–wildlife interaction or a people–people interaction (i.e. a contro-
versy) or both (Decker and Chase 1997).
Between 1998 and 2017, 546 forest entrants were attacked by tigers out of which
485 succumbed to their injuries with an average of 27.2 events per year (Das 2018).
Some 14 percent of the victims were honey collectors, 5 percent were woodcutters,
and as much as 80 percent were fishermen including crab collectors. About 1 percent
of the victims were forest staff (Das 2018).
Crocodile victims are generally of two types—fishermen and tiger prawn seed
collectors. In Sundarban, hundreds of people, mostly women, and young children
get engaged in the prawn seed collection every day. Wading through waist-deep or
even neck-deep water, they filter out of spawn of shrimps using fine nylon nets. In an
area where the scope of alternative employment is limited, this activity has become
popular in Sundarban during the last two decades as it yields very high returns
(Ghosh et al. 2018). It is also done on a commercial scale, with nets spread across
almost the entire width of the river with help of boats and buoys. According to a
survey, some 132 people were attacked by crocodiles during 2000 to 2019; out of
these, 61.16 percent succumbed to death on an average 7.9 persons every year.
Almost 80 per cent of victims were prawn seed collectors and belonged to the age
group of 11 to 50. They were mostly children and women. Male victims are slightly
lower (46.60%) than the females (53.40%). Most of the cases were recorded from
Gosaba (34%), followed by Patharpratima (25.24%) and Namkhana (18.45%) (Das
2018). Apart from the crocodiles, the persons being exposed to the creeks of
Sundarban are also vulnerable to attack from sharks locally called kāmots. Shark
bite is a relatively recent phenomenon in Sundarban that started to arise about 15 to
20 years ago (Das 2017). Increasing population pressure and dire poverty urge the
people to take the risk of facing natural hazards as well as attack from wild animals
as they venture into the jungle. Trespassers take undue advantage of this human
presence in the zone for pilferage of forest produces and poaching of wild animals. It
is also not uncommon for the animals to stray into human habitations and to cause
depredations. All these lead to conflict between humans and animals—the root cause
of which is socio-economical.
On an average, one forest user spent the time as 3–4 h/day in the minimum and
10–12 h/day in the maximum range, which accounted for about 15 days/month. In
the purpose of crab collection, the duration used to vary from 3–4 h/day to 10–12 h/
day, accounted for about 20–25 days/month. The collection of tiger prawn seeds
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 387

(prawn: Penaeus monodon) was considered to be a hectic job for which one used to
spend 6–7 h/day in the water, utilizing 21 days/month. In this activity, individuals
put their maximum effort with high concentration to segregate small seeds of tiger
prawn from heterogeneous species of juvenile fishes. Tiger prawn seeds had cost
around Rs. 4/piece. Comparatively, the wood collection was recorded to spend about
5 h/day and treated as a concurrent activity done simultaneously with a collection of
fishing and crab. Analyses of activities that led to a maximum exposure clearly
showed that entering into creeks in forested areas for fishing, collection of crab,
prawn seed, and fuelwood compelled the people to become most vulnerable.
Analyses of seasonal incidences showed that maximum exposure normally hap-
pened during winter (59.7%), followed by monsoon period (18.6%) and then the
summer months being remarkably low (1.7%) (Das 2017, 2018).

16.8 Case Study: Livelihood of Three Fringe Villages


of Sundarban

For identifying the livelihood pattern of the people of the fringe areas of Sundarban,
a household survey was done in three villages of the Sundarban under three blocks
such as Gosaba, Hingalganj, and Kultali (Fig. 16.2) with a total area of 22.95 km2
and a population of 17,728 (Census 2011). These mouzas are, here, known as fringe
villages of Sundarban as these are bordered with the Sundarban Reserve Forest (Das
2015). Second, these villages were selected because a large section of dwellers
depends on the forest resources of Sundarban for their daily livelihood.

16.8.1 Materials and Methods

A comprehensive database of the forest users was made through a door-to-door


household survey with questionnaires interviewed with earning member(s) of all
households during a period of May 2017 to April 2018. Collection of data on land
use and other socio-economic aspects like land holding (Table 16.7), occupation
(primary, secondary, and tertiary), caste, literacy, availability of drinking water, use
of electricity, and 100 days works initiated by Govt. of India for provision of jobs to
all. Total number of respondents was 1079 of 4161 households of three selected
villages in the fringe areas in Sundarban (Table 16.6). Of the total respondents, only
505 (46.8%) were female. The eldest respondent was 95 years old when the mean
age was 46.31 years.
Simultaneously, another in-depth interviewing was conducted on marginal
workers (livelihood for less than 6 months) with a separate questionnaire to
157 respondents [Samsernagar (52), Rajatjubille (68) and Kisorimohanpur (47)]
with a criteria of at least one member directly involved in activities like fishing,
honey collection, crab collection as the source of primary occupation. Data was
collected on dependency upon forest and forest resources, frequency of entering
forest, type of resources collected, rate of daily utilization of resources, amount of
388 C. S. Das

Fig. 16.2 Location of Fringe villages in Sundarban

Table 16.6 Sample study related population data


No. of Population Area No of Mean Marginal
Village households (2011) (km2) samples age workers
Samsernagar 1145 4394 6.83 345 48 558
Rajatjubillee 1745 6851 8.51 422 56 1937
Kisorimohanpur 1271 6483 7.65 372 52 1373
Total 4161 17,728 22.95 1079 50.3 3868

resources sold in the markets, income from sale of resources, type of fuel used,
average use of fuel on daily basis, financial source for procurement of essential
items, exposure to human-wildlife and its conflict, consequence of exposure to wild
life, and role of eco-development committee (EDC) or forest protection committee
(FPC) (Table 16.7).
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 389

Table 16.7 Land status of households of selected locations


Percentage of landless people
Samsernagar Rajatjubille Kisorimohanpur
Direct forest users category (n ¼ 167) (n ¼ 160) (n ¼ 155) Average
Fishermen 26.19 32.23 29.35 39.25
Crab catcher 30.91 19.20 31.44 27.18
PS collection 22.91 28.43 17.86 23.06
Honey collector 5.71 14.88 17.90 12.83
Nypa palm collector 14.28 5.26 3.45 7.66
Total 100.00 100.00 100.00 21.96
Source: Field Survey

16.8.2 Category of Occupation and Resources

The study listed 14 primary occupations (A), which constitute further 8 occupational
types (Table 16.8). Each primary occupation was distinct from another based on its
activity and tools used, though resources might be similar as biological perspective.
The term ‘primary occupation’ referred to the activities which supported the major
livelihood of forest dwellers and maximum time spent. The ‘secondary occupation’
(B) referred to those activities which were usually performed during the lean phase
of primary occupation and were treated as the supplementary revenue generation.
The ‘tertiary occupation’ (C) happened very rarely and was not much support for
livelihood. However, the concept of primary, secondary, and tertiary occupations
was relative to respective dwellers. The study considered the collection of aquatic
resources (collection of finfish, shellfish, prawn seed, and crab) as a primary occu-
pation to those having major earning for livelihood and maximum time spent. When
the same occupants took agriculture during the lean phase, the activity was consid-
ered as a secondary occupation. All these occupants were made four groups based on
their characteristic of dependency like resources dependent (RD) includes five
activities, wage earner (WE) to four activities, self-resilient (SR) to four activities
and Government servant (GS).
The major occupation of the villagers was based on a collection of aquatic
bio-resources available in surrounding creeks, rivers, and estuaries. Resource depen-
dency activities were the major occupations in all the regions of Sundarban with
nearly 60% population being engaged followed by wage earners and self-resilience.
Collection of aquatic bio-resources (fishing, crab, and prawn collection) was the
principal occupation of the study areas followed by important secondary occupations
like agriculture, honey collection, fuel wood collection working as labours and
engaging in business and trading.
390

Table 16.8 Occupation categories and types in three villages of Sundarban

Occupational Category of occupation (%)


category Occupation type Resources Samsernagar Rajatjubile Kisorimohanpur
Resource dependent Collection of aquatic Fin fish, shell fish crab, prawn seed 62.96 59.94 56.94
resource
Collection of forest Honey, wax, wood, timber, leaf
resource
Agriculture Paddy and vegetables
Wage labour Wage earning Wood processing daily useable, goods domestic 26.2 31.60 31.6
works
Self-resilience Business Domestic items, chemical fertilizer 8.67 6.29 9.46
Self-earning Wood and timber, fermented palm juice
Tuition School children
Government service Permanent salaried job Government department 2.17 2.17 2.0
Total 100 100 100
Source: Field Survey
C. S. Das
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 391

16.8.3 Relative Livelihood Index

Relative Livelihood Index (RLI) was developed to have a comparative assessment of


the degree of strength among occupations (Fig. 16.3). The objective of this index
was to assess whether the strength of primary occupation was alone robust to support
the occupants continuing livelihood. If the sum of the total strength of secondary and
tertiary occupations were greater than that of the primary one, then primary occupa-
tion was considered to be weak and dwellers belonging to it manifested a poor state
of livelihood.
Our of 8 major occupations (Fishing and crab collection, Agricultural labour,
Crab collection, Prawn seed collection, Casual labour, Honey collection, Fishing,
Carpentry) four occupations scored positively, whereas the other four showed a
negative trend (Table 16.9 and Fig. 16.4). Practically, occupants engaged in primary
occupations measured with negative scores depended on secondary and tertiary

Occupational Structure
Govt.
Servant Samsernagar
Category of Occupation

Rajatjubille
Self Kishorimohanpur
Resilient

Wage
Labour

Resource
Dependent

0 50 100 150 200


% of Occupation

Fig. 16.3 Occupational structure of selected villages of Sundarban

Table 16.9 Calculation of relative livelihood index (RLI)


Score (number)
Sl. No. Primary occupation Samsernagar Rajatjubille Kisorimohanpur
1 Fishing and crab collection 60.97 23.35 19.80
2 Agricultural labour 21.38 4.82 24.75
3 Crab collection 17.6 10.85 4.95
4 Prawn seed collection 14.59 8.88 4.95
5 Casual labour 4.82 12.62 9.90
6 Honey collection 23.68 27.78 43.56
7 Fishing 50.08 31.4 19.8
8 Carpentry 34.99 43.32 52.78
RLI ¼ [A  3  {(B  2) + (C  1)}], where A ¼ Primary occupation, B ¼ Secondary occupation
and C ¼ Tertiary occupation
Source: Field Survey
392 C. S. Das

Score
80
Fishing &
crab collection
60 Agricultural
labour
40 Crab collection
Score (Number)

20 Prawn seed
collection
Casual labour
0
Samsernagar Rjatjubille Kishorimohanpur Honey
–20 collection
Fishing
–40
Carpentry
–60

Fig. 16.4 Livelihood index of selected villages of Sundarban

Fig. 16.5 Food chain of mangrove ecosystem of Sundarban

occupations for livelihood, since primary occupations of that particular livelihood


were weak. Three occupations such as ‘Fishing & crab collection’, ‘Crab collection’,
and ‘Prawn seed collection’ which had positive scores among others were exclu-
sively resourced dependence. Occupants were inherently skillful in these
occupations in one way and found many earnings/catches in collecting natural
resources in others. On the other hand, occupations like casual labour, honey
collection, fishing and carpentry scored negatively meaning that these belonged to
a poor state of livelihood and for sustenance occupants had to depend on other
occupations as secondary or tertiary in spite of their lack of skillfulness (Figs. 16.4
and 16.5).
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 393

Strength of fishing and crab collection as a major primary occupation is observed


in Samsernagar and Rajatjubille, whereas in Kisorimohanpur agricultural labour
dominates others. However, in the entire region, the primary occupation like
honey collection, fishing, and carpentry scores negatives which justify that this
occupation was alone robust to support the occupants continuing livelihood.

16.9 Conclusion

Sundarban mangrove ecosystem warrants maintaining the food chain of terrestrial


and aquatic systems simultaneously apart from other ecosystem services necessary
for the livelihood of the rural people living in the fringe areas (Fig. 16.5). Their major
collection includes honey, fuelwood, fishes, and crabs both legally and illegally. The
dwellers of the Sundarban have to face both visible and not-so-visible uncertainties
in their traditional livelihoods like agriculture and fishing. On the other hand, huge
population pressure forces the dwellers to make use of these natural resources
terribly and indiscriminately, which gave rise to conflict between forest managing
authorities and local habitants. Over exploiting of resources cannot be controlled
without the execution of appropriate alternate livelihood arrangements including
satisfactory marketing transactions. As such, there is a need for appropriate water
and land care system for the entire Sundarban to make this fragile ecosystem into
climatically more resilient and sustainable. Cooperative basis modern aquaculture
practice and crab rearing, excavation of ponds for rainwater harvesting along
pisciculture are some of the alternatives to provide independency of villagers for
livelihood. Plantation with fast-growing trees of high calorific value can be initiated
as bio-resource which may be common property to be available largely on the
roadside, canal side, and river embankments, under the control of Panchayat.
These attempts will make the villagers become independent as well as self-resilient
in the collection of fuelwood in one way and discourage them to enter into the forest
in others and will thereby minimize the risk of exposure to man-eaters both in
terrestrial and water bodies. More research will promote more widespread efforts
to develop conservation and sustainable development policies that integrate steady
increasing salinity of the region, changes in mangrove dynamics, and the welfare
impacts on poor and marginal communities. The socio-economic characteristics of
these communities should be considered as the main criteria for designing any
conservation plan as well as a sustainable livelihood programme which would
reproduce both benefit-sharing and incentive measures applied to all stakeholders.

Acknowledgement I am grateful to Swapan Paul, a Research scholar of West Bengal State


University for helping to prepare Figures no.1 and 2. Thanks are also due to Semonti Das, Assistant
professor of Chandrakona Vidyasagar Mahavidyalaya for checking citation as well as references of
the manuscripts.
394 C. S. Das

References
Ajibola A, Chamunorwa JP, Erlwanger KH (2012) Nutraceutical values of natural honey and its
contribution to human health and wealth. Nutr Metab 9(1): 61.
Alongi DM (2002) Present state and future of the world’s mangrove forests. Environ Conserv 29(3):
331–349.
Angell CL (1994) Promotion of Small-scale Shrimp and Prawn Hatcheries in India and Bangladesh.
Bay of Bengal Program (BOBP), Madras, India.
Ashley C, Carney D (1999) Sustainable livelihoods: Lessons from early experience (Vol. 7, No. 1).
Department for International Development. London.
Badhsheeba MA, Vadivel V (2020) Physicochemical and phytochemical contents of the leaves of
Acrostichum aureum L. J Glob Biosci 9(4): 7003–7018.
Bandaranayake WM (1998) Traditional and medicinal uses of mangroves. Mangroves Salt Marshes
2: 133–148.
BARC (1998) Changes in cropping pattern in Sundarban over last two decades. Bangladesh
Agriculture Research Council, Dhaka, Bangladesh, p 10.
Bradbear N (2003) Bees and rural livelihoods. 16pp booklet in English, Portuguese and Spanish
editions. Bees for Development, Monmouth, UK.
Bradbear N (2005) Bees for Development Journal 74. Bees for Development, UK.
Bunt JS, Robertson AI, Alongi DM (1992) Introduction. In: Tropical mangrove ecosystem.
American Geophysical Union, Washington, DC, p 1–6.
Burgett M (2000) Honey hunters of the Sundarban, Bees Dev Journal 56: 6–7.
Census (2011) Primary census abstracts, Registrar General of India, Ministry of Home Affairs,
Government of India, Available at: http://www.censusindia.gov.in/2011census/PCA/pca_
highlights/pe_data.
Chakraborty K (1985) In: Abstract, Sundarbans Mangrove Biomass Productivity and Resources of
Utilization of mangrove. Proceedings of National Symposium of Biology, Utilization and
Conservation of mangrove.
Chambers R, Conway G (1992) Sustainable rural livelihoods: Practical concepts for the 21st
century. IDS Discussion Paper 296, IDS, Brighton.
Chapman VJ (1976) Mangrove vegetation. In: Cramer J, p 447.
Chatterjee T (2017) Climate change and its collision on fisheries resource in sundarban region of
south 24 pargana, West Bengal, Suresh Gyan Vihar University. Int J Environ Sci Technol 3(2):
10–13.
Chaudhuri AB, Choudhury A (1994) Mangroves of the Sundarbans. Volume 1, The IUCN Wetland
Program. India.
Chowdhury A, Naz A, Maiti SK (2017) Health risk assessment of ‘tiger prawn seed’ collectors
exposed to heavy metal pollution in the conserved mangrove forest of Indian Sundarbans: A
socio-environmental perspective. Hum Ecol Risk Assess 23(2): 203–224.
Chowdhury AN, Mondal R, Brahma A, Biswas MK (2020) Eco-psychiatry and environmental
conservation: Study from Sundarban Delta, India. Environ Health Insights 2(1): 61.
Danda AA (2007) Surviving in the Sundarbans: Threats and responses. Unpublished PhD, Univer-
sity of Twente. p 203.
Danda AA, Sriskanthan G, Ghosh A, Bandyopadhyay J, Hazra S (2011) Indian Sundarbans Delta:
A vision. World Wide Fund for Nature-India, New Delhi, p 40.
Darwin NT (2014) Livelihood strategies of people surrounding the Sundarban mangrove forest
(Doctoral dissertation, Charles Darwin University).
Das CS (2012) Tiger straying incidents in Indian Sundarban: Statistical analysis of case studies as
well as depredation caused by conflict. Eur J Wildl Res 58(1): 205–214.
Das CS (2015) Causes, consequences and cost-benefit analysis of the conflicts caused by tiger
straying incidents in Sundarban, India. Proceedings Zoological Soc 68(2): 120–130.
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 395

Das CS (2017) Human-wildlife conflicts in the Sundarbans, In: Danda AA, Joshi AK, Ghosh A,
Saha, R (eds) State of art report on biodiversity in Indian Sundarbans. World Wide Fund for
Nature-India, New Delhi. p 325–340.
Das CS (2018) Pattern and characterisation of human casualties in Sundarban by Tiger attacks,
India. J Sustain For 1(2): 1–10.
Das CS, Bandyopadhyay S (2012) Sharing space—human animal conflicts in Indian Sundarban.
Das P, Das A, Roy S (2016) Shrimp fry (meen) farmers of Sundarban Mangrove Forest (India): A
tale of ecological damage and economic hardship. Int J Agricult Food Res 5(2): 28–41.
Das CS & Mandal RN (2016) Coastal people and mangroves ecosystem resources vis-à-vis
management strategies in Indian Sundarban. Ocean & Coastal Management, 134, 1–10.
Dasgupta S, Sobhan I, Wheeler D (2017) The impact of climate change and aquatic salinization on
mangrove species in the Bangladesh Sundarbans. Ambio 46(6): 680–694.
Day Jr JW, Conner WH, Ley-Lou F, Day RH, Navarro AM (1987) The productivity and composi-
tion of mangrove forests, Laguna de Terminos, Mexico. Aquat Bot 27(3): 267–284.
Day Jr JW, Coronado-Molina C, Vera-Herrera FR, Twilley R, Rivera-Monroy VH, Alvarez-
Guillen H, Day R, Conner W (1996) A 7 year record of above-ground net primary production
in a southeastern Mexican mangrove forest. Aquat Bot 55(1): 39–60.
De DK, Chakaraborty RK, Subramanian M (1978) Preliminary survey on the availability of
culturable brackishwater prawn seed in the Haldia estuary around Baughata, West Bengal. J
Inland Fish Soc India 10: 142–143.
Decker DJ, Chase LC (1997) Human dimensions of living with wildlife: A management challenge
for the 21st century. Wildlife Soc Bullet (1973–2006) 25(4): 788–795.
DFID GS (2000) Sustainable livelihoods guidance sheets, Section 2. Framework.
García-Frapolli E, Ramos-Fernández G, Galicia E, Serrano A (2009) The complex reality of
biodiversity conservation through Natural Protected Area policy: Three cases from the Yucatan
Peninsula, Mexico. Land Use Pol 26(3): 715–22.
German Agro Action Rural India (2009) Land shaping for crop irrigation and to prevent water
stagnation in Sunderbans, West Bengal https://www.indiawaterportal.org/articles/ramakrishna-
mission-lokasiksha-parishad-rkmlsp-promotes-land-shaping-crop-irrigation-and-0
Ghosh A (2010) The shadow lines. Penguin Books India.
Ghosh A, Schmidt S, Fickert T, Nüsser M (2015) The Indian Sundarban mangrove forests: History,
utilization, conservation strategies and local perception. Diversity 7(2): 149–169.
Ghosh P (2014) Subsistence and biodiversity conservation in the Sundarban Biosphere Reserve,
West Bengal, University of Kentucky, Theses and Dissertations-Geography. p. 26. https://
uknowledge.uky.edu/geography_etds/26
Ghosh U, Bose S, Bramhachari R (2018) Living on the edge: Climate change and uncertainty in the
Indian Sundarbans. STEPS Working Paper 101, Brighton: https://opendocs.ids.ac.uk/opendocs/
handle/20.500.12413/13597
Ghoshal TK, De D, Biswas G, Kumar P, Vijayan KK (2019) Brackishwater aquaculture:
Opportunities and challenges for meeting livelihood demand in Indian sundarbans. In The
Sundarbans: A Disaster-Prone Eco-Region. Springer, pp. 321–349.
Giri C, Pengra B, Zhu Z, Singh A, Tieszen LL (2007) Monitoring mangrove forest dynamics of the
Sundarbans in Bangladesh and India using multi-temporal satellite data from 1973 to 2000.
Estuar Coast Shelf Sci 73(1–2): 91–100.
Glavovic B, Scheyvens R, Overton J (2003) Waves of adversity, layers of resilience. Exploring the
sustainable livelihoods approach. In: Overton DJ, Nowak B (eds) Contesting development:
Pathways to better practice. Proceedings of the Third Biennial Conference of the Aotearoa
New Zealand International Development Studies Network (Dev Net), Massey University, Dec
5–7, 2002. Institute of Development Studies, p 289–293.
Gong WK, Ong JE, Wong CH, Dhanarajan G (1984) Productivity of mangrove trees and its
significance in a managed mangrove ecosystem in Malaysia. In: Proceedings of the Asian
Symposium on Mangrove Ecosystem in Research and Management (Eds. Soepadmo E., Rao
A.N. and Macintosh D.J.) Kuala Lumpur: University of Malaya. p 216–225.
396 C. S. Das

GoWB (2005) Administrative Report: 2004–2005: Sundarbans Development Board, Sundarbans


Affairs Department, Government of West Bengal.
GoWB (2019) Annual report 2017–2018. Wildlife wing, Directorate of Forest, Govt. of West
Bengal: p 227.
Hasan R, Naser N (2016) Fishermen livelihood and fishery resources of the Sundarbans reserved
forest along the Mongla port area Bangladesh. Int J Fish Aquat Stud 4(3): 468–475.
Hazarika AK, Kalita U (2019) Conservation and Livelihood Conflict of Kaziranga National Park: A
World Heritage Site of Assam, India. Space Culture, India 7(3): 224–232. https://doi.org/10.
20896/saci.v7i3.656
Hazra S, Samanta K (2016) Temporal Change Detection (2001–2008): Study of Sundarban.
Working Papers. ID: 10526, eSocialSciences.
Hegde R, Suryaprakash S, Achoth L, Bawa KS (1996) Extraction of NTFPs in the forests of Biligiri
Rangan Hills, India: Contribution to rural income. Econ Bot 50(3): 243–251.
Hill DB, Webster TC (1995) Apiculture and forestry (bees and trees). Agroforest Syst 29 (3):
313–320.
Hogarth PJ (1999) The biology of mangroves. Oxford University Press (OUP).
Hoq ME (2007) An analysis of fisheries exploitation and management practices in Sundarbans
mangrove ecosystem, Bangladesh. Ocean Coast Manage 50(5–6): 411–427.
Hossain SJ, Iftekharuzzaman M, Haque MA, Saha B, Moniruzzaman M, Rahman MM, Hossain H
(2016) Nutrient compositions, antioxidant activity, and common phenolics of Sonneratia apetala
(Buch.-Ham.) fruit. Int J Food Propert 19(5): 1080–1092.
Islam MM, Borgqvist H, Kumar L (2019a) Monitoring mangrove forest landcover changes in the
coastline of Bangladesh from 1976 to 2015. Geocarto Int 34(13): 1458–1476.
Islam MN, Ratul SB, Sharmin A, Rahman KS, Ashaduzzaman M, Uddin GM (2019b) Comparison
of calorific values and ash content for different woody biomass components of six mangrove
species of Bangladesh Sundarbans. J Indian Acad Wood Sci 16(2): 110–117.
IUCN (1994) The World Conservation Union. In: Hossain Z, Acharya G (eds) Mangroves of
Sundarbans vol. 2: Bangladesh. Bangkok, Thailand: IUCN-The World Conservation Union,
p 230.
Junk WJ, Brown M, Campbell IC, Finlayson M, Gopal B, Ramberg L, Warner BG (2006) The
comparative biodiversity of seven globally important wetlands: A synthesis. Aquat Sci 68(3):
400–414.
Kamruzzaman M, Ahmed S, Osawa A (2017) Biomass and net primary productivity of mangrove
communities along the Oligohaline zone of Sundarbans, Bangladesh. Forest Ecosys 4(1): 1–9.
Kar A, Raut SK, Bhattacharya M, Patra S, Das BK, Patra BC (2017) Marine fishes of West Bengal
coast, India: Diversity and conservation preclusion. Reg Stud Mar Sci 16: 56–66.
Khan MN, Khatun S, Azad MS, Mollick AS (2020) Leaf morphological and anatomical plasticity in
Sundri (Heritiera fomes Buch.-Ham.) along different canopy light and salinity zones in the
Sundarbans mangrove forest, Bangladesh. Glob Ecol Conserv 23: e01127.
Kibria AS, Costanza R, Groves C, Behie AM (2018) The interactions between livelihood capitals
and access of local communities to the forest provisioning services of the Sundarbans Mangrove
Forest, Bangladesh. Ecos Serv 32: 41–49.
Komiyama A, Havanond S, Srisawatt W, Mochida Y, Fujimoto K, Ohnishi T, Ishihara S, Miyagi T
(2000) Top/root biomass ratio of a secondary mangrove (Ceriops tagal (Perr.) CB Rob.) forest.
Forest Ecol Manage 139(1–3): 127–134.
Komiyama A, Ogino K, Aksornkoae S, Sabhasri S (1987) Root biomass of a mangrove forest in
southern Thailand. 1. Estimation by the trench method and the zonal structure of root biomass. J
Tropic Ecol 1: 97–108.
Kothari A, Patel A (2006) Environment and human rights: An introductory essay and essential
readings. National Human Rights Commission.
Kothari MJ, Rao KM (1995) Dolichandrone spathacea (Lf) K. Schum.(Bignoniaceae)-a little
known plant for Goa. J Econ Taxon Bot 19(2): 323–324.
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 397

Kothari N (2015) The plight of the honey collectors in Sundarban: Challenges, uncertainties
strategies and survivals. J Human Soc Sci 20(2): 56–59.
Kumar KS, Bhowmik D, Biswajit C, Chandira MR (2010) Medicinal uses and health benefits of
honey: An overview. J Chem Pharm Res 2(1): 385–395.
Mahapatra BK, Chatterjee P, Saha D, Datta NC (1999) Declining trend in the abundance of seeds of
tiger shrimps, Penaeus monodon (fabricus) in the Sundarbans with suggestion for restorations.
Sundarbans Mangal 532–538.
Mahapatra BK, Sarkar UK, Lakra WS (2014) A review on status, potentials, threats and challenges
of the fish biodiversity of West Bengal. J Biodivers Biopros Dev 2: 140. https://doi.org/10.4172/
2376-0214.1000140
Mandal RN, Das CS, Naskar KR (2010) Dwindling Indian Sundarban mangrove: The way out. Sci
Cult 76(7–8): 275–282.
Mandal RN & Bar R (2018) Mangroves for Building Resilience to Climate Change. CRC Press.
Mandal RN, Meher PK, Naskar KR (2009) Effect of salinity on germination and seedling develop-
ment of Heritiera fomes Buch. Ham. Proc Nat Acad Sci India, B-Biol Sci 79: 153–160.
Mandal RN, Naskar KR (2008) Diversity and classification of Indian mangroves—a review. Trop
Ecol 49(2): 131–146.
Mandal RN, Saenger P, Das CS, Aziz A (2019a) Current status of mangrove forests in the trans-
boundary Sundarbans. In: Sen HS (Ed.) The Sundarbans: A disaster-prone eco-region. Springer
Nature Switzerland, p 93–131.
Mandal S, Burman D, Mandal UK, Lama TD, Maji B, Sharma PC (2017) Challenges, options and
strategies for doubling farmers’ income in West Bengal–reflections from coastal region. Agricul
Econ Res Rev 30: 89–100.
Mandal UK, Burman D, Bhardwaj AK, Nayak DB, Samui A, Mullick S, Mahanta KK, Lama TD,
Maji B, Mandal S, Raut S (2019b) Water logging and coastal salinity management through land
shaping and cropping intensification in climatically vulnerable Indian Sundarbans. Agric Water
Manage 216: 12–26.
Mann KH (1982) Ecology of coastal waters: A systems approach. University of California Press.
McDowell MC (1995) The development of the river Hoogli for navigation Commemoration
volume of Port Calcutta, p 69.
Miah G, Bari N, Rahman A (2003) Agricultural activities and their impacts on the ecology and
biodiversity of the Sundarban area of Bangladesh. J Natl Sci Foundation 31: 175–199.
Mistri A, Das B (2020) Environmental change, livelihood issues and migration. Springer.
Naskar K, Ghosh A (1989) Mangrove forest of the Sundarbans: Its impact on estuarine fisheries.
Proceeding in Coast Zone Management of West Bengal, Sea Explores Institute, Calcutta,
p A.47–59.
Naskar KR, Guha Bakshi DN (1987) Mangrove swamps of the Sundarbans–an ecological perspec-
tive. Naya Prokash, Calcutta.
Naskar KR, Mandal RN (1999) Ecology and biodiversity of Indian mangroves. Daya Books.
Nishat B (2019) Landscape narrative of the Sundarban: Towards collaborative management by
Bangladesh and India. The World Bank.
Pakrashi H. (2016) Is there a tomorrow?: The story of survival of Sunderbans against climate
change. In: Nautiyal S., Schaldach R., Raju K., Kaechele H., Pritchard B., Rao K. (eds) Climate
change challenge (3C) and social-economic-ecological interface-building. Environmental sci-
ence and engineering. Springer, Cham. https://doi.org/10.1007/978-3-319-31014-5_21
Peters CM, Gentry AH, Mendelsohn RO (1989) Valuation of an Amazonian rainforest. Nature 339:
655–656.
Pillai RS, Ong BL (1999) Effects of inorganic nitrogen availability on the sporophytes of
Acrostichum aureum L. Photosynthetica 36(1–2): 259.
Putz FE, Chan HT (1986) Tree growth, dynamics, and productivity in a mature mangrove forest in
Malaysia. Forest Ecol Manage 17(2–3): 211–230.
Ray P (2000) Sundarban–A goldmine for prawn fisheries. Paper presented at the Celebration of
Fourth Sundarbans Day, Kolkata Wildlife Society. Kolkata, p 59–64.
398 C. S. Das

Robertson AI, Alongi DM Boto KG (1992) Food chains and carbon fluxes. In: Robertson AI,
Alongi DM (eds), Tropical mangrove ecosystems. American Geophysical Union,
Washington. p. 193–326.
Rodrigues BF (2006) Biodiversity loss and its impacts on rural health/alternate systems of medi-
cine. In: Sonak S (ed) Multiple dimensions of global environmental change, p. 170–181.
Ros-Tonen M, Dijkman W, Lammerts van Bueren E (1995) Commercial and sustainable extraction
of non-timber forest products. Towards a policy and management oriented research strategy.
The Tropenbos Foundation, Wageningen.
Saberwal VK, Gibbs JP, Chellam R & Johnsingh AJT (1994) Lion-human conflict in the Gir Forest,
India. Conservation Biology, 8(2), 501–507.
Saenger P (2002) Mangrove ecology, silviculture and conservation. Springer Science & Business
Media.
Sahana M, Sajjad H (2019) Vulnerability to storm surge flood using remote sensing and GIS
techniques: A study on Sundarban biosphere reserve, India. Remote Sensing Appl: Soc Environ
13: 106–120.
Saintilan N (1997) Above-and below-ground biomasses of two species of mangrove on the
Hawkesbury River estuary, New South Wales. Mar Freshwat Res 48(2): 147–152.
Sánchez-Triana E, Enriquez S, Afzal J, Nakagawa A, Khan AS (2014) Cleaning Pakistan’s air:
Policy options to address the cost of outdoor air pollution. The World Bank.
Sánchez-Triana E, Ortolano L, Paul T (2018) Managing water-related risks in the West Bengal
Sundarbans: Policy alternatives and institutions. Int J Wat Resour Develop 34(1): 78–96.
Santhakumar V, Enamul Haque AK, Bhattacharya R (2005) An economic analysis of mangroves in
South Asia. In: Khan M (ed) Economic development in South Asia. Tata McGraw Hill, New
Delhi, p 368–437.
Sanyal P (1992) Tropical Ecosystem: ecology and Management, p. 309–313, (K. P. Singh and J. S.
Singh eds, Wiley Eastern Ltd. New Delhi).
Sarangi SK, Maji B, Singh S, Burman D, Mandal S, Sharma DK, Singh US, Ismail AM, Haefele
SM (2015) Improved nursery management further enhances the productivity of stress-tolerant
rice varieties in coastal Rainfed lowlands. Field Crops Res 174: 61–70.
Schwamborn R, Saint-Paul U (1996) Mangroves-forgotten forests? Nat Resour Develop 43(44):
13–36.
Sen A, Pattanaik S (2017) How can traditional livelihoods find a place in contemporary conserva-
tion politics debates in India? Understanding community perspectives in Sundarban, West
Bengal. J Political Ecol 24(1): 861–880.
Sen HS (2019) Climate-risk Sundarbans needs multi-pronged and unified approach for ecological
sustenance a necessity for improved livelihood: Summary and concluding remarks. In: The
Sundarbans: A disaster-prone eco-region, Springer, pp. 611–625.
Shackleton C, Shackleton (2004). The importance of non-timber forest products in rural livelihood
security and as safety nets: A review of evidence from South Africa. South Afr J Sci 100(11):
658–664.
Simmons IG (1981) The ecology of natural resources, 2nd edition. Edward Arnold (Pub) Ltd,
16–19: p 436.
Singh OP (2002) Interannual variability and predictability of sea level along the Indian coast. Theor
Appl Climatol 72(1–2): 11–28.
Singh A, Bhattacharya P, Vyas P, Roy S (2010) Contribution of NTFPs in the livelihood of
mangrove forest dwellers of Sundarban. J Hum Ecol 29(3): 191–200.
Situmorang YA, Zhao Z, Yoshida A, Abudula A, Guan G (2020) Small-scale biomass gasification
systems for power generation (<200 kW class): A review. Renewable Sustainable Energy Rev
117: 109486.
Spalding M, Kainuma M, Collins L (2010) World Atlas of mangroves. Earthscan: London, UK,
p. 319.
STR (2018) Annual report 2017–2018. Conservator of forest & field director. Sundarban Tiger
Reserve. p 236.
16 Livelihood of Forest Dependent Dwellers in Relation to the Exploitation of. . . 399

STR (2014) Annual report 2013–2014. Sundarban Tiger Reserve. Retrieved from www.
sundarbantigerreserve.org/news/Annual%20Report%202013-14.pdf.
Tamai S, Nakasuga T, Tabuchi R, Ogino K (1986) Standing biomass of mangrove forests in
southern Thailand. J Jpn Fore Soc 68(9): 384–388.
Thakur S, Ghosh S, Ghosh T (2016) Scope and constraint of beekeeping in Satjalia Island within
Indian sundarban delta. Int J Agric Environ Res 2(1): 109–114.
Tomlinson PB (1986) The botany of mangroves. Cambridge University Press, London.
Twilley RR, Chen RH, Hargis T (1992) Carbon sink in mangrove and their implications to Carbon
budget of tropical coastal ecosystem. Wat Air Soil Pollut 64: 265–288.
Untawale AG (2006) Change of coastal land use, its impact and management options. Multiple
dimensions of global environmental change, TERI Press, New Delhi, India. p 23–43.
Upadhyay VP, Ranjan R, Singh JS (2002) Human-mangrove conflicts: The way out. Curr Sci 83
(11): 1328–1336.
Vyas P (2012) Biodiversity conservation in Indian Sundarban in the context of anthropogenic
pressures and strategies for impact mitigation (Doctoral dissertation). Saurashtra University.
Wackernagel M, Larry O, Linares AC, Falfan ISL, Garcia JM, Geurrero AIS, Geurrero MGS (1997)
Ecological footprints of nations: How much nature do they use? How much nature do they
have? (Commissioned by the Earth Council, New Society Publishers).
Whittaker RH, Likens GE (1975) The biosphere and man. In: Lieth H, Whittaker RH (eds) Primary
productivity of the biosphere, Springer-Verlag Ecological Series 14: 305–328.
World Bank (2014) Building resilience for sustainable development of the Sundarban. Strategy
Report. Report. No. 88061, Sustainable Development Department, Environment and Water
Resources Management Unit, South Asia Region.
Zohora FT (2011) Non-timber forest products and livelihoods in the Sundarbans. Rural livelihoods
and protected landscapes: Co-management in the Wetlands and Forests of Bangladesh.
Nishorgo Network, Dhaka, p. 99–117.
The Roles of Mangroves in Sustainable
Tourism Development 17
Yarina Ahmad and Mohd Nazip Suratman

Abstract

Mangrove and its ecosystem have contributed significantly to the environment,


community, and economy. Mangrove forests became a tourism attraction for
decades, and the demand for nature-based tourism of mangrove forests is increas-
ing throughout the world. Unsustainable mangrove tourism efforts may lead to
many drawbacks, which include forest loss, depletion of natural resources, and
also increase of pollution in mangrove forests and their surroundings. While
aiming to maximise the economic earnings from mangrove tourism, it should
be balanced with mangrove ecosystem conservation’s efforts to ensure sustain-
able tourism development. This situation can be realised through implementing
effective strategies, including continuous conservation and restoration efforts of
mangroves; enhancing the policies and legislation on the use and management of
mangroves; enhancing infrastructure and facilities, and efficient use of resources
through sound technologies; and encouraging community participation and
engagement. Balancing the economic activity of mangrove tourism while
maintaining mangrove forests through preservation, conservation, and restoration
are the key success to achieve sustainable tourism development.

Y. Ahmad
Faculty of Administrative Science and Policy Studies, Universiti Teknologi MARA (UiTM),
Shah Alam, Malaysia
Institute for Biodiversity and Sustainable Development, Universiti Teknologi MARA (UiTM),
Shah Alam, Malaysia
M. N. Suratman (*)
Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 401
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_17
402 Y. Ahmad and M. N. Suratman

Keywords

Mangrove · Tourism · Sustainable tourism development · Conservation

17.1 Introduction

Mangrove forests offer myriad benefits to the environment, community, and econ-
omy; however, conversely, they have received major threats from imbalanced and
destructive industrial development initiatives such as excessive aquaculture and
agriculture activities as well as high-scale urban expansion. While noting that
mangrove forests have contributed to the local socio-economy sector for centuries,
the demand for larger economic contributions through the tourism industry, for one,
has placed the mangrove ecosystem in jeopardy. Balancing between economic
earnings and mangrove ecosystem conservation efforts needs to be scrutinised to
ensure sustainable tourism development.
This chapter presents the roles of mangrove in ensuring the sustainability of the
coastal ecosystem; and the tourism industry needs to be enhanced by emphasising
precisely four critical strategies. Firstly, through the continuous conservation and
restoration efforts of mangroves to help achieve a sustainable tourism development
and further promote the 2030 Sustainable Development Goals (SDGs). Secondly,
there should be a focus on enhancing the policies and legislations of the use and
management of mangroves, particularly in highlighting the importance of regulating
mangrove forests. Thirdly, by enhancing infrastructure and facilities, efficient use of
resources through sound technologies can be ensured to promote environmental
cleanliness. Finally, community engagement and participation in utilising mangrove
forests as a tourism attraction should be encouraged in maintaining and conserving
the area. Thus, all parties must play their roles to help establish an environmentally-
responsible community. This can be done by educating and creating awareness
through various initiatives.
Mangrove conservation and restoration efforts alongside tourism development
initiatives can be initiated by optimising the potential and values of mangrove forests
and highlighting their economic and social benefits while minimising environmental
impact. This effort will ensure a genuine sustainable tourism development through
utilising mangroves as economic earning, as well as conserving the mangrove
ecosystem for future generations.

17.2 Mangroves

The academic discourse surrounding mangrove and its ecosystem has been
discussed globally for centuries (Bacon 1987; Suratman 2008; Webber et al. 2016;
Spalding and Parrett 2019). According to Webber et al. (2016), the word “man-
grove” refers to both a specific vegetation type and the unique habitat which is also
called tidal forest, swamp, wetland, or mangal. Mangroves are a group of trees and
17 The Roles of Mangroves in Sustainable Tourism Development 403

shrubs which are formed in swampy areas with low oxygen conditions prevailing
below the first few centimetres (Suratman 2008). Further, mangrove trees have a
special mechanism which allows them to survive water logged soil conditions and to
grow in saline and brackish water (Pons and Fiselier 1991).
Mangrove forests consist of mangrove trees which grow in a muddy coastal
wetland distributed worldwide in tropical and warm temperate coastal areas
(Hakim et al. 2017; Spalding and Parrett 2019). Notably, mangrove forests represent
transitional ecosystems where the ocean, land, and freshwater meet (Suratman
2008), which collect nutrients to create their own growing sides as well as forming
breeding grounds for shellfish and fish (Pons and Fiselier 1991). Mangrove forests
are unique and considered as a complex major component of coastal zones in the
tropical and sub-tropical regions (Suratman 2008) which play a major role in
environmental services, economy, and social benefits (Hakim et al. 2017).
Mangrove and its ecosystem have been contributed to the socio-economy devel-
opment of coastal community for many years. Traditionally, mangrove forests have
been utilised by the indigenous and local people for a variety of purposes. The
coastal community has been dependent on mangrove waters for centuries for fishes,
shrimps, crabs, mollusks (Suratman 2008). In recent times, the community continues
to rely on these products from the mangrove habitats not only for their daily living,
but also trading for economic development.
While human has gained many benefits from mangrove ecosystem for centuries;
they also have exploited these natural resources from various aspects. Mangrove
forests continue to be degraded rapidly through different types of human activities
(Saenger et al. 1983; Hamilton and Snedaker 1984). Large areas of mangrove had
been lost in recent decades to aquaculture, agriculture, and urban expansion
(Spalding et al. 2010; Richards and Friess 2016; Castellanos-Galindo et al. 2015).
Undoubtedly, mangrove degradation affects the ecological stability of coastal zones
(Suratman 2008); gives a negative impact on the coastal economy (marine resources
and fisheries industry) (Hakim et al. 2017); and depletion of the natural resources
(Amir 2019). These situations open debates among environmentalists and scholars
on the needs to protect and conserve mangrove forests through balancing the
economic development and ensuring the sustainability of mangrove forests and its
ecosystem.

17.3 Mangrove Tourism

Mangrove is now recognised as being among the most important ecosystem in the
world that provide various benefits to environment, community, and the economy
(Spalding and Parrett 2019). Recently, mangrove forests are among the popular
tourism attractions, particularly for nature-based tourism. This is not something new,
as mangrove ecosystem has been a tourism attraction for decades. Since the 1970s,
mangrove tourism was already found in the Caribbean (Bacon 1987). Since then,
mangrove tourism and attractions have been associated with the general trend of
ecotourism (Balmford et al. 2009; Cisneros-Montemayor and Sumaila 2010).
404 Y. Ahmad and M. N. Suratman

In one of the recent studies undertaken by Spalding and Parrett (2019), mangrove
forests are found as tourist’s attractions in many countries across the globe. The
article explored mangrove recreation and tourism worldwide through online search
of popular travel website TripAdvisor. In their study, which involved 3,985 man-
grove “attractions” located at 93 countries and territories, revealed that mangrove
tourism attracts tens to hundreds of millions of visitors annually; and thus, consid-
ered as a multi-billion industry.
Mangrove tourism is associated with other attractions including facilities,
activities, and wildlife (Spalding and Parrett 2019). Among the facilities provided
at mangrove tourism and attraction include a boardwalk, viewing tower, information
centre, and information boards. The activities offered in the mangrove tourism areas
are mostly related to boating, such as airboat, canoe and kayak, stand up paddle
boarding; and other activities such as fishing and hiking. Wildlife is also another
attraction associated with mangrove tourism; among them are birdlife, biolumines-
cence, fireflies, monkeys, proboscis monkey, manatee/dugong, crocodile/alligator,
and many other wildlife attractions (Spalding and Parrett 2019). These are among the
strategies to utilise mangrove forests as a tourism attraction undertaken by the
tourism industry to benefit the community and boost the economy. However, rapid
development of mangrove tourism may lead to drawbacks which are not only for
short-term, but also will give long-term impact to mangrove habitat and its
environment.
There are always two sides of a coin. Despite the widespread discussion on the
values and benefits of mangrove forests and its ecosystem; there are debates about
the loss of mangrove areas, and this happens in many countries across the globe
(Webber et al. 2016). Unsustainable use for economic development purposes on
mangroves has led to an alarming loss of global mangrove forests (WWF Interna-
tional 2017). Notably, the loss of mangrove forests has been reported since the 1980s
(FAO 2007; Polidoro et al. 2010). Nearly half of all mangrove forests have
disappeared since the mid-twentieth century (WWF International 2017). Further,
the loss of mangrove forests globally was reported at a mean rate of 1-2 per cent per
year (Duke et al. 2007; FAO 2007). In addition, the global loss rate of mangroves is
3 to 5 times higher than terrestrial forests (WWF International 2017). However, the
loss of mangrove forests has been declining over the last three decades (Spalding
et al. 2010). This could be an indication of increasing resilience of the remaining
mangroves or the result of effective conservation and restoration/rehabilitation
efforts (Webber et al. 2016).
Mangrove ecosystems provide a leading exemplar of the potentially high value to
be obtained from multiple ecosystem services (Salem and Mercer 2012); and thus,
provide a critical argument for both protection and restoration (Spalding and Parrett
2019). Utilising mangrove as tourism attraction and other species attraction might in
turn generate powerful arguments for conservation and management (The Econo-
mist 2018); in particular for preserving the environment for future generations.
Hence, four critical strategies should be undertaken to balance mangrove conserva-
tion efforts alongside tourism development in achieving a genuine mangrove sus-
tainable tourism development.
17 The Roles of Mangroves in Sustainable Tourism Development 405

17.4 Sustainable Development of Mangrove Tourism

Countries around the world are responsible in improving the planet and life of
citizens to achieve a better and more sustainable future. The Sustainable Develop-
ment Goals (SDGs) address the global challenges with the 17 Goals, which are all
interconnected, with the aim of no one left behind (United Nation 2020). In
promoting sustainable tourism development of mangrove, Goal 14 of the sustainable
development agenda is highly related. The aim of Goal 14 is to “conserve and
sustainably use the oceans, sea and marine resources for sustainable development”.
In this context, the world’s ocean—their temperature, chemistry, and life—drive the
global systems that make the earth habitable for humankind (UNDP 2020). The
world’s ocean covers mangrove forests and its ecosystem; and thus, through this
goal, there is a high requirement for mangrove forests to be protected, conserved,
and managed effectively, to benefit the community, society, as well as to preserve
the environment for future generations. Other than that, promoting sustainable
development of mangrove tourism also helps to achieve other SDGs including
Goal 1: “No poverty”; Goal 2: “Zero hunger”; Goal 8: “Decent work and economic
growth”; Goal 13: “Climate action”; and Goal 15: “Life on land” (Blum and Herr
2017).
Hence, in achieving a genuine sustainable development of mangrove tourism
through balancing the economic needs and protecting the mangrove forests and its
ecosystem; four critical strategies are emphasised. These four strategies are interre-
lated to each other—conservation and restoration of mangroves require a clear
policy and legislation to ensure effective management of mangrove tourism; subse-
quently, infrastructure, facilities, community participation, and engagement must be
enhanced—these strategies will support the agenda to achieve sustainable develop-
ment of mangrove tourism holistically. Figure 17.1 presents the four critical
strategies to achieve sustainable development of mangrove tourism.

Fig. 17.1 Promoting sustainable tourism development through four critical strategies
406 Y. Ahmad and M. N. Suratman

17.4.1 Continuous Conservation and Restoration Efforts


of Mangroves

In debating this issue, the first choice in hand is to preserve the mangrove forests. In
other words, no human activity is conducted at and/or nearby the mangrove sur-
roundings. In some remote areas, there are mangrove forests which are not being
utilised and exploited. However, once the mangrove forests are discovered, their
benefits will be taken advantage by human, community, and industry for a source of
food, income, and also increasing profit. Are utilising mangrove forests for a living
and boosting the economy considered wrong? There is a degree of continuum in
debating about this issue. Generating income activities by community and industry
can be considered acceptable if the mangrove forests are protected, and conservation
and restoration efforts of mangroves are done appropriately and continuously.
In recent trends, mangrove tourism is one of the earning strategies utilised by
many countries as well as to promote their culture (Hakim et al. 2017; Spalding and
Parrett 2019). While mangrove ecosystem is identified as among the most threatened
habitats on the Earth, and with limited research on mangrove and conservation
efforts; this resulted in a limited understanding of mangrove ecosystem among
many parties (Suratman 2008). In Indonesia, for instance, there are abundant of
mangrove forests; and their government highly encourages for tourism development
in natural areas, including mangrove ecosystem (Hakim et al. 2017). While
acknowledging that the mangrove forests in Indonesia are threatened, the country
and its community have begun to take efforts to restore mangrove ecosystem.
Among the activities conducted including corporate social responsibilities (CSR)
performed by companies focusing on rehabilitation of mangrove forests and its
ecosystem, and continuous meeting with various government officers to take actions
to protect the environment (Hakim et al. 2017).
In Malaysia, for instance, actions taken by the government to gazette all
remaining mangrove forests within forest reserves and protected areas are considered
as an effective initiative to preserve mangrove forests. Other initiative to conserve
mangrove forests is through devising well-balanced coastal land-use plans, such as
maintaining sustainable limits in logging and other harvesting activities of its
resources. The Malaysian government has acted in retaining protective mangrove
buffers along coastlines and rivers to prevent erosion; managing mangrove forests as
fishery reserves to encourage environmentally-sensitive commercial aquaculture
activities; and finally introducing the social forestry schemes where damaged forest
areas can be planted and managed for small-scale village timber enterprises
(WWF-Malaysia 2020).
In terms of restoration, mangrove planting is considered as a popular action;
however, it was not always effective (Wetlands International 2016). Mangrove
planting has been widely recognised following the Indian Ocean Tsunami in 2004.
Since then, mangrove planting has become very popular, and many parties, includ-
ing the government, non-governmental organisations (NGOs), private sectors, and
communities have planted mangrove trees, and many fund-raising activities were
undertaken to support the idea of mangrove planting. Although thousands of
17 The Roles of Mangroves in Sustainable Tourism Development 407

hectares of mangrove have been actively planted across the globe, most planting
efforts fail to effectively restore functional mangrove forests. Among the factors
leading to the failure of mangrove planting include lack of support from local
communities, mono-species planting leading to non-functional mangroves, planting
at the area where the recovered mangrove would block sediment and water flow,
planting at the area of original cause of loss, at the area where mangroves are settling
naturally, and at the area where not previously covered by mangroves (Wetlands
International 2016).
While taking experiences on the failure of mangrove planting across the globe,
Wetlands International (2016) suggested considering two principles to ensure suc-
cessful mangrove restoration. These include ensuring biography conditions are
appropriate for mangrove recovery and the socio-economic conditions allow for
mangrove recovery. These two principles are the cornerstone of the Ecological
Mangrove Approach developed by Lewis III (2005). This approach has been
successfully applied in mangrove planting in Indonesia, which was called
Community-Based Ecological Mangrove Restoration; and is considered as a best
practice to be adopted by other countries (Wetlands International 2016).
The conservation and restoration of mangrove forests is thus an important
contribution to the achievement of the Agenda 2030 of the United Nations and
therein defined SDGs (WWF International 2017). Further, restoring and protecting
mangroves help fulfil multiple SDGs, including SDG14 which focuses on sustain-
ably governing of the oceans and coasts and recognises the immense value of
mangroves to local communities; SDG1 and SDG2 on eliminating poverty and
hunger; SDG8 in ensuring livelihoods and economic growth; SDG13 taking actions
against climate change impacts; and SDG15 on halting biodiversity loss (Blum and
Herr 2017).

17.4.2 Enhancing the Policies and Legislations of the Use


and Management of Mangroves

The loss of mangrove forests across the globe and its impact to the mangrove
ecosystem and environment has led to the need for policy and legislation on the
use and management of mangrove to be enhanced. In some countries, mangroves are
protected through legislation that limits or prohibits mangrove clearing (Webber
et al. 2016). For instance, in Brazil, there is legislation on Brazil’s Federal Forestry
Code (Brazil 2012) which prohibits the use of any components of mangrove trees or
plants. In the United States of America, there is an act to regulate trimming,
disturbance, or removal of mangroves in the state, which is referred to as the
Mangrove Trimming and Preservation Act enacted in 1996 in the 1 United Nations,
Treaty Series¸ vol. 996, No. 14583 # 2016 United Nations 7 state of Florida
(Webber et al. 2016).
Notably, not all countries, states, or regions have a specific legislation to protect
mangroves. In Tonga, there is no policy and legislation specifically designated to
regulate the use, management, and conservation of mangroves (The International
408 Y. Ahmad and M. N. Suratman

Union for Conservation of Nature n.d.); hence, the country proposed for law reform
on mangrove conservation and management. Similarly, in Malaysia, there is no
specific policy and act for mangrove protection and management. The National
Forestry Act 1984 provides for the administration, management, and conservation
of forests and forest developments within the states in Peninsular Malaysia. Mean-
while, the Environmental Quality Act 1974 (Amended 1985) protects the forest
environment and biodiversity, in particular the logging of natural forests. In terms of
policy, there are two policies that cover environmental issues generally, which are
National Policy on the Environment, and National Coastal Resource Management
Policy (Abd Shukor 2004). Other countries that have no specific policy and regula-
tion on mangrove protection and the environment are Pakistan, Thailand, and
Vietnam (Beresnev et al. 2016).
The absence of a policy and legislation on mangroves can be considered as a
pertinent issue that hinders mangrove protection, conservation, and thus lead to
weak management of mangrove forests. This situation will further provide unclear
guidance to relevant parties who deal with mangrove forests, not only for tourism
purposes. At present, the countries that have no specific policy and legislation rely
on various relevant policies and legislation that cover mangrove generally under the
name of “environment”, “forests”, “land”, and many others. These will create
confusion and further may lead to a weak understanding of mangrove, its ecosystem,
and how to protect them. Hence, there is an urgent need for the respected countries to
take action to develop a specific policy and legislation as undertaken in Brazil and
USA. This will support a clear agenda to protect, conserve, and manage mangroves
effectively.
While Van Lavieren et al. (2012) has emphasised on the need to establish
framework policy and legislation for mangroves at the national level; the authors
also have suggested that laws and regulations related to mangrove protection,
conservation, and management must be enacted and enforced accordingly. Further,
there should be a framework of mangrove management, which emphasised on the
right of ownership, access and use of mangrove forest; and enhance human, techni-
cal, legal, and financial capacity for mangrove management at different level.
Among the management measures and tools to maximise the benefits and help
secure the long-term future of mangroves recommended by Van Lavieren et al.
(2012) include:

• Increase mangrove restoration;


• Increase community involvement in mangrove management;
• Implement sustainable mangrove forestry practices;
• Encourage sustainable aquaculture practices;
• Establish protected areas in ensuring the protection of mangrove biodiversity;
• Develop cohesive management plans;
• Promote managed realignment to (re-) establish landward expansion of mangrove
habitat;
• Encourage and support mangrove ecotourism to generate income and employ-
ment for local communities and to improve outreach and education;
17 The Roles of Mangroves in Sustainable Tourism Development 409

• Enhance existing carbon stocks and reverse CO2 emissions by increasing protec-
tion and restoration of mangrove ecosystems, and build mangroves into emissions
trading and climate change mitigation planning;
• Utilise multilateral environmental agreements, together with the establishment of
national legal protection measures, to support mangrove management.

Once the policy, legislation, and management framework of mangrove are in


place to ensure effective mangrove management and conservation, the way forward
agenda is to coordinate action towards the protection and restoration of mangroves
needs to be embedded within the international policy arena, notably under biodiver-
sity, wetlands, sustainable development, and climate change agreements. Aligning
the national agenda of mangrove with the international standards will ensure a
holistic achievement of SDGs, in particular to promote sustainable development of
mangrove tourism in the world.

17.4.3 Enhancing Infrastructure and Facilities, and Efficient Use


of Resources through Sound Technologies to Environmental
Cleanliness

The healthy environment of mangrove forests attracts local and international tourists
to visit and vacation. The forests are able to offer “nature therapy” for people who are
always busy with lives in cities and urban areas. Wildlife, panorama views, and
stress-free environment encourage people to experience life at mangrove forests.
This situation urges the local authority to open the door for tourists. Various efforts
and initiatives need to be undertaken to ensure there is an efficient use of mangrove-
related resources in promoting environmental cleanliness. As demands for visit and
vacation at mangrove forests are increasing, tourism development is also growing.
Due to this situation, infrastructure and facilities provided at mangrove forests must
be eco-friendly and suitable to prevent mangroves from destruction. Otherwise,
mangrove forests remain in extreme risks due to degradation, erosion, and the
extinction of wildlife.
In today’s globalised world, the tourism development is integrated with sound
technologies and development plans, especially in the complex area likes mangrove
forest. One of the rapid tourism developments that have taken place in mangrove
forests is resort construction. In maintaining the environmental cleanliness in the
mangrove forests, site planning of resort development is extremely important (Said
and Rahman 2000). Site planning must consider the estuarine or intertidal wetland to
conserve the relationship between physical and biotic factors. The conservation of
the relationship is vital to safeguard the food chains from being broken as well as to
protect the microclimate of the mangrove ecosystem. This huge responsibility
depends mostly on the ability and capability of an architect who is accountable for
the resort development. Further, the architecture planning designed at mangrove
forests must take into account the infrastructure and facilities related to intensive
human activity and building development. For example, vehicle parking, registration
410 Y. Ahmad and M. N. Suratman

Fig. 17.2 ‘Building with Nature’ activity at Timbul Sloko village, Indonesia (Source: Spalding
et al. 2014b)

area, outdoor games, and dining should be concentrated at the core building area to
protect mangrove-related resources. This planning concept can optimise the oppor-
tunity for guests to experience the serenity of the forest’s natural setting (Said and
Rahman 2000). In ensuring tourists are able to experience memorable events at
mangrove forests, comprehensive tourist guideline must be developed and designed
alongside the architecture planning to regulate tourists’ behaviours while at the
forests. This initiative helps the resort management in promoting environmental
cleanliness and further maintains an eco-friendly infrastructure and facilities for the
long-term usage as well as to conserve and protect the mangrove ecosystem.
The “Building with Nature” concept is another approach that can be considered in
promoting environmental cleanliness at mangrove forests (Spalding et al. 2014a,
2017). This approach has shown positive results by combining green (nature-based)
with grey (engineered) solutions. This initiative represents a remarkable shift in
environmental paradigm—from fighting against nature, to collaboratively cooperate
with natural processes. For instance, Northern Java’s coasts are suffering from
severe erosion whereby the areas are retreated by three kilometres. Improper devel-
opment of infrastructure and aquaculture disrupts the protection of mangrove forests
and the sediment flows towards the coast and floodplains. Due to this situation, a
group of researchers used “Building with Nature” approach to encounter the destruc-
tive erosion of the coast, particularly in Timbul Sloko village in Central Java. The
approach entails the placement of permeable dams that break the waves and trap
sediment, thus reclaiming land. Once the land is back, mangroves can recolonise the
area and help protect the coastline against erosion. The waves are clearly much lower
inside the grid of permeable dams than outside. In some cases, pioneering mangrove
trees are already becoming established. In a direct response to these initial tests, the
village signed a decree in 2014, establishing 100 hectares of the recently lost land as
protected area, ensuring that, upon recovery, it will not be damaged or destroyed
again. Figure 17.2 shows the activity of “Building with Nature” at Timbul Sloko
village (Spalding et al. 2014b).
In addition, mangrove forests and its restoration sites can also be strategically
placed to contribute to upgrading infrastructure with greater adoption of
17 The Roles of Mangroves in Sustainable Tourism Development 411

environmentally sound technologies through applying “green-grey infrastructures”


for coastal protection (Blum and Herr 2017). This is in line with the SDGs target to
upgrade infrastructure and retrofit industries to make them sustainable, with
increased resource-use efficiency and greater adoption of clean and environmentally
sound technologies and industrial processes, with all countries taking action in
accordance with their respective capabilities by 2030. These actions further build
the resilience of vulnerable coastal communities by reducing their exposure to
climate-related extreme events and environmental shocks and disasters (Blum and
Herr 2017). Similarly, the SDGs targeted to build the resilience of the poor and those
in vulnerable situations as well as to reduce their exposure and vulnerability to
climate-related extreme events and other economic, social, and environmental
shocks and disasters by 2030.

17.4.4 Encouraging Community Participation and Engagement

The benefits of mangrove forests bring immeasurable value to the local community
such as providing business opportunities, source of protein, and wild attraction
among tourists. For this reason, it is important to discuss the roles of local
communities in protecting and conserving mangroves forests to retain the long-
standing benefits of mangroves. The local community has been recognised as the
cornerstone in managing and protecting mangrove forests from extinction
(Valenzuela et al. 2020). The continuous protection and conservation of mangrove
forests does not only require the participation of local community; nonetheless, trust,
norms, and networks are also important in coordinating the actions which are
associated as “social capital”. This concept incorporated with two important aspects,
which are (1) it is created by building social relationships, and (2) it is a mechanism
to acquire more resources that can further offer more opportunities (Valenzuela et al.
2020). The combination of human capital and social capital helps in upgrading the
local community participation to the next level which is community engagement.
This situation further creates strong responsibility of local community in protecting
and conserving mangrove forests from harm and destruction.
The role of local communities in protecting and conserving mangrove forests is
closely discussed with Community-Based Conservation (CBC), which has been
developed as a viable alternative for sustainably managing the forests (Datta et al.
2012). This concept is aimed at empowering local community in protecting and
conserving mangrove forests (Walters 2004). Notably, there are no absolute
strategies and approaches outlined by the CBC. This is because different approaches
and strategies to achieve sustainability will be implemented by considering various
aspects including ecological, economic, and social sustainability in particular places
(Boyer et al. 2016).
However, there are various barriers that prevent the effectiveness of CBC in
promoting alternatives to manage mangrove forests, such as socio-cultural transfor-
mation of the local community (Datta et al. 2012). In today’s challenging world,
globalisation and modernisation have provided significant impact towards mangrove
412 Y. Ahmad and M. N. Suratman

forest management. Examples of poor forest management may lead to excessive


fishing activity in tropical coastlines among local community and others in the
destruction of the mangrove forests. The rapid tourism development also creates
countless problems towards the growth of mangroves. Again, this situation has two
sides of the coin, which are (1) devalue the role of the local community, and (2) open
for more opportunity of research and innovation in properly managing the forests.
Ironically, ineffective management of mangrove forests does not only affect the
sustainability of mangrove tourism; but it has been argued to bring the local
community back into poverty (Wesenbeeck et al. 2015).
In order to encourage participation and engagement among the local community,
several alternatives can be undertaken. One of the highly preferred alternatives
among the local community is by receiving incentives. However, debates related
to this aspect revealed that generating participation of the local community with
incentives is considered as challenging due to the difficulty in identifying their real
needs and preferences (Pons and Fiselier 1991). If the actions of local communities
in protecting and conserving mangrove forests are determined by incentives,
questions arise in the situation if and when the incentives are withdrawn. This
situation does not guarantee the continuity and successfulness of mangrove restora-
tion. While agreeing that incentives can act as “carrots” to motivate such behaviour;
however, incentives cannot create self-volunteerism behaviours among local
communities in protecting and conserving mangrove forests.
Another alternative that can be conducted to encourage local community partici-
pation and engagement is through implementing relevant regulations (Astuti et al.
2017; Majesty and Fadmastuti 2018). Mangrove is regarded as a threatened species
on the Earth, which requires legal enforcement (Official Website of Global Man-
grove Alliance n.d.). This initiative requires a clear direction from the government to
state and local authorities. Since mangrove forests are the meeting point of land, sea
water, and fresh water, the legal responsibility lies on various entities and agencies.
This situation creates conflict in the inter-agency cooperation between relevant
agencies which further reduces the effectiveness of policies and regulations
implemented (Beresnev et al. 2016).
In Thailand, for instance, mangroves are regulated and managed by various
agencies, including the Ministry of Natural Resources and Environment
(MoNRE), Ministry of Interior, Department of Fisheries, and Ministry of Agriculture
and Cooperatives (Beresnev et al. 2016). Historically, policies and regulations
pursued by the different entities have come into conflict, and the national decentrali-
sation process has also led to disruptions. Conflict has been related primarily to the
simultaneous pursuit of mangrove conservation and shrimp farm expansion,
although in recent years, the decline in the profitability of shrimp production has
reduced demands on land in mangrove areas (Memon and Chandio 2011). Further,
this situation weakens the community participation and engagement in protecting
and conserving mangrove forests.
The next alternative that can be undertaken has received less attention and
preference; yet, it is a powerful strategy in encouraging the community participation
and engagement in protecting and conserving mangrove forests. It is education and
17 The Roles of Mangroves in Sustainable Tourism Development 413

knowledge (Quarto 1999). There are a number of education programmes undertaken


by the NGOs across the globe about conservation and restoration of mangroves.
Among the organisations with projects around the world include the Mangrove
Action Project, Western Indian Ocean (WIO) Mangrove Network, the Mangrove
Alliance, and Mangrove Watch, as well as domestic organisations, including Honko,
mangrove conservation and education organisation in Madagascar, and the Man-
grove Forest Conservation Society of Nigeria, and many others (Webber et al. 2016).
These actions can be examples for other countries to continuously educate and create
awareness among relevant parties with direct or indirect responsibility to protect and
conserve mangrove forests particularly for tourism attractions.

17.5 Conclusions and Recommendations

Mangroves and its ecosystem undoubtedly have contributed numerous benefits to


environment, community, and the economy. Mangrove forests nowadays are among
popular tourism attractions across the globe. Their natural environment and wildlife
attractions offer economic opportunities for community, businesses, and developers
to take advantage for income generation and profit making. Ideally, the aim is to
achieve sustainable tourism development through balancing the economic activity of
mangrove tourism while maintaining mangrove forests through preservation, con-
servation, and restoration (Refer to Fig. 17.3).
While mangrove tourism help boost the country’s economy and promote their
culture; it also has its drawbacks, which include forest loss, depletion of natural
resources, and also an increase of pollution in mangrove forests and its surroundings.
Overemphasising on the economic aspect, as well as for the development and
urbanisation purposes will lead to an imbalance in sustainable tourism development.
This is particularly when mangrove preservation, conservation, and restoration are
not done appropriately and effectively. The situation continues to worsen when there
is no specific or clear policy, regulation, and ineffective management of mangrove
undertaken to utilise mangrove forests (Fig. 17.4). Additionally, many mangrove

Fig. 17.3 Sustainable tourism development (Ideal)


414 Y. Ahmad and M. N. Suratman

Sustainable Tourism Development (Reality)

• No specific policy and


legislation
• Weak mangrove
management
• Weak mangrove
restoration
• Lack of infrastructure
and facilities
• Low community
participation and
engagement
Development

Urbanisation
Tourism Preservation

Restoration
Imbalance
Conservation

• Economic pressure
• Loss of mangrove
forests
• Pollution
• Diminishing natural
resources

Fig. 17.4 Sustainable tourism development (Reality)

restoration or planting are also found ineffective due to inappropriate strategies and
actions. Further, when mangrove forests are utilised as tourism attractions, suitable
infrastructure and facilities are not provided. This contributes to other problems such
as pollution. Finally, low community participation and engagement in the manage-
ment of mangrove, as well as in promoting mangrove tourism hinders the effective-
ness of sustainable tourism development. Undermining these aspects creates gaps for
an imbalance of sustainable development of mangrove tourism.
Hypothetically, the ideal model of sustainable tourism development is the aim of
any mangrove tourism attractions. With this ideal model in mind, there is a need for
balancing the two aspects of economic pressure (tourism, development, and urbani-
zation) with the environmental aspect (mangrove preservation, conservation, and
restoration). This can be realised through implementing effective strategies: (1) Con-
tinuous conservation and restoration efforts of mangroves; (2) Enhancing the
policies and legislations on the use and management of mangroves; (3) Enhancing
infrastructure and facilities, and efficient use of resources through sound
technologies; and (4) Encouraging community participation and engagement.
17 The Roles of Mangroves in Sustainable Tourism Development 415

References
Abd. Shukor AH (2004) The use of mangroves in Malaysia. Promotion of mangrove-friendly
shrimp aquaculture in Southeast Asia . Aquaculture Department, Southeast Asian Fisheries
Development Center 136-144.
Amir AA (2019, May 17) Preserving mangroves and sustainable development. New Straits Times.
https://www.nst.com.my/opinion/columnists/2019/05/489425/preservingmangrooves-and-sus
tainable-development.
Astuti S, Muryani C, Rindarjono MG (2017) The community participation on mangrove conserva-
tion in Sayung, Demak from 2004-2016. IOP Conference Series: Earth and Environmental
Science 145.
Bacon PR (1987) Use of wetlands for tourism in the insular Caribbean. Ann Tour Res 14(1):
104-117.
Balmford A, Beresford J, Green J, Naidoo R, Walpole M, Manica A (2009). A global perspective
on trends in nature-based tourism. PLoS Biol 7(6): e1000144.
Beresnev N, Phung T, Broadhead J (2016) Mangrove-related policy and institutional frameworks in
Pakistan, Thailand and Viet Nam. Food and Agriculture Organizations of the United Nations
Regional Office for Asia and the Pacific International Union for Conservation of Nature.
Blum J, Herr D (2017) Can restoring mangroves help achieve the Sustainable Development Goals?
International Union for Conservation of Nature. Retrieved August 2, 2020. https://www.iucn.
org/news/forests/201703/can-restoring-mangroves-help-achievesustainable-development-
goals.
Boyer R, Peterson ND, Arora P, Caldwell K (2016) Five approaches to social sustainability and an
integrated way forward. Sustainability 8.
Castellanos-Galindo GA, Cantera J, Saint-Paul U, Ferrol-Schulte D (2015). Threats to mangrove
social-ecological systems in the most luxuriant coastal forests of the Neotropics. Biodivers
Conserv 24: 701-704.
Cisneros-Montemayor AM, Sumaila UR (2010) A global estimate of benefits from ecosystem-
based marine recreation: potential impacts and implications for management. J Bioecon 12:
245-268.
Datta D, Chattopadhyay RN, Guha P (2012) Community based mangrove management: A review
on status and sustainability. J Environ Manag 107: 84-95.
Duke NC, Meynecke JO, Dittmann S, Ellison AM, Anger K, Berger U, Cannicci S, Diele K, Ewel
KC, Field CD, Koedam N, Lee SY, Marchand C, Nordhaus I, Dahdouh-Guebas F (2007) A
world without mangroves? Science 317(5834): 41–42.
FAO (2007) The World’s Mangroves 1980-2005. FAO Forestry Paper No. 153. Rome, Forest
Resources Division, FAO 77.
Hakim L, Siswanto D, Makagoshi N (2017) Mangrove conservation in East Java: The ecotourism
development perspectives. J Tropic Life Sci 7(3): 277-285.
Hamilton LS, Snedaker SC (1984) Handbook for mangrove area management, IUCN/ UNESCO/
UNEP, Honolulu: East-West Center.
Lewis III R (2005). Ecological engineering for successful management and restoration of mangrove
forests. Ecol Eng 24(2005): 403-418.
Majesty KI, Fadmastuti M (2018) Degree of community participation in mangrove resources
management as livelihood support in West Java, Indonesia. International Conference Series
on Life Cycle Assessment: Life Cycle Assessment as a Metric to Achieve Sustainable Devel-
opment Goals (ICSoLCA 2018).
Memon J, Chandio A (2011) Critical appreciation of restoration and conservation of degraded
mangroves in Thailand. Int J Environ Rural Dev 2: 108-113.
Official Website of Global Mangrove Alliance (n.d.) The Global Mangrove Alliance. http://www.
mangrovealliance.org/wp-content/uploads/2017/08/global-mangrove-alliance_strategy.pdf.
Polidoro BA, Carpenter KE, Collins L, Duke NC, Ellison AM, Ellison JC, Joana C, Ellison JC,
Farnsworth EJ, Fernando ES, Kathiresan K, Koedam NE, Livingstone SR, Miyagi T, Moore
416 Y. Ahmad and M. N. Suratman

GE, Nam VN, Ong JE, Primavera JH, Salmo III SG, Sanciangco JC, Sukardjo S., Wang Y,
Yong JWH (2010) The Loss of Species: Mangrove Extinction Risk and Geographic Areas of
Global Concern. PLoS ONE 5(4): e10095.
Pons LJ, Fiselier JL (1991) Sustainable Development of Mangroves. Landsc Urban Plan 20(1-3):
103-109.
Quarto A (1999) Local community involvement in mangrove rehabilitation: Thailand’s Yadfon. In
Streever, WJ (ed.), An international perspective on wetland rehabilitation. Springer, Dordrecht
139-142.
Richards DR, Friess DA (2016) Rates and drivers of mangrove deforestation in Southeast Asia,
2000-2012. Proc Natl Acad Sci, USA 113(2): 344–349.
Saenger P, Hegerl EJ, Davie JDS (1983) Global status of mangrove ecosystems. The Environmen-
talist 3 (supplement 3).
Said IA, Rahman MA (2000) Site planning and architectural design considerations of resorts in
mangroves forest. Proceedings of the CASLE 2000.
Salem ME, Mercer DE (2012) The economic value of mangroves: A meta-analysis. Sustainability.
Spalding M, Kainuma M, Collins L (2010) World atlas of mangroves. Earthscan.
Spalding MD, Ruffo S, Lacambra C, Meliane I, Hale LZ, Shepard CC, Beck MW (2014a) The role
of ecosystems in coastal protection: Adapting to climate change and coastal hazards. Ocean
Coast Manag 90: 50-57.
Spalding M, McIvor A, Tonneijck FH, Tol S, Eijk VP (2014b) Mangroves for coastal defence:
Guidelines for coastal managers & policy makers. Wetlands International and the nature
conservancy 42.
Spalding MD, Burke L, Wood S, Ashpole J, Hutchison J, Ermgassen PZ (2017) Mapping the global
value and distribution of coral reef tourism. Mar Policy 82: 104─113.
Spalding M, Parrett CL (2019). Global patterns in mangrove recreation and tourism. Mar Pol 110
(2019).
Suratman MN (2008). Carbon sequestration potential of mangroves in Southeast Asia. In Bravo F,
Jandl R, LeMay V, Von Gadow K. (eds.), Managing forest ecosystems: The challenge of
climate change. managing forest ecosystems. Springer, Dordrecht 17: 297-315.
The Economist (2018). Money swinging from trees. The Economist Magazine. London.
The International Union for Conservation of Nature (n.d.). Review of policy and legislation relating
to mangroves, the use and management of mangroves in Tonga. The International Union for
Conservation of Nature. https://www.iucn.org/sites/dev/files/content/documents/Tonga_pol
icy_and_legislative_review_report.pdf.
United Nations (2020). About the sustainable development goals. Sustainable Development Goals.
https://www.un.org/sustainabledevelopment/sustainable-development-goals/
UNDP (2020). Goal 14: Life below water. United Nations Development Programme. https://www.
undp.org/content/undp/en/home/sustainable-development-goals/goal-14-life-below-water.html
Valenzuela RB, Youn Y-C, Mi SP, Chun J-N. (2020). Local people’s participation in mangrove
restoration projects and impacts on social capital and livelihood: A case study in the Philippines.
Forests 11(580).
Van Lavieren H, Spalding M, Alongi D, Kainuma M, Clüsener-Godt M, Adeel Z (2012). Securing
the future of mangroves: A policy brief, UNU-INWEH, UNESCO-MAB with ISME, ITTO,
FAO, UNEP-WCMC and TNC. London: Earthscan.
Walters BB (2004). Local management of mangrove forests in the Philippines: Successful conser-
vation or efficient resource exploitation? Hum Ecol 32(2): 177-195.
Webber M, Calumpong H, Ferreira B, Granek E, Green S, Ruwa R, Soares M (2016). Mangroves.
United Nations. https://www.un.org/Depts/los/global_reporting/WOA_RPROC/Chapter_48.
pdf.
17 The Roles of Mangroves in Sustainable Tourism Development 417

Wesenbeeck BV, Balke T, Eijk PV, Tonneijck FH, Siry HY (2015). Aquaculture induced erosion of
tropical coastlines throws coastal communities back into poverty. Ocean Coast Manag 116:
466-469.
Wetlands International (2016, November 7). Mangrove restoration: To plant or not to plant?
Wetlands.org. https://www.wetlands.org/publications/mangrove-restoration-to-plant-or-not-
toplant/.
WWF International (2017). ‘Mangroves – A Life-saving Coastal Ecosystem: Scaling up protection
and restoration for achieving SDGs’, https://www.mangrovealliance.org/wp-content/uploads/
2017/10/more-information-on-Save-Out-Mangroves-Now-.pdf.
WWF-Malaysia (2020). ‘Mangrove Forests’, https://www.wwf.org.my/about_wwf/what_we_do/
forests_main/the_malaysian_rainforest/types_of_forests/mangrove_forests/.
Aquaculture in Mangroves
18
Tengku Mohd Zarawie Tengku Hashim,
Engku Azlin Rahayu Engku Ariff, and Mohd Nazip Suratman

Abstract

Ever since man realized the benefits of mangroves, the habitat has been impacted,
but peaked in the twentieth century. Approximately 35% of the world’s mangrove
area was lost between 1980s and 1990s with deforestation rates ranging from 1%
to 8%. The major drivers of mangrove deforestation in recent times include
aquaculture, agriculture, urban expansion, forest product extraction, salt pond
conversion, and the oil and gas industry. The boom in the aquaculture industry
from 1970s onwards resulted in almost 28% of the habitat being lost in Asia
(Bangladesh, India, China, Thailand, Vietnam, and Indonesia) and South Amer-
ica (Ecuador, Brazil, Peru) but by country, wise losses ranged from 7% to 63%. In
South East Asia alone mangrove loss to aquaculture amounted to approximately
30% (1.66 million hectares). The total global economic value of mangrove loss to
aquaculture is amounted at US$3.78–17.01billion/year. Three types of organisms
are generally cultured in mangroves, namely fish, shrimp/prawns, and crabs.
Mangrove conversion to aquaculture is a response to food security which is
mainly to an increase in demand for protein and a decrease in marine capture
fisheries. This is however is not without costs, such as habitat destruction, loss of
ecosystem services, water quality reduction, exotic species introduction, and
disease.

T. M. Z. Tengku Hashim · M. N. Suratman (*)


Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]
E. A. R. Engku Ariff
Faculty of Applied Sciences, Universiti Teknologi MARA (UiTM) Pahang Campus, Bandar Tun
Razak Jengka, Malaysia

# The Author(s), under exclusive license to Springer Nature Singapore Pte 419
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_18
420 T. M. Z. Tengku Hashim et al.

Keywords

Mangrove · Aquaculture · Ecosystem services · Food security · Livelihoods

18.1 Introduction

The growth of the world population in the past century does not receive much
attention from researchers, as it does not consider as a major threat to the human
survivors until the end of the twentieth century. However, since the exponential
growth of the human population exceeded 1 billion in the nineteenth century and
world population growth projection is expected to reach 9.3 billion in the year 2060,
uncertainties and the impacts of the growing population are started to emerge and
visible (Goujon 2019). The significant increase in human population will put
irreversible pressure on natural ecosystems as demand for the increase of natural
resources in order to satisfy the consumption needs for the human population. One of
the problematic scenarios that need to be considered by human throughout this era as
consequences from the increase in human population is to address the enormous gap
on how the world solve the arising issues in maintaining food security.
Since the great migration and cultural changes of the human race that took place
in the mid-Holocene, the coastal area has always been the focal place for this
purpose. With an increase in population density, the coastal area holds countless
benefits that always generate interest from the human perspective to live and exploit
the availability of resources in the area especially the mangrove ecosystems.
Mangroves are recognized as a group of trees, shrubs, and palms that can endure
and survive in harsh coastal conditions and can easily be distinguished by their
unique characteristics in the coastal line of subtropical and tropical areas across the
globe (Giri et al. 2011). The adaptability of mangroves to survive in a harsh
condition is tremendously admired as it can be considered as the dominant ecosys-
tem in the coastal area as compared to the other types of vegetation.
The impacts of saline seawater, tidal inundation period, and elevation play a
major role in the distribution, zonation pattern, and growth rate of mangroves (Win
et al. 2019). To endure the rigorousness in surviving in an unfriendly environment
such as in the coastal climate, mangroves develop numerous adaptation strategies.
One of the important morphological adaptations of mangroves is by possessing a
unique root structure that functions as a stabilizing mechanism that spreads and acts
like cable roots that anchor the mangroves to the soft muddy soils. Perhaps the most
remarkable root adaptations of the mangroves are the stilt roots of Rhizophora, the
pneumatophores of Avicennia, Sonneratia, and Lumnitzera, the root knees of
Bruguiera, Ceriops, and Xylocarpus, and the buttress roots of Xylocarpus and
Heritiera (Kathiresan and Bingham 2001). The root system of mangroves does not
only act as a stabilizer, but also function as a gas exchange and breathing tool for
mangroves to survive in the muddy and anaerobic place (Suratman 2008; Naidoo
2016). To cope with the saline environment, mangroves develop certain mechanisms
to excrete the excessive salt caused by the seawater. For example, Avicennia marina
18 Aquaculture in Mangroves 421

secretes the excessive salt by a special salt gland that exists on both of its leaf surface
(Krauss and Ball 2013) while other species such as Rhizophora, Bruguiera, and
Ceriops possess ultra-filters in their root systems that filter the excessive salt while
maintaining water uptake from the soil (Kathiresan and Bingham 2001).
Despite existing in the extreme coastal climate, mangroves are considered as
natural jewel to the coastal ecosystem in terms of providing crucial ecological
functions and pivotal roles in socioeconomic needs. Mangroves play a remarkable
role in sheltering and reducing the damage from the impacts of the tsunami and
strong wave-current in the coastal areas while offering protection to the coastal
communities (Ahmadun et al. 2020). Mangroves also provide essential benefits in
stabilizing coastal banks from erosion that is caused by strong tidal wave by trapping
and binding sediment through its unique root systems (Sánchez-Núñez et al. 2019).
Another crucial ecological function provided by the mangroves which could be a
key component in mitigating the ever-challenging global climate change is their
ability to sequester carbon and store it as part of their biomass for a long time period
(Suratman 2008; Hashim et al. 2015, 2020). In terms of socioeconomic values,
mangroves are considered as an important ecosystem to the coastal community by
providing woods for timber, charcoal, and other commercial products (Abdul Aziz
et al. 2015). Moreover, mangroves are proven to be an important ecosystem for
commercial fishing and coastal aquaculture industries as they home many high grade
marine habitats (Hutchison et al. 2014). Nevertheless, mangroves also provide job
opportunities like generating household income to the local communities through
coastal aquaculture industries and fishing activities in the mangroves.
Aquaculture activity is of the fastest growing industries on the planet as it
contributes to the world food production system. As human population trend is
exponentially increased in the upcoming decade, aquaculture seems to be one of the
important solutions in maintaining, reducing the gap, and fulfilling the increasing
demand for food supply. According to FAO (2020), it was estimated that in 2018,
the global fish production had reached 179 million tonnes where aquaculture was
among the highest contributors for this production with 82 million tonnes which was
valued at USD 250 billion. As things stand right now, aquaculture is a more
favourable technique to be used to produce mass food products as compared to the
traditional method which is becoming less effective for supplying fishery resources
to the market in the next decade (FAO 2018, 2020). With aquaculture industry
gaining pace to expand even further, the coastal area especially mangroves always
attract the attention of the governments, entrepreneurs, and local community
throughout the globe as an area that has humongous potential to be transformed
into an aquaculture hub. Furthermore, the condition of mangroves clicks all the
necessary criteria such as brackish water, water quality suitability, and protection
against natural elements such as storms and tsunamis (Hutchison et al. 2014;
Ahmadun et al. 2020), adding some additional values why this area can be a
successful aquaculture farming hub.
Despite the importance of mangroves to the coastal communities in providing
numerous ecological and socioeconomic supports, this ecosystem has degraded at an
alarming rate mainly due to global climate change and anthropogenic activities.
422 T. M. Z. Tengku Hashim et al.

Although previous studies indicated that the loss of mangroves worldwide are
caused by unprecedented climate change such as rising of sea level, storms, and
tsunamis as well as the increase of sea currents (Richards and Friess 2016; Das and
Mandal 2016) throughout the decade, evidence suggested that rapid and uncontrol-
lable anthropogenic activities especially the aquaculture activities and land clearing
are the main culprits for the reduction of mangroves worldwide (Truong and Do
2018). The unsustainable management and high conversion rate of mangroves to
aquaculture pond throughout the decade have left the mangrove ecosystem in
complete chaos. According to Valiela et al. (2001), in the past decade, mangroves
had reduced to half of their size in Southeast Asia, South and Central America to
make way for fish and shrimp farming.
Given the massive impacts of aquaculture activities to the mangroves and adja-
cent communities, it is crucial to find a solution to address this problem in order to
ensure the sustainability and survival of mangrove ecosystems. This chapter aims at
reviewing the roles of mangroves in supporting the aquaculture industries. The
benefits and impacts of aquaculture industries to the mangrove ecosystem and the
surrounding community are also highlighted.

18.2 The Role of Mangroves in Fisheries Habitat

Mangroves provide unique ecosystem functions that exist between the land and the
sea as they consist of a rich assemblage of biodiversity. The unique ecosystems are
not only crucial in providing a sustainable environment for many marine species, but
also contribute to important commercial fisheries resources such as fishes,
crustaceans (crabs and shrimps), and molluscs, which are sources of protein and
income for the local community (Hutchison et al. 2014; Barbier et al. 2011).
Mangroves play an essential ecological function to a large proportion of aquatic
species and marine habitat by serving as a nursery ground, good breeding ground
and act as shelter for juvenile and small fishes from predation (Hutchison et al.
2014). The effective and complex structure of mangrove habitats provides shelter for
the aquatic organism from predation thus creating more microhabitat availability and
increasing the amount of food (Mason et al. 2005).
Mangroves play a crucial role as the basis of the food chains that support a wide
range of marine habitats in the coastal area. The high level of primary productivity
from the mangrove vegetation such as litter, branches, and trunks as well as other
primary producers is important in establishing a complex food web (Hutchison et al.
2014). With mangroves being considered as one of the highly productive ecosystems
(Alongi 2012; Hutchison et al. 2014; Kamruzzaman et al. 2017), the main source of
the primary productivity that acts as the base of the coastal ecosystem food chain
comes from the mangrove trees, periphyton (algal that grows on tree roots), and
phytoplankton (Hutchison et al. 2014). Meanwhile, other additional sources of food
that enrich the mangroves come from adjacent ecosystems such as rivers, coral reefs,
and sea grass by depositing material from upstream rivers and carrying material from
the ocean by wave action (Igulu et al. 2013). In a study to determine the usage of
18 Aquaculture in Mangroves 423

mangroves by juvenile fish, Verweij et al. (2006) found that species richness of
herbivore juvenile fish such as Acanthurus chirurgus and A. Bahianus was more
abundant in the mangroves due to the presence of algae that grow on top of the
mangrove roots. Another study that was conducted in Ryukyu Island, Japan in an
effort to study the dependence of fish on mangroves also reported that the abundance
of fish that can be found in the mangroves is due to the presence of food and shelter
(Pantallano et al. 2018). Both of these studies are consistent with the hypothesis
highlighted by Nanjo et al. (2014a), where the greater availability of food that
accumulates in mangroves attracts fish and other marine habitats to the area.
Another important role provided by the mangroves in supporting fish habitat is by
providing a suitable and safe area for juvenile fish and small fish against predators.
The unique structures of mangroves such as the complex prop root system and
canopy shades that exist from the combination of branches and leaves form a
strategic hideout spot from predators (Nanjo et al. 2014b; Hutchison et al. 2014;
Verweij et al. 2006). This comes to an agreement with a study conducted in South
East Queensland, Australia using an artificial mangroves structure in a tank which
concluded that abundance of marine fishes, especially the juvenile tends to be higher
in a place that have more complex structure such as the mangroves in order to limit
the risk of predation (Laegdsgaard and Johnson 2001). Another study also suggested
that mangroves are used by small and juvenile fish as a refuge strategy by staying in
mangrove’s border for feeding purposes and flee into the mangroves when a large
predator approaches them (Sheaves et al. 2016). Pantallano et al. (2018) in their
study also revealed that fish species such as Apogon amboinensis, Neopomacentrus
taeniurus, and Pomacentrus taeniometopan are the highest species of fish that can be
found in the study area where they depend largely on the complexity of mangroves’
roots and structures in order to seek protection against bigger predators. More recent
studies conducted by Bradley et al. (2019) in Hinchinbrook, Northern Australia
found that complex mangrove ecosystem in the coastal water provides a refuge area
and a predominant habitat for coastal and reef juvenile fish species such as the
Acanthopagrus pacificus, A. australis, Lutjanus argentimaculatus, L. russellii,
Epinephelus coioides, and E. malabaricus. This is consistent with the main findings
in a study that was conducted in Nahoon Estuary, South Africa, which compared the
fish abundance in both salt marsh and mangroves indicated that fish species richness
were higher in the mangroves as compared to salt marshes due to the complex
structural and ecological benefit of the mangroves (Keur et al. 2019). Shrimps also
utilize the mangroves as a place for seeking protection against predation. According
to Rönnbäck et al. (1999) shrimps such as Penaeus merguiensis use the complex
structure of mangrove roots as a place for seeking shelter from predators. Further-
more, the study found out that the abundance of mangroves in the coastal area could
be the key in increasing the density of shrimp as it reduced the rate of predation while
the murky water in the mangroves during tide helps in reducing the visibility of
predator to prey.
Changes of water level in the mangrove forests which are directly influenced by
the tide activity have created a variation of the water level that helps the aquatic
species to navigate freely from an open water into parts of the forest thus enabling
424 T. M. Z. Tengku Hashim et al.

them to receive the benefits from the mangroves (Truong and Do 2018; Hutchison
et al. 2014). The complex structure of the mangrove trees that directly come in
contact with the water generates friction that slows the water speed in the mangroves
and creates an ideal habitat for small juveniles fishes and prawns (Hutchison et al.
2014). These kinds of conditions create an ideal environment for breeding and
nursery grounding for many commercial corals and freshwater fish where
mangroves’ structure traps fine sediment and creates a soft muddy floor that is
crucial for larvae and egg to settle down in the mangrove environment and eventu-
ally will increase fish and invertebrate’s population. According to Jamizan and
Chong (2017), a study conducted in five different mangrove areas across West
Peninsular Malaysia, Malaysia revealed that the geomorphological condition of
mangroves which is connected by creeks has a strong relationship with abundance
and distribution of fish (families of Lutjanidae, Haemulidae and Serranidae) and
shrimp (penaeid shrimps). Furthermore, the study also stated that a large mangrove
area that had many upstream rivers could be an ideal place for nursery grounding for
many high commercial fishes and shrimps. This comes to an agreement with a study
conducted by Tanaka et al. (2011) in Matang Mangrove Reserve, Malaysia where
they found out that large mangrove areas play an important role in life cycle stages
for many juvenile Lutjanus johnii. The study also revealed that the young juvenile of
L. johnii could migrate up to 13 km upstream from the coastal area towards the
mangroves using rich connective rives and were able to shift their diet from coastal
food web to mangroves’ dependence food web. El-Regal and Ibrahim (2014) also
concluded that mangroves are the essential ground for nursery area where they found
a large number of Gerres oyena and Sardinella maderensis juvenile during their
study in three coastal mangroves area along the Egyptian Red Sea.

18.3 Aquaculture Activities

Aquaculture is believed to be the fastest growing and profitable industry around the
globe that is responsible for supplying food resources to satisfy the growing needs of
human food consumption (FAO 2018, 2020; Ahmed and Thompson 2019). The
intense growth of cultivation of aquatic animals (fishes, shrimps, mussels, and
oysters) sometimes plants (seaweed) in marine, brackish, and fresh water worldwide
are considered as a blue revolution (Bavington and Banoub 2016; Ahmed and
Thompson 2019). The blooming of the aquaculture industry throughout this decade
is corresponding with the meteoric increase of world population and shifting pattern
in human consumption behaviour towards fishery products. According to FAO
(2020), the pattern of global fish consumption by human had outpaced the rate of
annual world population growth (1.6%) which saw an increased rate of 3.1% from
1961 to 2017. In terms of per capita fish food consumption, the rate of increase was
about 1.5% per year, which grew from 9 kg in 1961 to 20.5 kg in 2018 (FAO 2020).
The world of the aquaculture industry has contributed significantly in minimizing
the gap between food supply and demand for human consumption. As mentioned
previously, the aquaculture industry contributed about 82 million tonnes in 2018
18 Aquaculture in Mangroves 425

where the industry was dominated by finfish (54.3 million tonnes), molluscs, mainly
bivalve (17.7 million tonnes) and crustacean (9.4 million tonnes) (FAO 2020). With
aquaculture activity seems to be one of the major tools to address the food security
scenario, rapid expansion of aquaculture farm is visible throughout the globe.
Mangroves are one of known areas which are synonym with aquaculture activities
that are rapidly explored worldwide for the purpose of coastal aquaculture activity
(Hutchison et al. 2014; Ahmed and Glaser 2016; Truong and Do 2018). Even though
coastal aquaculture is significantly practiced all over the globe, this type of activity is
more concentrated in South, Southeast and East Asia, and Latin America (FAO
2018, 2020). The condition of mangrove environment and suitable climate in those
regions can be an ideal place to cultivate and practice the aquaculture activity.

18.3.1 Mangrove Ecosystem Services and Aquaculture Activities

Mangrove ecosystems play a tremendous role by providing numerous ecological


supports that give an impact towards a successful coastal aquaculture activity. The
role of mangroves in anticipating and reducing the impacts of tsunami and storms in
the coastal area can reduce the damage and ensure the safety of aquaculture products
and structures. Tsunamis and storms could have a devastating impact on aquaculture
activities as it increases the water level in the area which could cause over flow to
the aquaculture pond that might alter the salinity level and eventually will affect the
growth and production of the aquaculture products (Nguyen et al. 2018). The
catastrophic cyclone Sidr occurred in 2007 that hit Bangladesh had resulted in
fatal impact to shrimp farming in the local area where the loss was about USD
36 million (IRIN 2008). According to Sarker and Azam (2007), the catastrophe of
cyclone Sidr had resulted more than 90% loss of aquaculture farm in the regions of
Morelgonj and Saronkhola Upazilas. Even though mangroves do not have the
capability to captivate all the impacts of storms and tsunamis, the complex structure
of mangroves can help in reducing the turbulence impact from storms and tsunamis
thus benefitting the aquaculture industry in the area (Truong and Do 2018). The
complex root structure of mangroves along the aquaculture site could be important in
protecting the aquaculture pond by shielding and binding the soil from wave action
thus reducing the erosion problem.
Mangroves are considered as an ecosystem that has high biodiversity that consists
of many organisms that inhabit the area. As mentioned previously, mangroves are
strategic place for many aquatic habitats for nursery ground, feeding area and
seeking refuge from predators (Hutchison et al. 2014). The abundance of organisms
in the mangroves will benefit the aquaculture product (fish or shrimp) by providing
additional nutrients for feeding. According to Nagelkerken et al. (2008), after the
process of spawning finish, the eggs dispersed all around mangroves and after a
certain period of time, they would turn to planktonic larvae. The planktonic larvae
would move or carried by currents into other parts of mangroves or in the aquacul-
ture cage in the mangrove area that could be an additional source of food for other
aquatic habitats as well as for fish and shrimp in the aquaculture cage.
426 T. M. Z. Tengku Hashim et al.

The complex biodiversity of mangroves does not only provide additional food to
fish or shrimp in aquaculture ponds, but also provide natural production of larvae
and juveniles. Even though juveniles and larvae nowadays are spawned using
breeding technique in the captivity facilities, the dependent on mangroves in
providing wild seeds for the aquaculture industry is largely practiced in many
countries. According to Beveridge et al. (1997), the aquaculture industry, especially
shrimp farming was still largely depending on wild caught seeds or gravid females
for the aquaculture industry. Rönnbäck et al. (1999) stated that the natural produc-
tion of larvae and juveniles that scattered abundantly in mangroves is an important
source of seeds for aquaculture industry which they entered naturally by flowing
current or artificially caught. In addition, the ecological functions of mangroves in
providing suitable climates such as sustaining water quality mitigate variation of
salinity and turbidity also influence the growth rate and survival of aquaculture
product.
According to Venkatachalam et al. (2018), a study they conducted in Tamil Nadu,
India to determine the survival and growth of Lates calcarifer found out that the
integration between mangroves and aquaculture produced a higher rate of survival
(11%) and had a higher growth rate (12.5%) as compared to an aquaculture area that
did not have mangroves. In Vietnam, mangroves are regarded as an important
ecosystem for integration with aquaculture activity due to their ecological functions
provided. According to Vietnamese and international experts, the integration
between mangroves and aquaculture activity does not only improve the resilience
of aquaculture product, but also reduce disease outbreak because mangroves act as a
bio-filtering entity that improves water quality and buffers the impact of temperature
fluctuation (Joffre et al. 2015).

18.3.2 Importance of Aquaculture Activities

18.3.2.1 Job Opportunities


The coastal areas always attract the attention of human beings due to the attractive
socioeconomic function they offered. It is believed that the landscape of coastal area
is immensely altered from time to time due to vigorous land use/cover changes that is
caused by human growth and pressure, especially in the mangrove areas. According
to Thomas (2017), a total of 120 million people worldwide occupied the mangroves
and highly depending on mangroves’ resources for survival. However, the
communities that live in a coastal area are always regarded as communities
surrounded by poverty. The blooming of the aquaculture industry in recent decade
benefits the local community by creating a positive impact on the employment sector
where it generates business opportunities and job creations hence boosting the
socioeconomic livelihoods for the surrounding coastal community (Mialhe et al.
2013; Hussain et al. 2018). It was estimated that in 2018, a total of 59.51 million of
people was directly involved in the fishery sector where from that number, 20.58
million of people engaged in aquaculture sector (FAO 2020). The rapid growth of
the fishery sector, especially the aquaculture industry in the Asia region has resulted
18 Aquaculture in Mangroves 427

in an increase in job opportunity where 85% of the jobs in the fishery sector are
concentrated in this region, followed by Africa (9%), America (%), Europe (1%),
and Oceania (1%) (FAO 2020).
According to a study conducted by Mialhe et al. (2013), in their effort to
determine the impact of shrimp farming to land use, employment and migration in
Peru found out that there was a significant increase in job employment due to the
extensive shrimp farming activities in the area. The study revealed that during the
year 2001, only 439 people was directly involved in aquaculture activity. However,
with the economic revenue that was brought by this industry in terms of household
income, a total of 2660 people opted to join this industry as a full-time job in 2006.
In Bangladesh, the blue economy phenomenon, especially aquaculture industry
holds enormous potential in generating job opportunities for local coastal commu-
nity and this could hold the essential key in reducing the shackle of poverty within
this community (Hussain et al. 2018). According to Moni et al. (2018), in
Bangladesh alone a total of 600,000 people were working in the shrimp aquaculture
sector and benefited almost 3.5 million dependents. A study conducted in
Bangladesh concluded that the aquaculture sector through shrimp farming produc-
tion in the county has benefited almost 4.8 billion households by providing jobs to
1.2 million of people (Islam 2008). A similar trend of interim job opportunity was
observed in Vietnam after the introduction of shrimp aquaculture activities. It was
reported that the rapid production of aquaculture products in the CA Mau province
had resulted an exponential increase in aquaculture workers from 85,000 in 1997 to
312,000 in 2003 (Moni et al. 2018).
In Malaysia, the aquaculture sector has received much attention from the govern-
ment as it holds a potential economic gain as well as one of the alternative measures
to improve the economic gap for rural society especially the coastal community. One
of the alternatives that is implemented under the government policy (National Agro
Food Policy NAP (2011–2020)) in conjunction with the 10th Malaysian Plan is to
increase the production of aquaculture activities and products thus create and
increase job opportunities in this sector (Yusoff 2014). According to Yusoff
(2014), in 2012 the involvement of people in the aquaculture sector in Malaysia
was 29,494 and from this total about 22.7% or 6715 were directly involved in
aquaculture which was concentrated in mangrove areas.

18.3.2.2 Food Security


One of the main important criteria of the aquaculture industry towards human
sustainability is minimizing the gap for food consumption (FAO 2018, 2020). The
importance of aquaculture production in providing a source of food cannot be
ignored as it contributes to almost 50% of world fish consumption alongside deep
sea and inland fish capture (Barua and Rahman 2020). Recent advancement and
evolution of aquaculture technology and vast global market trade in recent years
have resulted in an increase in the production of aquaculture products and have a
significant impact to the consumption rate of human worldwide. According to FAO
(2020), in 2018 the per capita of fish consumption worldwide was about 20.5 kg, and
this figure had outpaced other types of animal protein consumption products such as
428 T. M. Z. Tengku Hashim et al.

milk, meat, egg, etc. This scenario is significantly impacted by the recent technolog-
ical development in processing, cold chain transport as well as diverse shipping and
distribution hub that are efficiently used to exploit the high market demand for
aquaculture products in the United States of America (USA), China, and the
European Union (EU) (FAO 2020). Moreover, with fast changing and improvement
in aquaculture distribution chain, it can have a significant increase to human fishery
consumption of 36 kg per capita by the year of 2020 (FAO 2016).
According to Pradeepkiran (2019), it was estimated that a total of 1 billion people
worldwide are largely depending on fish products. The high quality source of
proteins, essential amino acids, vitamins, and minerals that exist in fishery products
provide 16% of animal protein to human diet globally due to the abundance of
resources and cheaper as compared to other types of animal protein in the market
(Pradeepkiran 2019). According to Kawarazuka and Béné (2011) in 2010,
22 countries that could be categorized into low income and food deficient countries
were highly dependable on fish product as the source of animal protein in their diet.
In Malaysia, it was estimated that the fish consumption index among Malaysians is
expected to increase from 53.1 kg in 2011 to 61.1 kg in 2020 due to the shifting
patterns of diet for more healthy food based on fish protein (Yusoff 2014). The
increasing pattern of consumption that is based on fish protein among Malaysians is
also resulted from its growing population where the demand for fish products is
expected to increase from 1.7 million tonnes in 2011 to 1.93 million tonnes by 2020
(Yusoff 2014).

18.3.2.3 Aquaculture Monetary Benefits


The aquaculture sector is considered as one of the highest profitable business in this
decade that has a high market demand throughout the globe. According to the new
statistic reported by FAO (2020), the aquaculture sector has recorded another
milestone where in 2018 the aquaculture industry production produced an astonish-
ing record of 114.5 million tonnes in live weight which was equal to the sale value of
USD 263.6 billion. From the total aquaculture production, finfish was the highest
producer with 53.3 million tonnes or USD 139.7 billion where it harvested from
inland aquaculture (47 million tonnes, 104.3 billion) and marine and coastal aqua-
culture (7.3 million tonnes, USD 35.4 billion) (FAO 2020). In Southeast Asia, the
aquaculture sector is considered as an important economic activity and source of
income to the nation in this region as it contributed to almost USD 10.87 billion in
2017 as compared to 0.31 billion in 1984 (FAO 2018).
In Bangladesh, aquaculture activity, especially shrimp farming is the second most
important export of goods that significantly contributes to the economy of the nation.
The shrimp farming industry in Bangladesh is highly influenced by the high market
demand from developed countries such as the USA, United Kingdom (UK),
Belgium, Germany, and Japan due to the high quality grade of shrimp species
such as Penaeus monodon that cultivate in Bangladesh mangroves (Islam and
Bhuiyan 2016). According to Uddin et al. (2013), it was estimated that the export
of shrimp in Bangladesh for the years of 2011–2012 was about USD 449.56 million
as compared to export for years of 2001–2002 which was about USD 276.11 million.
18 Aquaculture in Mangroves 429

In Malaysia during the year of 2012, brackish water aquaculture activity, especially
shrimp farming activity (P. monodon and P. vannamei) contributed to
139129.51 tonnes of product that worth USD 0.39 million to the economy (Yusoff
2014). Being blessed with an abundance of resources, Malaysia considers aquacul-
ture as one of the important sectors that can contribute to the development of the
nation. According to Queiroz et al. (2013), with the financial benefit gained from the
aquaculture industry, it has transformed Brazil’s approach on aquaculture activity
from the experimental production of shrimp farming to an extensive industry that
was based on framework of rapid expansion and highly profitable. During this
booming era of shrimp aquaculture, this approach has resulted Brazil to produce
approximately 61,000 tonnes of shrimp production that valued USD 244.5 million
in 2013.
With coastal aquaculture activity being regarded as an important source of
income to the nation while producing economic activity for the coastal community,
it is expected that this sector will give a huge impact to the coastal ecosystem
especially the mangroves. By acknowledging the strategic and suitability of
mangroves in coastal area as a hub for aquaculture activity, it can be estimated
that more mangrove areas need to be cleared and converted to make way for these
industries. Many studies reported that this activity contributed to the largest man-
grove losses elsewhere in the world (Romañach et al. 2018; van Wesenbeeck et al.
2015; Richards and Friess 2016; Ottinger et al. 2016). The next section in this
chapter will highlight the impacts of aquaculture activity to mangrove ecosystems.

18.4 Impact of Aquaculture in Mangroves

During the last few decades, aquaculture has benefited from scientific progress and
the growth has been exponential. This increasing trend will need to remain in order
to support the demand of the food supply of more than 9 billion people by the year of
2050. Unfortunately, upon the successful aquaculture, there have also been undesir-
able social and environmental issues, including unsustainable water or natural feed
use, mangrove destruction, and biodiversity loss. According to Vo et al. (2013), the
destruction of mangrove forest is due to the development of settlements, aquaculture
and agriculture.

18.4.1 Eutrophication

Nutrient enrichment is one of the most serious threats to the coastal ecosystem
(Downing et al. 1999; Cloern 2001). Previous studies have shown that growth of
intertidal mangrove forests is stimulated with enhanced nutrient availability; how-
ever, the growth of plants is favoured in the aboveground parts compared to the root
parts (Tilman 1991; Lambers and Poorter 1992). Consequently, plants exposed to
high levels of nutrient availability must face greater vulnerability to environmental
stressors that require large investment in roots for tolerance such as during the
430 T. M. Z. Tengku Hashim et al.

drought (Chapin 1991). Previous study was conducted by Lovelock et al. (2009) on
12 study sites where mangrove trees have been fertilized by using urea (nitrogen, N)
or triple superphosphate (Phosphorus, P) for 3 years. The study found that N
fertilization elevated the tree mortality in the region of high salinity and aridity
and trees that fertilized with P tended to have higher probability of survival than
those fertilized with N. From this increasing pattern across species and biogeo-
graphic regions, it solidly determines the influence of eutrophication in climatic
interactions with the intertidal landscape. In addition, results showed that mangrove
ecosystems that are exposed to the excess nutrient availability enhance the mortality
rate, especially during drought, and that nutrient-induced mortality is higher in sites
which has low rainfall, low humidity, and high sediment salinity.
The analysis of sediment cores collected over the Chinese island Hainan revealed
that the land conversion had a significant impact on the biogeochemistry of the
estuarine bay (Herbeck et al. 2020). The study found that the increase in Corg content
up to 2.7% in the mangrove core (1) which was being collected directly adjacent to a
riverine mangrove forest, and <0.7% in the lagoon and channel cores (2) which was
taken in the estuarine lagoon approximately 500 m from the now aquaculture-
dominated shore and core (3) which was collected from the middle of the estuarine
channel approximately 5 km from the outlet. Furthermore, in the lagoon and channel
cores (2 and 3), the increase in δ15N values was found due to a simultaneous increase
in land cover change to aquaculture. The higher value of δ15N was a result of strong
relation with the uptake and assimilation of dissolved inorganic nitrogen (DIN) by
phytoplankton in the aquaculture ponds as well as in the estuary (Herbeck et al.
2020). The increment of δ15N was 7.8% at the top sediment layer of the lagoon core
(2) are comparable to those suspended matters that were measured inside aquacul-
ture ponds and suspended matters at estuarine (Herbeck et al. 2011; Herbeck and
Unger 2013; Bao et al. 2013).
Water discharge from aquaculture shrimp farm and sewage from the developing
urban areas near mangroves which contain high level of nutrients are ordinary
sources that contribute to pollution in the Northeast of Brazil (Queiroz et al. 2013;
Suárez-Abelenda et al. 2014). A study related to eutrophication from aquaculture
was done at three selected mangroves in Ceará State, Brazil, at the estuaries of
Jaguaribe, Cocó, and Pacoti River by analysing water and soil physiochemical and
biogeochemical parameters, and soil solid-phase P fractionation (Barcellos et al.
2019). The mangrove at Jaguaribe River was chosen because this river has been
intensively impacted by shrimp farming ponds (Lacerda et al. 2008; Kauffman et al.
2018). However, the mangrove at the Cocó River was the region that received
diffuse contamination from urban sewage (Molisani et al. 2007; Nóbrega et al.
2015); and the last region was mangrove at the Pacoti River that had some degree
of disturbance with ponds for salt production, but has been reclaimed and protected
by Brazilian environmental agencies since 1990 (Lacerda et al. 2007). Results from
discriminant analysis showed that the most labile P forms increased gradually and
significantly from control (0.073 mg L 1) to sewage (0.042 mg L 1) to shrimp farm
(0.167 mg L 1) impacted mangroves as observed by increasingly dissolved ortho-
phosphate (PO43 ) content in water and the exchangeable/soluble P (Exch-P)
18 Aquaculture in Mangroves 431

extracted from soils (Barcellos et al. 2019). Hence, these aquaculture activities
would increase the P forms soluble and readily available to algae and plant growth,
with direct impacts for eutrophication and pollution in mangrove ecosystems (Bai
et al. 2015).
A large number of nutrients such as N and P discharged into the coastal water can
cause many ecological problems, such as severe eutrophication with frequent occur-
rence of harmful algal blooms (Wang et al. 2011). A study on the impact of nutrient
release from fish cage aquaculture has been done in Daya Bay, located in
Guangdong Province in southern China. The main components of DIN released by
fish were NH4–N and NO3–N. Those chemical compounds that contained nitrogen
were beneficial for the growth of phytoplankton and macroalgae (Troell et al. 2003,
2009). From the findings, the annual amount of N released from fish cage culture
was 205.6 metric tons (hereafter tons) including 142.7 tons of dissolved inorganic
nitrogen (DIN) and 39.2 tons of P which included 15.1 tons of dissolved inorganic
phosphorus (DIP). Hence, the contribution of DIN and DIP from fish culture was
about 7.0% and 2.7%, respectively (Qi et al. 2019). The ratio of cage-derived N and
P was 21.1, higher than the ratio of coastal seawater which was 27.1, thus the cage
culture may enhance the growth of phytoplankton by changing the nutrient structure
(with regard to the N:P ratio) of seawater around farming regions (Qi et al. 2019).

18.4.2 Microplastic Pollution

Recent studies have identified high abundance of microplastics in mangrove and salt
marsh habitats and suggested that the vegetation of wetlands is an effective way to
retaining media of microplastics (Nor and Obbard 2014; Sutton et al. 2016;
Weinstein et al. 2016). Qinzhou Bay in Guangxi Province, southwest China is
well known as location for natural largest breeding area of Magallana rivularis.
Wide range of mollusc farming in this area can contribute to microplastics
pollutions. Styrofoam was one of the common microplastics found which was
widely used in hanging-culture farms for mollusc. Published study by Li et al.
(2018) in Quizhou Bay found that high concentrations of microplastics were
observed near mollusc farms and polystyrene (PS) was the major microplastics
polymer found in the area which was >98%. The result was expected since PS
was broadly used to set up mollusc rafts in the Qinzhou Bay. Besides PS, polypro-
pylene (PP) and polyethylene (PE) were also observed in the area since this polymer
was used in woven bag and fish net. Study concluded that microplastic pollution in
sediments of Qinzhou Bay was mainly a result of intensive mollusc aquaculture
(Li et al. 2018).

18.4.3 Destruction of Mangrove Ecosystems

The issues on impact of coastal aquaculture on the environment, biodiversity, and


society are argued and debated around the world (Primavera 2006; Hall 2011).
432 T. M. Z. Tengku Hashim et al.

Globally, mangrove conversion to shrimp farming has been fully condemned


because of its environmental and socioeconomic impacts (Primavera 1997; Deb
1998; Naylor et al. 1998; Páez-Osuna 2001; Lebel et al. 2002; Bush et al. 2010). In
addition, high demands and economic return in the international market, as well as
unplanned and unregulated shrimp farming have caused major destruction of
mangroves in many countries, including Malaysia, Philippines, Bangladesh,
China, Indonesia, Sri Lanka, Myanmar, India, Thailand, Brazil Mexico, and
Vietnam (FAO 2007). In Asia, governments of many countries have been encourag-
ing the conversion of mangrove forest to aquaculture pond in order to deal with food
security and livelihoods since the 1960s and 1970s (Hishamunda et al. 2009).
Consequently, a study by Hamilton (2013) found out that over 54% of mangrove
deforestation (28% of the former mangrove areas) were due to conversion of
aquaculture ponds. Southeast Asia has 35% of the world’s 18 million ha of man-
grove forests (Spalding et al. 1997), however, between 2000 and 2012, aquaculture
was accounted for 30% of the mangrove forest loss. Besides aquaculture, the
relevance of other land uses has also increased, such as oil palm plantations in
Malaysia, Indonesia, rice agriculture in Myanmar, and urban development in
Vietnam (Richards and Friess 2016)

18.4.4 Introduction of Non-Native Species

Introduction of new species for aquaculture has led to an enormous invasion of


species diversity and trophic food chain in mangrove ecosystem, for e.g., introduc-
tion of the Indo-Pacific lionfish to the Atlantic (Whitfield et al. 2002) and the killer
algae, Caulerpa taxifolia to the Mediterranean (Jousson et al. 1998). In addition,
tropical white shrimp (Penaeus vannamei) native to the eastern Pacific coast of
Central and South America has been strongly established around the world since the
1970s and in 2000. Presently, it has become the principle cultured shrimp species in
Asia. However, diseases from shrimp aquaculture are much concern since the release
of untreated shrimp farm discharge water can increase the nutrient load in the open
water thus may lead to a disease outbreak in shrimp farms. This condition will
provide mangrove unhealthy water bodies (Wolanski et al. 2000; Senarath and
Visvanathan 2001). Tilapias are native to Africa and were introduced from 1946
onwards to Asia for breeding to support aquaculture industry. Unfortunately, they
are now widely colonizing brackish water areas with mangroves, even though they
are a freshwater group of fishes (Bagarinao and Primavera 2005).

18.5 Conclusion

Mangrove forests are found in sheltered saline coastal environments in tropical and
subtropical latitudes and are among the most productive ecosystems in the world as
they fulfil a range of ecologically and socioeconomically important functions
including shoreline stabilization, storm and wave protection, reduction of coastal
18 Aquaculture in Mangroves 433

erosion as well as feeding and nursery habitat for numerous commercially important
invertebrate and vertebrate species such as fishes and prawns. Besides, mangroves
play a vital role in nutrient dynamic and coastal sediment. Mangroves also act as a
carbon sink for atmospheric CO2 and crucial source of organic carbon to the coastal
sea. Since 1960s and 1970s, the conversion of mangrove forests to aquaculture
ponds was encouraged by the governments of many countries and aquaculture
production is the fastest growing industry on the planet which also responsible for
the provision of high quality animal protein worldwide. As the demand for food
supply increases due to the increasing of human population, aquaculture seems to be
one of the important solutions in maintaining, reducing the gap and fulfilling the
need. Nevertheless, the increase in aquaculture on the other hand tends to negatively
impact the environment through the spread of diseases, destruction of wetlands and
mangroves, declining in biodiversity of natural fish populations by the escape of
non-native fish, and the pollution of surface and groundwater by effluent discharge.
Hence, sustainable management can confirm a sustainable progress and the advan-
tage of aquaculture.

References
Abdul Aziz A, Dargusch P, Phinn S, Ward A (2015) Using REDD+ to balance timber production
with conservation objectives in a mangrove forest in Malaysia. Ecol Econom 120: 108–116.
Ahmadun FL-R, Wong MMR, Mat Said A (2020) Consequences of the 2004 Indian Ocean
Tsunami in Malaysia. Saf Sci 121: 619–631.
Ahmed N, Glaser M (2016) Coastal aquaculture, mangrove deforestation and blue carbon
emissions: is REDD+ a solution? Mar Pol 66: 58–66.
Ahmed N, Thompson S (2019) The blue dimensions of aquaculture: A global synthesis. Sci Total
Environ 652: 851–861.
Alongi DM (2012) Carbon sequestration in mangrove forests. Carbon Manag 3: 313–322.
Bagarinao TU, Primavera JH (2005) Code of practice for sustainable use of mangrove ecosystems
for aquaculture in Southeast Asia, aquaculture department, Southeast Asian Fisheries Develop-
ment Center.
Bai XL, Zhou YK, Sun JH, Ma JH, Zhao HY, Liu XF (2015) Classes of dissolved and particulate
phosphorus compounds and their spatial distributions in the water of a eutrophic lake: A 31 P
NMR study. Biogeochemistry 126: 227–240.
Bao H, Wu Y, Unger D, Du J, Herbeck LS, Zhang J (2013) Impact of the conversion of mangroves
into aquaculture ponds on the sedimentary organic matter composition in a tidal flat estuary
(Hainan Island, China). Cont Shelf Res 57: 82–91.
Barbier EB, Hacker SD, Kennedy C, Koch EW, Stier AC, Silliman BR (2011) The value of
estuarine and coastal ecosystem services. Ecol Monogr 81: 169–193.
Barcellos D, Queiroz HM, Nóbrega GN, De Oliveira Filho RL, Santaella ST, Otero XL, Ferreira TO
(2019) Phosphorus enriched effluents increase eutrophication risks for mangrove systems in
northeastern Brazil. Mar Pollut Bull 142: 58–63.
Barua P, Rahman SH (2020) aquatic health index of coastal aquaculture activities at South-Eastern
Coast Of Bangladesh. Water Conserv Manag 4: 53–69.
Bavington D, Banoub D (2016) Marine fish farming and the blue revolution: Culturing cod
fisheries. London Journal of Canadian Studies 31: 35–44.
Beveridge M, Phillips M, Macintosh D (1997) Aquaculture and the environment: the supply of and
demand for environmental goods and services by Asian aquaculture and the implications for
sustainability. Aquac Res 28: 797–807.
434 T. M. Z. Tengku Hashim et al.

Bradley M, Baker R, Nagelkerken I, Sheaves M (2019) Context is more important than habitat type
in determining use by juvenile fish. Landsc Ecol 34: 427–442.
Bush SR, Van Zwieten PA, Visser L, Van Dijk H, Bosma R, De Boer WF, Verdegem M (2010)
Scenarios for resilient shrimp aquaculture in tropical coastal areas. Ecol Soc 15:15.
Chapin FS (1991) Integrated responses of plants to stress. BioScience 41: 29–36.
Cloern JE (2001) Our evolving conceptual model of the coastal eutrophication problem. Mar Ecol
Prog Ser 210: 223–253.
Das CS, Mandal RN (2016) Coastal people and mangroves ecosystem resources vis-à-vis manage-
ment strategies in Indian Sundarban. Ocean Coast Manag 134: 1–10.
Deb AK (1998) Fake blue revolution: environmental and socio-economic impacts of shrimp culture
in the coastal areas of Bangladesh. Ocean Coast Manag 41: 63–88.
Downing J, Mcclain M, Twilley R, Melack J, Elser J, Rabalais N, Lewis W, Turner R, Corredor J,
Soto D (1999) The impact of accelerating land-use change on the N-cycle of tropical aquatic
ecosystems: current conditions and projected changes. Biogeochemistry 46: 109–148.
El-Regal MAA, Ibrahim NK (2014) Role of mangroves as a nursery ground for juvenile reef fishes
in the southern Egyptian Red Sea. Egypt J Aquat Res 40: 71–78.
FAO (2007) The world’s mangroves 1980–2005. FAO Rome, Italy.
FAO (2016) Aquaculture Department (2010) The state of world fisheries and aquaculture. Food and
Agriculture Organization of the United Nations, Rome.
FAO (2018) The State of World Fisheries and Aquaculture 2018. Meeting the sustainable develop-
ment goals.Rome.
FAO (2020) The State of World Fisheries and Aquaculture 2020. Sustainability in Action. FAO
Rome, Italy.
Giri C, Ochieng E, Tieszen LL, Zhu Z, Singh A, Loveland TR, Masek J, Duke NC (2011) Status
and distribution of mangrove forests of the world using earth observation satellite data. Glob
Ecol Biogeogr 20: 154–159.
Goujon A (2019) Human Population Growth. In: Fath, B. (ed.) Encyclopedia of Ecology (Second
Edition). Oxford: Elsevier.
Hall SJ (2011) Blue frontiers: managing the environmental costs of aquaculture, WorldFish.
Hamilton S (2013) Assessing the role of commercial aquaculture in displacing mangrove forest.
Bull Mar Sci 89: 585–601.
Hashim T, Suratman M, Jaafar J, Hasmadi IM, Abu F (2015) Field assessment of above ground
biomass (AGB) of mangrove stand in Merbok, Malaysia. Malays Appl Biol 44: 81–86.
Hashim T, Suratman M, Singh H, Jaafar J, Bakar A. (2020) Predictive model of mangroves carbon
stocks in Kedah, Malaysia using remote sensing. IOP Conference Series: Earth and Environ-
mental Science, 2020. IOP Publishing, 012033.
Herbeck LS, Krumme U, Andersen TJ, Jennerjahn TC (2020) Decadal trends in mangrove and pond
aquaculture cover on Hainan (China) since 1966: mangrove loss, fragmentation and associated
biogeochemical changes. Estuar Coast Shelf Sci 233: 106531.
Herbeck LS, Unger D (2013) Pond aquaculture effluents traced along back-reef waters by standard
water quality parameters, δ15N in suspended matter and phytoplankton bioassays. Mar Ecol
Prog Ser 478: 71–86.
Herbeck LS, Unger D, Krumme U, Liu SM, Jennerjahn TC (2011) Typhoon-induced precipitation
impact on nutrient and suspended matter dynamics of a tropical estuary affected by human
activities in Hainan, China. Estuar Coast Shelf Sci 93: 375–388.
Hishamunda N, Ridler NB, Bueno P, Yap WG (2009) Commercial aquaculture in Southeast Asia:
Some policy lessons. Food Pol 34: 102–107.
Hussain MG, Failler P, Karim AA, Alam MK (2018) Major opportunities of blue economy
development in Bangladesh. J Indian Ocean Reg 14: 88–99.
Hutchison J, Spalding M, Ermgassen P (2014) The role of mangroves in fisheries enhancement. The
Nature Conservancy and Wetlands International: 54.
18 Aquaculture in Mangroves 435

Igulu M, Nagelkerken I, Van Der Velde G, Mgaya Y (2013) Mangrove Fish Production is Largely
Fuelled by External Food Sources: A Stable Isotope Analysis of Fishes at the Individual,
Species, and Community Levels from Across the Globe. Ecosystems 16, 1336–1352.
Irin (2008) Bangladesh: Cyclone-hit Shrimp Farmers Face Uncertain Future, Integrated Regional
Information Network, Nairobi.
Islam MS (2008) From pond to plate: towards a twin-driven commodity chain in Bangladesh
shrimp aquaculture. Food Pol 33: 209–223.
Islam SDU, Bhuiyan MAH (2016) Impact scenarios of shrimp farming in coastal region of
Bangladesh: an approach of an ecological model for sustainable management. Aquac Int 24:
1163–1190.
Jamizan AR, Chong V (2017) Demersal Fish and Shrimp Abundance in Relation to Mangrove
Hydrogeomorphological Metrics. Sains Malaysiana 46: 9–19.
Joffre OM, Bosma RH, Bregt AK, Van Zwieten PA, Bush SR, Verreth JA (2015) What drives the
adoption of integrated shrimp mangrove aquaculture in Vietnam? Ocean Coast Manag 114:
53–63.
Jousson O, Pawlowski J, Zaninetti L, Meinesz A, Boudouresque CF (1998) Molecular evidence for
the aquarium origin of the green alga Caulerpa taxifolia introduced to the Mediterranean Sea.
Mar Ecol Prog Ser 172: 275–280.
Kamruzzaman M, Ahmed S, Osawa A (2017) Biomass and net primary productivity of mangrove
communities along the Oligohaline zone of Sundarbans, Bangladesh. For Ecosys 4: 1–9.
Kathiresan K, Bingham BL (2001) Biology of mangroves and mangrove ecosystems. Adv Mar Biol
40, 81–251.
Kauffman JB, Bernardino AF, Ferreira TO, Bolton NW, Gomes LEDO, Nobrega GN (2018)
Shrimp ponds lead to massive loss of soil carbon and greenhouse gas emissions in northeastern
Brazilian mangroves. Ecol Evol 8: 5530–5540.
Kawarazuka N, Béné C (2011) The potential role of small fish species in improving micronutrient
deficiencies in developing countries: building evidence. Public Health Nutr 14: 1927–1938.
Keur R, James N, Rajkaran A (2019) A tale of two habitats: preliminary comparison of fish
abundance and diversity between saltmarsh- and mangrove-dominated creeks in the Nahoon
Estuary, South Africa. Afr J Mar Sci 41: 449–454.
Krauss KW, Ball MC (2013) On the halophytic nature of mangroves. Trees 27: 7–11.
Lacerda LD, Molisani MM, Sena D, Maia LP (2008) Estimating the importance of natural and
anthropogenic sources on N and P emission to estuaries along the Ceará State Coast NE Brazil.
Environ Monit Assess 141: 149–164.
Lacerda LDD, Menezes MOTD, Molisani MM (2007) Changes in mangrove extension at the Pacoti
River estuary, CE, NE Brazil due to regional environmental changes between 1958 and 2004.
Biota Neotropica 7: 67–72.
Laegdsgaard P, Johnson C (2001) Why do juvenile fish utilise mangrove habitats? J Exp Mar Biol
Ecol 257: 229–253.
Lambers H, Poorter H (1992) Inherent variation in growth rate between higher plants: a search for
physiological causes and ecological consequences. Advances in ecological research. Elsevier.
Lebel L, Tri NH, Saengnoree A, Pasong S, Buatama U (2002) Industrial transformation and shrimp
aquaculture in Thailand and Vietnam: pathways to ecological, social, and economic
sustainability? AMBIO: A Journal of the Human Environment 31: 311–323.
Li HX, Ma LS, Lin L, Ni ZX, Xu XR, Shi HH, Yan Y, Zheng GM, Rittschof D (2018) Microplastics
in oysters Saccostrea cucullata along the Pearl River estuary, China. Environ Pollut 236:
619–625.
Lovelock CE, Ball MC, Martin KC, Feller IC (2009) Nutrient enrichment increases mortality of
mangroves. PLoS One 4: e5600.
Mason NW, Mouillot D, Lee WG, Wilson JB (2005) Functional richness, functional evenness and
functional divergence: the primary components of functional diversity. Oikos 111: 112–118.
Mialhe F, Gunnell Y, Mering C (2013) The impacts of shrimp farming on land use, employment
and migration in Tumbes, northern Peru. Ocean Coast Manag 73: 1–12.
436 T. M. Z. Tengku Hashim et al.

Molisani M, Lisieux R, Cavalcante M, Maia L (2007) Effects of water management on hydrology


and water quality of a semi-arid watershed in the Northeast of Brazil. Braz J Aquat Sci Tech 11:
43–49.
Moni NN, Al Masud MM, Hossen SS (2018) Trade liberalization and socio-economic vulnerabil-
ity: a cross country analysis on the shrimp sector. Khulna University Studies Volume 15 (1):
71─90
Nagelkerken I, Blaber SJM, Bouillon S, Green P, Haywood M, Kirton LG, Meynecke JO, Pawlik J,
Penrose HM, Sasekumar A, Somerfield PJ (2008) The habitat function of mangroves for
terrestrial and marine fauna: A review. Aquat Bot 89: 155–185.
Naidoo G (2016) The mangroves of South Africa: An ecophysiological review. South Afr J Bot
107: 101–113.
Nanjo K, Kohno H, Nakamura Y, Horinouchi M, Sano M (2014a) Differences in fish assemblage
structure between vegetated and unvegetated microhabitats in relation to food abundance
patterns in a mangrove creek. Fish Sci 80: 21–41.
Nanjo K, Kohno H, Nakamura Y, Horinouchi M, Sano M (2014b) Effects of mangrove structure on
fish distribution patterns and predation risks. J Exp Mar Biol Ecol 461: 216–225.
Naylor RL, Goldburg RJ, Mooney H, Beveridge M, Clay J, Folke C, Kautsky N, Lubchenco J,
Primavera J, Williams M (1998) Nature’s subsidies to shrimp and salmon farming. American
Association for the Advancement of Science.
Nguyen XH, Nguyen XT, Nguyen HH, Tran TT, Le DQ (2018) Assessment of storm surge risk in
aquaculture in the Northern coastal area of Vietnam. Vietnam J Sci Technol Engineer 60: 89–94.
Nóbrega GN, Ferreira TO, Artur AG, De Mendonça ES, Raimundo ADO, Teixeira AS, Otero XL
(2015) Evaluation of methods for quantifying organic carbon in mangrove soils from semi-arid
region. J Soils Sediments 15: 282–291.
Nor NHM, Obbard JP (2014) Microplastics in Singapore’s coastal mangrove ecosystems. Mar
Pollut Bull 79 (1–2): 278–283.
Ottinger M, Clauss K, Kuenzer C (2016) Aquaculture: Relevance, distribution, impacts and spatial
assessments–A review. Ocean Coast Manag 119: 244–266.
Páez-Osuna F (2001) The environmental impact of shrimp aquaculture: causes, effects, and
mitigating alternatives. Environ Manag 28: 131–140.
Pantallano ADS, Bobiles RU, Nakamura Y (2018) Dependence of fish on subtropical riverine
mangroves as habitat in the Ryukyu Islands, Japan. Fishe Sci 84: 613–625.
Pradeepkiran JA (2019) Aquaculture role in global food security with nutritional value: a review.
Transl Anim Sci 3: 903–910.
Primavera JH (1997) Socio-economic impacts of shrimp culture. Aquac Res 28: 815–827.
Primavera JH (2006) Overcoming the impacts of aquaculture on the coastal zone. Ocean Coast
Manag 49: 531–545.
Qi Z, Shi R, Yu Z, Han T, Li C, Xu S, Xu S, Liang Q, Yu W, Lin H (2019) Nutrient release from fish
cage aquaculture and mitigation strategies in Daya Bay, southern China. Mar Poll Bull 146:
399–407.
Queiroz L, Rossi S, Meireles J, Coelho C (2013) Shrimp aquaculture in the federal state of Ceará,
1970–2012: Trends after mangrove forest privatization in Brazil. Ocean Coast Manag 73:
54–62.
Richards DR, Friess DA (2016) Rates and drivers of mangrove deforestation in Southeast Asia,
2000–2012. Proc Natl Acad Sci 113: 344–349.
Romañach SS, Deangelis DL, Koh HL, Li Y, Teh SY, Raja Barizan RS, Zhai L (2018) Conserva-
tion and restoration of mangroves: Global status, perspectives, and prognosis. Ocean Coast
Manag 154: 72–82.
Rönnbäck P, Troell M, Kautsky N, Primavera JH (1999) Distribution pattern of shrimps and fish
among Avicennia and Rhizophora micro habitats in the Pagbilao Mangroves, Philippines.
Estuar Coast Shelf Sci 48: 223–234.
Sánchez-Núñez DA, Bernal G, Pineda JEM (2019) The Relative Role of Mangroves on Wave
Erosion Mitigation and Sediment Properties. Estuar Coast 42: 2124–2138.
18 Aquaculture in Mangroves 437

Sarker T, Azam M (2007) Super Cyclone SIDR 2007: climate change adaptation mechanisms for
coastal communities in Bangladesh. Sarker, T. and Azam, M.(2012). Super cyclone SIDR:
85–105.
Senarath U, Visvanathan C (2001) Environmental issues in brackish water shrimp aquaculture in Sri
Lanka. Environ Manag 27: 335–348.
Sheaves M, Johnston R, Baker R (2016) Use of mangroves by fish: new insights from in-forest
videos. Mar Ecol Prog Ser 549: 167–182.
Spalding MD, Blasco F, Field CD eds (1997) World mangrove atlas. The international Society for
Mangrove Ecosystems, Okinawa, Japan. 178 pp.
Suárez-Abelenda M, Ferreira T, Camps-Arbestain M, Rivera-Monroy V, Macías F, Nóbrega GN,
Otero X (2014) The effect of nutrient-rich effluents from shrimp farming on mangrove soil
carbon storage and geochemistry under semi-arid climate conditions in northern Brazil.
Geoderma 213: 551–559.
Suratman MN (2008) Carbon sequestration potential of mangroves in southeast Asia. Managing
forest ecosystems: The challenge of climate change. Springer.
Sutton PC, Anderson SJ, Costanza R, Kubiszewskic I (2016) The ecological economics of land
degradation: Impacts on ecosystem service values. Ecological Economics Volume 129:
182–192.
Tanaka K, Hanamura Y, Chong VC, Watanabe S, Man A, Kassim FM, Kodama M, Ichikawa T
(2011) Stable isotope analysis reveals ontogenetic migration and the importance of a large
mangrove estuary as a feeding ground for juvenile John’s snapper Lutjanus johnii. Fish Sci 77:
809.
Thomas N (2017) The importance of mangroves to people: a call to action.
Tilman D (1991) Relative growth rates and plant allocation patterns. Am Nat 138: 1269–1275.
Troell M, Halling C, Neori A, Chopin T, Buschmann A, Kautsky N, Yarish C (2003) Integrated
mariculture: asking the right questions. Aquaculture 226: 69–90.
Troell M, Joyce A, Chopin T, Neori A, Buschmann AH, Fang JG (2009) Ecological engineering in
aquaculture—potential for integrated multi-trophic aquaculture (IMTA) in marine offshore
systems. Aquaculture 297: 1–9.
Truong TD, Do LH (2018) Mangrove forests and aquaculture in the Mekong river delta. Land Use
Policy 73: 20–28.
Uddin MB, Sultana T, Rahman MM, Uddin MB, Sultana T, Rahman MM (2013) Shrimp export
marketing of Bangladesh. Bangladesh Res Pub J 8: 146–151.
Valiela I, Bowen JL, York JK (2001) Mangrove forests: one of the world’s threatened major
tropical environments. Bioscience 51: 807–815.
Van Wesenbeeck BK, Balke T, Van Eijk P, Tonneijck F, Siry HY, Rudianto ME, Winterwerp JC
(2015) Aquaculture induced erosion of tropical coastlines throws coastal communities back into
poverty. Ocean Coast Manag 116: 466–469.
Venkatachalam S, Kathiresan K, Krishnamoorthy I, Narayanasamy R (2018) Survival and growth
of fish (Lates calcarifer) under integrated mangrove-aquaculture and open-aquaculture systems.
Aquac Rep 9: 18–24.
Verweij M, Nagelkerken I, De Graaff D, Peeters M, Bakker E, Van Der Velde G (2006) Structure,
food and shade attract juvenile coral reef fish to mangrove and seagrass habitats: a field
experiment. Mar Ecol Prog Ser 306: 257–268.
Vo QT, Oppelt N, Leinenkugel P, Kuenzer C (2013) Remote sensing in mapping mangrove
ecosystems—An object-based approach. Remote Sens 5: 183–201.
Wang X, Hao F, Cheng H, Yang S, Zhang X, Bu Q (2011) Estimating non-point source pollutant
loads for the large-scale basin of the Yangtze River in China. Environ Earth Sci 63: 1079–1092.
Weinstein JE, Crocker BK, Gray AD (2016) From macroplastic to microplastic: Degradation of
high‐density polyethylene, polypropylene, and polystyrene in a salt marsh habitat. Environ
Toxicol Chem 35: 1632–1640.
438 T. M. Z. Tengku Hashim et al.

Whitfield PE, Gardner T, Vives SP, Gilligan MR, Courtenay Jr WR, Ray GC, Hare JA (2002)
Biological invasion of the Indo-Pacific lionfish Pterois volitans along the Atlantic coast of North
America. Mar Ecol Prog Ser 235: 289–297.
Win S, Towprayoon S, Chidthaisong A (2019) Adaptation of mangrove trees to different salinity
areas in the Ayeyarwaddy Delta Coastal Zone, Myanmar. Estuar Coast Shelf Sci 228: 106389.
Wolanski E, Spagnol S, Thomas S, Moore K, Alongi D, Trott L, Davidson A (2000) Modelling and
visualizing the fate of shrimp pond effluent in a mangrove-fringed tidal creek. Estuar Coast
Shelf Sci 50: 85–97.
Yusoff A (2014) Published. Status of resource management and aquaculture in Malaysia. Resource
Enhancement and Sustainable Aquaculture Practices in Southeast Asia: Challenges in Respon-
sible Production of Aquatic Species: Proceedings of the International Workshop on Resource
Enhancement and Sustainable Aquaculture Practices in Southeast Asia 2014 (RESA), 2015.
Aquaculture Department, Southeast Asian Fisheries Development Center, 53–65.
Ecological Valuation and Ecosystem
Services of Mangroves 19
Hong Tinh Pham, Thi Hong Hanh Nguyen, and Sy Tuan Mai

Abstract

Knowledge of the economic value of mangrove ecosystem plays an important


role in assisting managers and policy-makers in determining priorities, policies,
and actions for conservation of the mangrove ecosystem. Methods to evaluate
natural and environmental resources have been widely applied to raise awareness
as to the value of mangrove ecosystems. In this chapter, we describe the economic
and non-economic contribution of mangrove standings to both communities and
society. The services which are related to the various components and ecological
functions of a mangrove ecosystem are summarized. The chapter also describes
how mangrove ecosystem services are valuated and what their estimated values
are towards the provision of these services.

Keywords
Mangrove ecosystem · Ecological functions of mangroves · Ecosystem service ·
Ecological value · Valuation Method

H. T. Pham (*) · T. H. H. Nguyen


Faculty of Environment, Hanoi University of Natural Resources and Environment, Hanoi, Vietnam
e-mail: [email protected]
S. T. Mai
Mangrove Ecosystem Research Center, Hanoi National University of Education, Hanoi, Vietnam

# The Author(s), under exclusive license to Springer Nature Singapore Pte 439
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_19
440 H. T. Pham et al.

19.1 Introduction

Globally mangroves are distributed in the inter-tidal zone of many tropical and
subtropical coastal wetlands, forming some of the most productive and biologically
diverse ecosystems. Mangroves cover an estimated area of 137,600 km2 (Fig. 19.1)
with Asia accounting for 38.7% (53,278 km2), followed by Latin America and the
Caribbean with 27,939 km2 (20.3%), Africa with 27,465 km2 (20.0%), Oceania with
16,329 km2 (11.9%), North America with 11,563 km2 (8.4%), and Europe with 1026
km2 (0.7%) (Bunting et al. 2018). Mangroves exist at the terrestrial and marine
interface and are considered to be one of the most productive and biologically
diverse of all ecosystems. Mangrove leaf litter makes up the majority of net primary
production, acting as a key source of organic matter within directly-associated food
chains. Detritus-based food is also exported by water currents, providing nutrients
for marine/aquatic organisms within a large coastal area (Naylor et al. 2000;
Sheridan and Hays 2003; Aburto-Oropeza et al. 2008). Mangroves reduce the effects
of larger waves and elevated water levels as a result of storm surges and tsunamis
(Mazda et al. 2006; Zhang et al. 2012) providing coastal protection and helping local
communities become more resilient. Mangroves also serve as a protected nursery for
many marine species (Rönnbäck 1999; Walters et al. 2008).
The ecological importance of mangroves is widely recognized and the services
that they provide are recognized by academics. However, there is often a lack of
understanding or underestimation of the value of mangrove ecosystems and the
benefits they provide by policy makers. There is also an increasing pressure to
exploit resources by the local communities with mangroves sacrificed to make
way for shrimp ponds, aquaculture, and other short-term land use purposes, and
the impacts of anthropogenic pollution. This has led to a serious decline in the total
area and quality of mangrove forests and resulted in the development and refinement
of tools and models to assess and highlight the economic value of this precious
ecosystem.
It is acknowledged that ecosystems in general and mangrove ecosystems, in
particular, provide people with non-economic benefits (Tri 2006) and placing a
monetary value on a natural ecosystem do not always recognize the tangible and
non-use benefits of the system. However, economic valuation is a useful tool for

Fig. 19.1 Mangrove extent per country (Adapted from Bunting et al. 2018)
19 Ecological Valuation and Ecosystem Services of Mangroves 441

policy makers to make rational decisions against the pressures of economic devel-
opment and contributing to the sustainable management of coastal resources.
Providing an economic valuation of the mangrove ecosystem requires interdisci-
plinary expertise and knowledge of both mangrove ecology and a theoretical under-
standing of valuation science. Experience and practical knowledge of both fields is
also essential for application in specific cases. This chapter presents the current
knowledge and a theoretical basis for developing methods for valuing mangrove
ecosystems and application of a systematised approach to utilising research findings.

19.2 Components and Functions of the Mangrove Ecosystems

Mangroves are mainly woody trees and shrubs, characterized by the aerial roots and
seedlings that germinate on the plant. They grow at the land–sea interface in the
tropical and subtropical regions of the world (Latitude 30oN and 30o). Mangroves,
along with associated organisms (e.g. microorganisms, fungi, and animals), interact
with abiotic factors, regulating a distinct coastal tide-inundated ecosystem
(Kathiresan and Bingham 2001) (Figs. 19.2 and 19.3, Table 19.1) that undergoes
spatial-temporal movement (Tri et al. 2002).
The mangrove ecosystem is characterized by the coastal climate, alluvial soil, and
tidal mud flats that are inundated by salt water from diurnal or semi-diurnal tides and
mixing with fresh water from rivers resulting in brackish water and the physico-
chemical properties of the water (e.g. salinity, pH, and temperature) is constantly
changing with space and time. The biotic components of the mangrove ecosystem
include marine and terrestrial organisms, and organisms adapted to the mangrove
forest (Table 19.1).
Complex, functional, and dynamic interactions take place within the system
which are capable of self-regulating and sustaining the mangrove ecosystem in a
relatively steady state. However, negative anthropogenic activities can affect

Fig. 19.2 Mangrove ecosystem at Ba Lat estuary, Vietnam (Image source: Pham Hong Tinh)
442 H. T. Pham et al.

Fig. 19.3 A woman catching


crabs under mangrove canopy
(Image source: Pham Hong
Tinh)

ecosystem stability. The basic functions of the ecosystem are determined by the
following interactive relationships:
Interaction between abiotic factors: This is clearly shown in such relationships
as: salinity decreasing during periods of extended rainfall; turbidity increases when
the amount of sediment input from the river is high; the substrate may become highly
anaerobic during extended periods of flooding caused by sea water inundation.
Interaction between organisms: This is shown in food chains and webs, trophic
levels, and moving energy flow in the ecosystem. These are the relationships
between producers and consumers (mangrove plants - insects; mangrove leaves
fallen on the forest floor - crabs, algae - zooplankton), predators and preys (seabass
eat shrimp, shrimp eat plankton), symbiotic organisms, saprophytes (pathogenic
fungi), parasites (worms, pathogenic bacteria), commensalism.
Interaction between biotic and abiotic factors: For example, abiotic factors (soil,
water, air environment) affect the distribution, zoning, growth and development of
mangrove species and other organisms in the forest. While mangroves can influence
the microclimate, sequester carbon and the creation of burrows by crabs improves
soil conditions. The relationship between biotic and abiotic factors is also reflected in
the biogeochemical cycle, such as cycles of water, carbon, nitrogen, phosphorus, and
sulfur. For example, the nutrient cycle in the mangrove ecosystem shows that
mangroves and algae synthesize organic matter via photosynthesis using sunlight,
CO2 in the air, water and minerals obtained from the soil. These organic matters are
stored in plant body tissues that are leaves, stems, and roots. Plant organic matters
are directly eaten by other organisms (worms, bugs, and crabs) or after being
decomposed, (decomposed food) help build up body tissues and energy for the
movement of animals. When these animals die, they are broken down by
microorganisms and fungi into minerals, returned to the soil, and then sucked up
by plants as nutrients. This cycle is called the nutrient cycle or biogeochemical cycle
because of the involvement of biological processes (food chains and food webs),
19 Ecological Valuation and Ecosystem Services of Mangroves 443

Table 19.1 Basic characteristics of components in a mangrove ecosystem


Component in the ecosystem Sea factor Land factor Mangrove factor
Abiotic Atmosphere Sea climate, The inland climate A place where there
components wind, storms, depends on the is a mixture of
high humidity, distribution of marine and inland
... landscape and climate.
vegetation factors
Soil Anaerobic soils Mud flats are Mud flats in
dependent on gradually elevated estuarine and
frequency and coastal areas:
level of tidal anaerobic, salty,
inundation acid sulfate.
Water Saline seawater Fresh water from the Brackish water
river (a mixture of
seawater and fresh
water)
Biotic Micro- Marine micro- Inland micro- Micro-organisms,
components organisms, organisms, organisms, fungi, fungi, algae in
fungi, moss, fungi, moss, moss, in mangrove brackish water, on
algae algae in sea plants mangrove roots
water
Mangrove Associate mangrove True mangrove
trees species species
Invertebrates Shrimps, Insects and bugs Some species of
oysters, clams, (ants, butterflies, insects, crabs, ... can
crabs, worms, ... bees) living in live both in
living in the mangrove plants mangrove plants
marine and in seawater
environment.
Vertebrates Fish, reptiles, Reptiles, birds, Typical species of
birds, animals animals living in mangrove forests
living in the mangrove trees and and seasonal
marine those from the inland migratory species
environment area coming here for (fish, sea turtles,
food, nesting snakes, birds,
dugongs)

chemical reactions in metabolic, decomposition, mineralization, processes, physical,


geographic, geomorphologic processes of the environment.

19.3 Goods and Services of Mangrove Ecosystems

In society, the relationships between people are economically reflected in the form of
goods and services supplied by one person to the other and vice versa. In an
ecosystem, goods and services are products of nature. These products have been
developed during a long-term evolution process. While many people mistakenly
believe that these products are readily available in nature for supply and continued
exploitation by humans (Lebel et al. 2002).
444 H. T. Pham et al.

Mangrove ecosystems provide goods to meet local people’s demand for building
materials, energy for cooking and food. These types of goods are traded and
exchanged at the market, including plant products such as wood for house construc-
tion, furniture and firewood, nipa leaves for roofing. In addition, food resources,
including honey, shrimp, crabs, fish, and oysters are an important source of nutrition
for local people, and a valuable source of export income (Adger et al. 1995).
However, the value of the mangrove ecosystem is not only shown in goods, but
also in its ability to provide essential and important services (Barbier et al. 1991)
closely related to ecosystem function. The services ecosystems provide are divided
into groups, including: provisioning services (food, raw materials), regulating
services, supporting services, and cultural services (de Groot et al. 2012; MEA
2005). The services provided by mangroves are summarized in Table 19.2.

19.4 Valuation Methods of Mangrove Ecosystem Services

The valuation of an ecosystem in general and a mangrove ecosystem in particular is


difficult, with authorities involved in managing and formulating policies from the
central to the local level facing difficulties in making decisions on forest conserva-
tion or economic development. Mangrove ecosystem valuation is very useful if it is
used as a tool to contribute to identifying priorities, policies, and actions for
conservation. Economists have developed theoretical methods for evaluating natural
resources and the environment. Those methods are divided into four groups
(Table 19.3).
Detailed descriptions of some of the most widely adopted methods are given
below:
Market price method: uses the market price to estimate the value of an ecosystem
good or service, such as firewood or aquatic products sold at the market. Other
ecosystem services such as absorbing wastes that help clean water sources, increas-
ing fisheries productivity, etc. are determined through the benefits of end products.
Intangible services as the aesthetics of the landscape, inspiration for the creation of
art and maintenance of moral fortitude cannot be sold at market, but the prices which
people want to pay or the degree of willingness to pay for these services can be
valued. Pricing, quantity, preference, and cost data can all be easily collected and
used to provide an economic evaluation. However, this method can only be applied
to a limited number of goods and services brought about by ecosystems. It does not
reflect the value of all ecosystem functions, nor it is fully reflective of the economic
value of goods and services, as market prices are dependent on the policies, seasons,
and other factors.
Benefit transfer method: estimates economic values for ecosystem services by
applying assessment findings of previous research from other locations provided that
these locations have the same conditions to reduce finance, time, and human
resources. When this method is applied, it is necessary to carefully assess the
transferability of findings; especially the features of residents, tradition, educational
level (knowledge and responsibilities) as well as their interests. This requires the
19 Ecological Valuation and Ecosystem Services of Mangroves 445

Table 19.2 Ecosystem services provided by mangroves


Ecosystem Provisioning Regulating Supporting Cultural
components services services services services
Mangrove trees – Provide: – Absorb carbon – Nutrient cycle – Scientific
(Avicennia, • Building and pollutants, – Seed dispersal research
Rhizophora, materials ameliorate the – Trap alluvial – Maintain
Sonneratia, • Fuel effects of sediment, expand cultural
Ceriops, . . .) (firewood, climate change the area of mud heritage
charcoal) – Filter water flats – Historical
• Herbal – Reduce the – Reduce Erosion value,
medicine impacts of – Provide a Ecotourism
– Mangrove winds, storms, spawning ground and
resin is used in and cyclones for aquatic landscape
the processing – Regulate species, and tourism
technology of temperature, nesting sites for – Create
varnish, ink humidity birds; jobs for
– Limit inland – Provide a local people
saline intrusion refuge for
– Maintain vertebrates
Ecological
balance
Aquatic and – Provide food – Filter water, – Food sources – Scientific
marine resources for people reduce water for other species research
( fish, shrimp, – Animal feed pollution – Nutrient cycle – Create
crabs, oysters, to serve – Maintain – Reserve source jobs for
snails, . . .) husbandry Ecological – Organic people
– Herbal balance decomposition
medicine – Maintain Soil
– Genetic structure
resources for
aquaculture
Other species – Food for – Maintain – Food sources – Scientific
(Pythons, snakes, people Ecological for other species research
turtles, otters, – Herbal balance – Gene reserves – Maintain
birds, monkeys, medicine – Maintain Soil – Help flowers cultural
honeybees, ...) – Genetic structure pollinate heritage
resource – Seed dispersal – Create
jobs for
people
Soil and water in – Shelter – Regulate – Soil – Historical
the coastal – Farmland climate improvement value
mangrove area – Supply fresh – Conserve – Prevention – Maintain
water from rain, water and land against forest fire cultural
groundwater resources – Shelter for wild heritage
– Ecological animals – Create
balance jobs for
people

consultation/comments of experts, scientists, officials, and local people. Researchers


have to understand fully the study that is being utilized, including raw data, actual
survey data, number of people interviewed, etc. However, this method should only
446 H. T. Pham et al.

Table 19.3 Methods of economic valuation of mangrove ecosystem services


Category Method Application
Real market Market price Economic values of a mangrove product or service are
methods estimated through its prices in commercial markets.
Productivity Economic values of a mangrove product or service are
change estimated through its contribution to the production of
the mangrove ecosystem products or services that are
bought/sold in commercial markets.
Replacement Economic values of a mangrove product or service are
cost estimated by using the costs of replacing ecosystem
products or services.
Avoided cost Economic values of a mangrove product or service are
estimated by using the costs of avoided damages
resulting from lost ecosystem services.
Surrogate Travel cost Economic values of mangrove products or services are
market methods estimated by asking people about their willingness to pay
to travel to visit the mangrove ecosystem.
Hedonic price Economic values of a mangrove product are decomposed
its total value into the value of its several attributes.
Production Economic values of a mangrove product or service are
function estimated by modelling the physical contribution of the
that product or service to economic output.
Hypothetical Contingent Economic values of a mangrove product or service are
market methods valuation estimated by asking people about their willingness to pay
for it in a hypothetical scenario.
Choice Economic values of a mangrove product or service are
modeling estimated by asking people to make tradeoffs among
them; willingness to pay is inferred from tradeoffs that
include cost.
Other methods Benefit transfer Economic values of a mangrove product or service are
estimated by using existing estimates from similar studies
in other sites
Ecosystem Economic values of a mangrove service are estimated by
value multiplying the mangrove area by global averaged value.
coefficient

be applied in special cases because collecting and evaluating primary data for a
particular area is more reliable. Costanza et al. (1997) and de Groot et al. (2012) used
the benefit transfer method to estimate the global value of ecosystem services. These
authors used values from many other studies to calculate average values which were
then multiplied by biome area estimated from a global land use map.
Productivity change method: estimates economic values of ecosystem products
or services that contribute to the production of commercially marketed goods (input).
For example, the economic benefits of improving water quality can be measured by
increases in crop yields, increased aquatic species due to the availability of organic
humus as a suitable source of food and the habitat mangrove forests provide. This
method, while not complicated, requires coordination between ecologists and
economists; the data collected should reflect changes in the quality and quantity of
19 Ecological Valuation and Ecosystem Services of Mangroves 447

raw materials from natural resources, including costs of production for the final
goods; supply and demand for end products; supply and demand for other factors of
the production process. This information is used in relation to the impact of
quantitative and qualitative changes leading to amendments to producer or consumer
surpluses, based on which economic benefits can be evaluated.
Travel cost method: is used to value ecosystem services related to tourism and
entertainment, in which the cost–benefit analysis is based on: entrance fee and the
area and quality of the environment in the tourist area. This approach is based on the
time and costs tourists spend on travel to a given location. Therefore, the degree of
willingness of tourists to pay for their visit to a given location can be determined by
the number of times they visit and associated travel costs. This method is quite
simple, inexpensive, and easily applied, however, one must assume that tourists visit
a site for only one purpose. For travellers who visit a location with multiple
purposes, valuing services for each purpose is more complex. On the other hand,
travel time can be used in many ways and the time spent visiting a tourist site can be
viewed as an “opportunity cost” (Tri 2006).
Methods of avoided cost, replacement cost: are economic valuation methods
based on the cost of avoided damages due to preventing the loss of ecosystem
services, the cost of ecosystem services replacement or the cost of providing
additional ecosystem services. These methods do not rigorously measure the eco-
nomic value, but are based on willingness to pay for specific goods or services.
These methods can be applied in cases such as: valuing erosion prevention services
of a forest or wetland by assessing the cost of dredging downstream river beds;
valuing the water cleaning service that wetlands provide by calculating the cost of
filtering and chemically treating water; valuing sea dyke protection services of
mangrove ecosystems by calculating the costs of constructing permanent dyke
sections; valuing the service of providing nurturing and breeding grounds for coastal
marine/aquatic species through cost assessment of spawning ground rehabilitation
programs and companies providing nurseries for and supplying aquatic seed source.
Contingent valuation method: is also known as the “stated preference” method or
“revealed preference” because it is directly related to interviewing people; to establish,
for example, willingness to pay a certain amount of money for specific environmental
services (Tri et al. 1998). These values are not related to the purchasing and selling at
the market and are therefore also known as passive values. They are composed of
everything from basic life support associated with ecosystem health or biodiversity, to
the enjoyment of natural beauty and responsibilities for leaving a legacy to future
generations. . . For example, voluntarily contributing to mangrove conservation today
will help improve knowledge for future generations such as the salt secreting mecha-
nism of Avicennia mangroves and the existence of the otters in the coastal area (Tri
2006). When awareness is raised, people are willing to pay for environmental benefits
regardless of the passive use value or non-use value. How much money each person is
willing to pay depends on behavioral, psychological, cognitive, and cultural factors
which limits the reliability of this method of assessment.
Hedonic price method: is used to estimate the economic value of environmental
or ecosystem services that are directly related to and affect market prices. The basis
of this approach is the determination of a price for a marketed good associated with
448 H. T. Pham et al.

Table 19.4 Number of studies of mangrove ecosystem service valuation during 2007–2016
(Adapted from Himes-Cornell et al. 2018)
Number of studies of evaluation of mangrove ecosystem services
Evaluation Provisioning Regulating Supporting Cultural
methods services services services services Total
Avoided cost 3 1 4
Damage cost 2 2
Market price 35 10 2 11 58
Net price 1 1
method
Replacement 10 10
cost
Production 8 3 4 1 16
function
Travel cost 2 2
Choice 2 1 3
modeling
Contingent 1 1 2 4
valuation
Benefit transfer 29 49 15 18 111
Others 11 20 6 9 46
Total 87 98 28 44 257

the characteristics and services it provides. Therefore, it is possible to evaluate each


characteristic or determine the price a customer is willing to pay when each charac-
teristic is improved in accordance with their preferences.
The application of this method is flexible and relevant to one or a group of
mangrove ecosystem services. As discussed in a previous review study (Himes-
Cornell et al. 2018) on economic evaluation methods for mangrove ecosystem
services, market price is predominantly used to value directly paid mangrove
ecosystem services (e.g., food, raw materials, carbon accumulation, leisure travel/
tourism), while the benefit transfer method is widely used to estimate the value of all
other ecosystem services. The other remaining valuation methods are not very
frequently used.
The number of studies on the evaluation of mangrove ecosystem services by each
research method is shown in Table 19.4.

19.5 Economic Value of Mangrove Ecosystem Services

Over the past two decades, many studies assessing the economic value of mangrove
ecosystems have been carried out. These studies have mainly focused on a specific
ecosystem or geographic area (Camacho-Valdez et al. 2014; Amarnath and Mouna
2016; Atkinson et al. 2016; Jerath et al. 2016; Mashayekhi et al. 2016; Mojiol et al.
2016; Sopheak and Hoeurn 2016; Susilo et al. 2016). However, some studies have
19 Ecological Valuation and Ecosystem Services of Mangroves 449

Raw material (21) 746.32

2131.25
Water (2)

12436.24
Food (25)

0 5000 10000 15000 20000

Values (USD/ha/year)

Fig. 19.4 Values of providing services of mangrove ecosystem (Based on data compiled by
Himes-Cornell et al. 2018)

Fig. 19.5 Seadykes protected by mangroves (a) and seadyke eroded due to absence of protection
mangroves (b) in Northern Vietnam (Image source: Phan Hong Anh)

attempted to calculate the global value of mangrove ecosystem services (Costanza


et al. 1997; de Groot et al. 2012). The value of providing food, raw materials,
moderating extreme events, preventing erosion, and maintaining life cycles of
migratory species has been widely documented. However, other ecosystem services
such as pollination, ornamental resources, and cultural benefits have received less
attention (Himes-Cornell et al. 2018).
Mangroves provide a wide range of products (as detailed previously) are directly
related to net primary production (Saenger 2002). This provisioning service has been
evaluated by many scientists in different geographic regions around the world.
Figure 19.4 shows estimates of the economic value of ecosystem services
provided by mangroves, ranging from 746.32 USD/ha /year (raw material provi-
sioning service) or 2131.25 USD/ha/year (water related services) and up to
12,436.24 USD/ha/year (aquatic product provisioning service). Estimates shown in
Figs. 19.4 and 19.5 are compiled and calculated from 48 different studies,
concerning the value of services providing food, raw materials, and water.
450 H. T. Pham et al.

Maintenance of soil fertility


640.00
and nutrient cycling (1)

Regulation of water flows (2) 600.00

Moderation of extreme events (14) 1205.75

Erosion prevention (10) 1016.23

Climate regulation (11) 38226.69

Waste treatment (7) 2826.86

0 20000 40000 60000 80000

Values (USD/ha/year)

Fig. 19.6 Values of regulating services provided by mangrove ecosystems (Based on data
compiled by Himes-Cornell et al. 2018)

Regulating services of a mangrove ecosystem, namely, quality maintenance,


storm, flood and erosion control and climate regulation have been also widely
studied and evaluated (Saenger 2002; Walters et al. 2008). In contrast limited
attention (in terms of quality and quantity analysis) has been given to other
regulating services including air quality, water flow (salt water intrusion), and
biological control (Ilman et al. 2011).
Figure 19.6 lists the values of regulating services and highlights significant
variation, not only between services but also within the same service. For example,
the value of climate regulation services ranges from 2.2 USD/ha/year to 414,411
USD/ha/year, and the value of water treatment service varies from 31 USD/ha/year
to 11,000 USD/ ha/year. Figure 19.6 also shows that climate regulation has the
largest economic value of 38,226.69 USD/ha/year, the remaining services are much
less such as water treatment 2826.86 USD/ha/year, erosion control 1016.23 USD/ha/
year or flow regulation 600 USD/ha/year.
Mangrove ecosystems are considered to play a very important role in supporting
coastal and offshore fishing as well as aquaculture by providing a breeding and
nursery ground and habitat to various fish and crustaceans (Rönnbäck 1999; Walters
et al. 2008). Therefore, supporting services are valued through the amount of fish or
other species caught or available per area of mangrove (Baran 1999; Pauly and
Ingles 1999) or the relative contribution of an area to a given harvest (Pauly and
Ingles 1999; Rönnbäck 1999).
Figures 19.7 and 19.8 presents estimates of supporting service values of man-
grove ecosystems based on 14 reference materials. Mangroves play a very signifi-
cant role in maintaining the life cycle of migratory species with an estimated average
value of 1725 USD/ha/year. Meanwhile, the value of maintaining biodiversity of
mangrove ecosystems is estimated at 82.41 USD/ha/year.
19 Ecological Valuation and Ecosystem Services of Mangroves 451

Fig. 19.7 A crab shielded in a mangrove tree hollow (Image source: Vo Van Thanh)

82.41
Maintenance of genetic diversity (6)

Maintenance of life cycles of 1725.00


migratory species (8)

0 500 1000 1500 2000 2500

Values (USD/ha/year)

Fig. 19.8 Values of supporting services of mangrove ecosystem (Based on data compiled by
Himes-Cornell et al. 2018)

A mangrove ecosystem provides local communities with many opportunities for


aesthetic and recreational experiences, cultural and artistic inspiration, and spiritual
and religious enrichment (Mastaller 1997; Rönnbäck et al. 2007). These cultural
services are often intimately tied to local culture, heritage, and traditional knowledge
452 H. T. Pham et al.

Fig. 19.9 Mangrove ecotourism in Morondova, Madagascar (Image source: Pham Hong Tinh)

4595.03
Opportunities for recreation and tourism (10)

276.15
Information for cognitive development (4)

Aesthetic information (1) 500

0 2000 4000 6000 8000 10000

Values (USD/ha/year)

Fig. 19.10 Values of cultural services of mangrove ecosystems (Based on data compiled by
Himes-Cornell et al. 2018)

(Walters et al. 2008). However, cultural services are generally difficult to quantify or
map, as they are determined by the appreciation and sentiments of local people (Van
Oudenhoven et al. 2014).
A number of studies have been conducted to value cultural services provided by
mangrove ecosystems. These values vary greatly among different cultural services or
among different studies of the same service. Figures 19.9 and 19.10 shows estimated
cultural values of a mangrove ecosystem, in which opportunities for recreation and
tourism see the highest estimated average value of 4595.03 USD/ha/year, followed
by aesthetic information at 500 USD/ha/year and information for cognitive develop-
ment at 276.15 USD/ha/year.
19 Ecological Valuation and Ecosystem Services of Mangroves 453

19.6 Perspective and Conclusion

Mangrove ecosystems play important roles in the sustainable development of coastal


zones through their ecosystem services, such as provisioning services, regulating
services, supporting services, and cultural services. The economic values of those
mangrove ecosystem services have been estimated in many parts of the world. In
which, some services such as providing foods and raw materials, moderation of
extreme events, erosion prevention, and climate regulation are frequently evaluated,
while others, such as pollination and ornamental resources and the cultural category
of ecosystem services have not yet been given adequate attention. Various evalua-
tion methods have been applied to estimate economic values of mangrove ecosystem
services with market price and benefit transfer methods most frequently used.
However, other methods like avoided/damage, choice modelling, travel cost, con-
tingent valuation, etc. are rarely used to evaluate one or several mangrove ecosystem
services. The estimated values also vary greatly across different ecosystem services
or within an ecosystem service, but with different locations and evaluation methods.

References
Aburto-Oropeza O, Ezcurra E, Danemann G, Valdez V, Murray J, Sala E (2008) Mangroves in the
Gulf of California increase fishery yields. Proc Natl Acad Sci 105: 10456–10459.
Adger N, Brown K, Cervigini R, Moran D (1995) Total economic value of forests in Mexico.
Ambio 24: 286–296.
Amarnath JS and Mouna A (2016) Environmental impact assessment of coastal ecosystem in Tamil
Nadu, India with hedonic and travel cost models. Int. J. Mar. Sci 66: 1–8.
Atkinson SC, Jupiter SD, Adams VM, Ingram JC, Narayan S, Klein CC, Possingham HP (2016)
Prioritising mangrove ecosystem services results in spatially variable management priorities.
PLoS One. doi: 10.1371/journal.pone.0151992.
Baran E (1999) A review of quantified relationships between mangroves and coastal resources.
Phuket Marine Biological Center Res Bull 62: 57–64.
Barbier EB, Costanza E, Twilley RR (1991) Guidelines for tropical wetland evaluation. CATIE:
Turrialba, Costa Rica, p 1-57.
Bunting P, Rosenqvist A, Lucas RM, Rebelo L-M, Hilarides L, Thomas N, Hardy A, Itoh T,
Shimada M, Finlayson CM (2018) The Global Mangrove Watch—A New 2010 Global Baseline
of Mangrove Extent. Remote Sens 10(10):1669, doi:https://doi.org/10.3390/rs10101669.
Camacho-Valdez V, Ruiz-Luna A, Ghermandi A, Berlanga-Robles CA, Nunes PALD (2014)
Effects of land use changes on the ecosystem service values of coastal wetlands. Environ
Manage 54: 852–864.
Costanza R, D’Arge R, de Groot R, Farber S, Grasso M, Hannon B, Limburg K, Naeem SV,
O’Neill R, Paruelo J, Raskin RG, Sutton P, van den Belt M (1997) The value of the world’s
ecosystem services and natural capital. Nature 387: 253–260.
De Groot R, Brander L, Van Der Ploeg S, Costanza R, Bernard F, Braat L, et al (2012) Global
estimates of the value of ecosystems and their services in monetary units. Ecosyst Serv 1: 50–61
Himes-Cornell A, Grose SO and Pendleton L (2018) Mangrove ecosystem service values and
methodological approaches to valuation: Where Do We Stand? Front Mar Sci 5: 376.
Ilman M, Wibisono ITC, Suryadiputra INN (2011) State of the art information on mangrove
ecosystems in Indonesia. Wetlands International - Indonesia Programme, Bogor.
454 H. T. Pham et al.

Jerath M, Bhat M, Rivera-Monroy VH, Castaneda-Moya E, Simard M, Twilley RR (2016) The role
of economic, policy, and ecological factors in estimating the value of carbon stocks in
Everglades mangrove forests, South Florida, USA. Environ. Sci Policy 66: 160–169.
Kathiresan K, Bingham BL (2001) Biology of mangroves and mangrove ecosystems. Adv Mar Biol
40: 81–251.
Lebel L, Tri NH, Saengnoree A, Pasong S, Buatama U, Thoa LK (2002) Industrial transformation
and shrimp aquaculture in Thailand and Vietnam: pathways to ecological, social and economic
sustainablity? AMBIO: A Journal of the Human Environment 31(4): 311-323.
Mashayekhi Z, Danehkar A, Sharzehi GA, Majed V (2016) Coastal communities WTA compensa-
tion for conservation of mangrove forests: a choice experiment approach. Knowl Manag Aquat
Ecosyst 417: 20, doi: 10. 1051/kmae/2016007.
Mastaller M (1997) Mangroves, the Forgotten Forest between Land and Sea. Tropical Press Sdn.
Bhd, Kuala Lumpur, Malaysia.
Mazda Y, Magi M, Ikeda Y, Kurokawa T, Asano T (2006) Wave reduction in a mangrove forest
dominated by Sonneratia sp. Wet Ecol Manag 14: 365–378.
MEA (2005) Millennium Ecosystem Assessment. Washington, DC: World Resources Institute.
Mojiol AR, Guntabid J, Lintangah W, Ismenyah M, Kodoh J, Chiang LK, Sompud J (2016)
Contribution of mangrove forest and socio-economic development of local communities in
Kudat District, Sabah Malaysia. Int J Agric For Plant 2: 122–129.
Naylor RL, Goldburg RJ, Primavera JH, Kautsky N, Beveridge MCM, Clay J, Folke C,
Lubchenco J, Mooney H, Troell M (2000) Effect of aquaculture on world fish supplies. Nature
405: 1017–1024.
Pauly D, Ingles J (1999) The relationship between shrimp yields and intertidal vegetation (man-
grove) areas: a reassessment. In Yáñez-Arancibia A, Lara-Domínguez AL (eds) Ecosistemas de
Manglar en América Tropical. Instituto de Ecología A.C., México, UICN/ORMA, Costa Rica,
NOAA/NMFS Silver Spring (MD), p 311-318.
Rönnbäck P (1999) The ecological basis for economic value of seafood production supported by
mangrove ecosystems. Ecol Econ 29: 235–252.
Rönnbäck P, Crona B, Ingwall L (2007) The return of ecosystem goods and services in replanted
mangrove forests: perspectives from local communities in Kenya. Environmental Conservation
34: 313-324.
Saenger P (2002) Mangrove ecology, silviculture and conservation. Kluwer, Dordrecht.
Sheridan P, Hays C (2003) Are mangroves nursery habitat for transient fishes and decapods?
Wetlands 23: 449–458.
Sopheak K, Hoeurn C (2016) An estimation of the production function of fisheries in Peam Krasaob
Wildlife Sanctuary in Koh Kong Province, Cambodia. EEPSEA Research Report No 2016022.
Laguna: Economy and Environment Program for Southeast Asia (EEPSEA).
Susilo E, Purwanti P, Lestariadi RA (2016) Mangrove management in Dumas beach: economic and
institutional analysis. Int J Manag Adm Sci 3: 11–24.
Tri NH (2006) Mangrove ecosystem service valuation – principle and application. Pubishing House
of Vietnam National Economic University, Hanoi, Vietnam (in Vietnamese).
Tri NH, Adger WN, Kelly PM (1998) Natural resource management in mitigating climate impacts:
the example of mangrove restoration in Vietnam. Glob Environ Change 8(1): 49-61.
Tri NH, Hong PN, Adger WN, Kelly PM (2002) Mangrove conservation and restoration for
enhanced resilience. In Rapport DJ, Lasley WL, Rolston DE, Nielsen NO, Qualset CO,
Damania AB (eds) Managing for healthy ecosystems. Lewis Publishers, Boca Raton, Florida,
USA, p 389–402.
Van Oudenhoven APE, Siahainenia AJ, Sualia I, Tonneijck FH, Van der Ploeg S, de Groot RS
(2014) Effects of different management regimes on mangrove ecosystem services in Java,
Indonesia. Wageningen University (Wageningen) and Wetlands International (Ede & Bogor).
Walters BB, Rönnbäck P, Kovacs JM, Crona B, Hussain SA, Badola R, Primavera JH, Barbier E,
Dahdouh-Guebas F (2008) Ethnobiology, socio-economics and management of mangrove
forests: A review. Aquat Bot 89: 220–236.
Zhang K, Liu H, Li Y, Xu H, Shen J, Rhome J, Smith TJ III (2012) The role of mangroves in
attenuating storm surges. Estuar Coast Shelf Sci 102-103: 11–23.
Management Action Plans for Development
of Mangrove Forest Reserves 20
Mohamad Danial Md Sabri, Mohd Nazip Suratman, and
Nur Hajar Zamah Shari

Abstract

Mangroves are recognized as a productive and pivotal ecosystem by providing


habitats for many types of biodiversity, sources of food and medicines, heighten
socio-economic of local communities, contribute towards climate regulations and
coastal protection from sporadic windstorm events. These ecosystems are widely
distributed throughout tropical and subtropical regions of the world. Mangroves
are basically generated from sediment complexes which has evolved over the last
several thousand years. This resulted in a combination of geomorphic processes
operating at a sea level close to its current position and ultimately acting as a
bridge connecting the land and sea. Among tree communities found in this
productive area are from some species such as Rhizophora spp., Avicennia spp.,
and Bruguiera spp. Overexploitation of mangroves through human activities such
as illegal logging and encroachment, wildlife poaching, expansion of agricultural
as well as aquaculture practices, apparently becomes among rapid drivers for
degradation to this majestic ecosystem. Without proper control over these
activities will serve the extinction on mangrove forests. Therefore, proper man-
agement action plans towards establishment of forest reserves in mangrove areas
are needed to overcome such threats. Moreover, the management action plan is
critical to guide policy and decision makers, local authorities, stakeholders, and
communities to play their roles to ensure the sustainability of mangrove forests

M. D. Md Sabri · N. H. Z. Shari
Forestry and Environment Division, Forest Research Institute Malaysia, Kepong, Selangor,
Malaysia
M. N. Suratman (*)
Faculty of Applied Sciences and Institute for Biodiversity and Sustainable Development, UiTM,
Shah Alam, Malaysia
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 455
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_20
456 M. D. Md Sabri et al.

and their resources. This chapter reviews related available policies, process, and
stages that encapsulated within recent scientific studies on mangrove forests
globally towards the development of effective management action plans to
promote the sustainability of mangrove forest reserves.

Keywords

Mangroves forest · Management action plan · Productive ecosystem · Forest


reserves

20.1 Introduction

Mangrove ecosystems ecologically inhabiting the interface between land and sea at
low latitudes which occupy a harsh environment, regularly subject to tidal changes
in temperature, water, and salt exposure, and varying degrees of conditions. Man-
grove forests and their inhabitants are fairly robust and highly adaptable to life in
waterlogged saline soils within warm, subtropical, and tropical seascapes. The
forests occupy approximately 4.79 million ha of global coastline where the largest
areas are found in Asia (5,547,000 ha) followed by Africa (3,240,000 ha) and North
and Central America (2,571,000 ha) regions (FAO 2020). These forests provide
various benefits and services, from the provision of materials (such as timber,
medicinal plants, and fisheries) to being a regulating influence (such as protection
from sporadic windstorm events, and carbon sequestration). Moreover, these forests
offer cultural and spiritual benefits to the surrounding communities.
Albeit crucial for supporting local livelihoods, mangrove ecosystems are becom-
ing increasingly threatened all over the world as they are undergoing rapid degrada-
tion. These mainly due to anthropogenic activities such as urban construction,
infrastructure development for tourism, conversion for aquaculture and agriculture,
overharvesting of mangrove resources, pollution, and human-induced climate
change (Barbier 2016; Arumugam et al. 2020). According to Food and Agriculture
Organization (FAO), the area of global mangroves declined by 1.04 million ha
between 1990 and 2020 where the rate of loss more than halved over three decades.
The global disappearance of mangroves poses a major impact on the vulnerability of
coastal populations and property in developing countries, especially with respect to
damaging and life-threatening storms and floods (Alongi 2008; Spalding et al.
2014).
As mangroves are declining over decades, understanding the consents of a
diverse range of stakeholders in mangrove forest managements is crucial
(Arumugam et al. 2020). The impact of forest management within mangroves
areas is depending on the way of local policies and legislation that are implemented.
Without proper management planning and actions, these important forest resources
will deplete over time. To ensure the sustainable services provide by the mangroves,
proactive actions are needed to compensate mangrove loss by gazetting more
mangroves areas to become permanent forest reserves. Through this initiative,
20 Management Action Plans for Development of Mangrove Forest Reserves 457

mangrove resources will be managed properly according to current legislation.


Therefore, this chapter reviews previous studies on mangrove management and
policies in order to develop an action plan for sustainability of mangrove forest
reserves.

20.2 Current Distribution and Status of Global Mangroves

Mangrove ecosystems are widely recognized for their pivotal role in ecological
sustainability by providing habitat for fish and crustacean of commercial value as
well as for effective sediment trapping, nutrient recycling, and protection of
shorelines from erosion (Datta et al. 2012). Besides that, freshwater floods occur
during the monsoon seasons, while brackish water or seawater floods in other
periods occasionally happened at these productive forest areas (Koh et al. 2018;
Yahya et al. 2020). This situation will be affecting the trend of plant communities,
diversity, and conditions in mangrove forests. These forests are mostly dominated by
halophytic plants and predominantly found along the tropical and subtropical
coastlines which eventually offer important ecosystem functions and services for
local and regional communities (Biswas et al. 2018).
Globally, mangrove forests are distributed within 113 countries and territories
(FAO 2020). In 2010, the mangrove areas were estimated to be 137,600 km2 of
coastline area globally (Bunting et al. 2018) which contributed to the economic
value on the order of 200,000–900,000 USD ha 1 (UNEP-WCMC 2006). Asia
region has recognized to have the largest extent and the highest species diversity
of mangrove ecosystems in the world (ITTO 2002) with the five largest mangrove-
holding nations include Indonesia, Brazil, Malaysia, Papua New Guinea, and
Australia which represent nearly 61.6% of global mangrove cover (Blasco et al.
2001; Hamilton and Casey 2016; Martínez-Espinosa et al. 2020). In addition, other
major mangrove areas are also found in Guinea Bissau, Mozambique, Madagascar,
the Philippines, Thailand, and Vietnam (Giri et al. 2011). Mangrove areas recorded
absent in Europe region as shown in Table 20.1. Based on previous information,
Oceania has the smallest areal extent and Africa has the least species diversity
(Ricklefs and Latham 1993).

Table 20.1 Global mangrove areas by region (FAO 2020)


Region Mangrove area (ha) Percentage (%)
Africa 3,240,000 21.91
Asia 5,547,000 37.52
Europe 0 0.00
North and Central America 2,571,000 17.39
Oceania 1,298,000 8.78
South America 2,130,000 14.41
World 14,786,000 100.00
458 M. D. Md Sabri et al.

Many developing countries nowadays are concerned with the importance of


conserving the mangrove forests and coastal populations for the benefit towards
the local livelihood and ecological-economic values (Datta et al. 2012). These
countries are taking efforts to maintain the extension of their mangrove forests.
The largest extension of mangrove forest recorded in neo-tropical latitudes which
found in Brazil (9627 km2), Mexico (741,917 km2), and Cuba (421,538 km2)
(Suárez-Abelenda et al. 2014). While in Asia, Bangladesh is having a huge area of
mangrove forest around 601,700 ha, which equal to 38.12% of the entire forest land
and 4.13% of the land mass in Sundarbans (Khan et al. 2020). In Pakistan, 1050 km
long coastline encompassing mangrove forests that provide many valuable ecosys-
tem services. These mangroves are found along the entire coastline of Sindh but
occur occasionally on the coast of Balochistan (Farooq and Siddiqui 2020). Mean-
while, in India, vast coastline of 8118 km has a rich mangrove ecosystem that offers
a livelihood to several thousand people across the nation. These mangrove forests
cover 4740 km2 area along the Indian coastline of which 617 sq. km occur in the
Andaman and Nicobar Islands (Kiruba-Sankar et al. 2018). These are among
developing countries which thrive to maintain the existence of mangrove forest
within their political territories.
Mangroves are mainly divided into three major categories which include true
mangroves, mangroves, and mangrove associates (Wan Juliana et al. 2010). The
total number of true mangrove plant species in the world is about 70 and which
belong to 17 families. Apart from that, 51 species are present in the South Asian
mangrove forests (Wan Ismail et al. 2018). Interestingly, Southeast Asia, which can
be perceived as home for world largest mangrove where constitute area approxi-
mately over 6.8 million ha or equal to 34%–42% out of total mangroves worldwide
(Giesen et al. 2007). A study by Giesen et al. (2007) also described that the greatest
areas were found in Indonesia, which has the most extensive coverage of mangroves
(60%), followed by Malaysia (11.7%), Myanmar (8.8%), and Thailand (5%).
Besides that, 268 flora species were recorded, including 129 trees and shrubs,
50 terrestrial herbs consisting of 27 types of grass and grass-like plants, 28 climbers,
28 epiphytes, 24 ferns, 7 palms, 1 screw pine (pandan), and 1 cycad. However, from
all of these, only 52 species were known as true mangrove species where the habitat
is exclusively in mangrove ecosystems.
Ecologically, mangroves which consist of salt tolerant plants comprised of tree,
palm, shrub, and fern communities were found in the intertidal zone, at the interface
of land and sea, transitional zone of the coasts, estuaries, and along rivers draining
into the sea (Hutchison et al. 2014; Shah et al. 2016). Among countries with a high
diversity of mangroves are Indonesia with 45 species followed by Malaysia (40 spe-
cies), Thailand (34 species), and Singapore (31 species) (Kathiresan and Rajendran
2005). In the case of Malaysia, mangrove forests are mainly found in the marine
alluvium along sheltered coasts and estuaries both in Sabah, Sarawak, and Peninsu-
lar Malaysia. Among tree genus distribute along the coast are Rhizophora,
Avicennia, Bruguiera, Sonneratia, and Xylocarpus spp. in species-specific belts
depending on soil and inundation patterns (Wilkie and Fortuna 2003). The presence
of mangrove species and diversity are depending on the soil types, predations,
20 Management Action Plans for Development of Mangrove Forest Reserves 459

disturbance, weather stability, intertidal patterns, management regimes, and more


based on geographical restricted, the region of their origin. Malaysia has shown the
effort to conserve mangrove forests through the establishment of the Matang Man-
grove Forest Reserve where this reserve is considered as the best managed mangrove
forest in the world (Walters et al. 2008).

20.3 Drivers of Mangrove Loss

Mangrove forests, like other ecosystems, are subject to various disturbances that
vary in their nature (e.g., geological, physical, chemical, biological) in time and
space (Alongi 2008). The drivers of these disturbances may come from natural and
anthropogenic sources. However, the great loss of mangrove forests is likely
threatened by a range of anthropogenic activities. Loss of mangrove cover has
increased significantly since 1970 (Giri et al. 2011; Richards and Friess 2016), and
the remaining forest patches are under pressure from different sources of distur-
bance. According to previous records, mangroves are disappearing at a global loss
rate of 1–2% per year (Spalding et al. 2010), and the loss rate reached 35% during the
last 20 years (Polidoro et al. 2010; Carugati et al. 2018). These rates may be as high
as or higher than rates of losses of upland tropical wet forests, and current loss rates
are expected to continue unless mangrove forests are protected as a valuable resource
(Polidoro et al. 2010).
The drivers in discussing this chapter are perceived as major threats for mangrove
loss in many places around the world. Previous studies showed that one quarter of
the world’s mangroves have been lost due to human action, mainly through forest
conversion to aquaculture, agriculture, and urban land uses (Barbier 2016). Other
studies also showed that over 35% of mangroves worldwide have disappeared since
the 1980s, mainly due to clear-cutting and conversion to aquaculture ponds (Valiela
et al. 2001; Richards and Friess 2016). In this context, aquaculture was identified as
the main source of degradation of mangrove forests. Tropical coastlines in Southern
Thailand where the mangroves can be found have been converted into aquaculture
ponds, consequently lost the important ecosystem services (Thampanya et al. 2006).
The depreciation services such as the loss of sediment stabilization is currently
causing problems along multiple aquaculture coasts worldwide in the form of coastal
erosion. Additionally, erosion of aquaculture areas can be as severe as several
kilometres per year where finally putting coastal communities in danger (Van
Wesenbeeck et al. 2015). Furthermore, many large companies abandoned the
ponds once their profits decrease, turning to other pristine mangrove areas for
development of new and more productive ponds (Huitric et al. 2002).
Indonesia had lost more than 200,000 ha of its mangroves by the 1960s due to
various threats where the largest losses were reported from Java and Sumatra (Koh
et al. 2018). Besides that, Indonesia was reported among ten largest fish producers in
the world. In the year of 1999 to 2003, fish production in Indonesia increased by an
average of 8.5% per year, from 4952 thousand tonnes in 1999 to an estimated 5961
thousand tonnes in 2003, with one-fifth of the production coming from aquaculture.
460 M. D. Md Sabri et al.

Fig. 20.1 Aquaculture trend in Indonesia from year 1980–2018 (FAO undated)

Aquaculture is deemed an important source of employment, providing livelihoods to


an estimated around 2.4 million households in this country (http://www.fao.org/
fishery/countrysector/naso_indonesia/en). Figure 20.1 shows that the trend of aqua-
culture in Indonesia from the year 1980 to 2018 as reported by FAO FishStat.
Halting the aquaculture activity will affect some part of local economics, thus,
proper management and action are believed to be crucial to ensure the availability
of these mangrove forests to upkeep the ecosystem services.
Another major driver in mangrove loss is shrimp farming. Shrimp farming is
among popular aquaculture activity which leads to land use changes that occurred in
mangrove areas. Eastern African mangroves are currently facing additional threat
from shrimp aquaculture, where the practices were originating from Southeast Asia.
This shrimp aquaculture practices are now reaching to West Africa (Ajonina et al.
2008). While in Brazil, shrimp farming production in tropical coastal regions has
increased significantly during the last 30 years due to food demand and high
economic value (Queiroz et al. 2013). Despite it is now being recognized that the
conversion of mangrove forests to shrimp farming becomes unsustainable in the
long term due to the environmental impacts (e.g., eutrophication) and strong disrup-
tion of local economies (Suárez-Abelenda et al. 2014), some countries are politically
extending the area to provide values for locals. Conversely, recent studies came out
with shrimp farming model development for Vietnamese Mekong Delta to reduce
the ecological impact from this activity. This model is beneficial to the local
livelihoods and together with enhancement of knowledge in coastal protection
(Nguyen et al. 2020). In this context, shrimp farming also plays greater role in
mangrove forests with some pros and cons on the impacts to the ecosystem or socio-
economic well-being.
20 Management Action Plans for Development of Mangrove Forest Reserves 461

Forest degradation is also believed to be a serious issue in management of


mangrove ecosystems. Although the rate of global mangrove loss has declined, the
future of global mangrove still remains uncertain as new territories of deforestation
are opening up, mainly in Southeast Asia and West Africa (Arumugam et al. 2020).
Southeast Asia, the region of largest and most biodiverse mangrove ecosystems on
Earth (Spalding et al. 2010; Giri et al. 2011; Appeltans et al. 2012), is currently
facing a highest rate of deforestation since 2000 (Hamilton and Casey 2016;
Richards and Friess 2016). While in Africa, mangroves also have been subjected
to enormous pressures and threats within the last past decades with great losses, for
example, over 20–30% of the mangroves notably in the west and central Africa have
lost since past 25 years. The loss happened through many factors, especially
deforestation for fish-smoking (Ajonina et al. 2008). Another study in Bangladesh,
Sundarbans mangrove forests were reported as being continuously degraded where
total tree cover has been reduced by 50 per cent over the past 20 years (Khan et al.
2020). High demand, poor management, and overexploitation within mangrove
forests without replacement plan were associated with severe degradation to the
ecosystems.

20.4 Importance of Mangrove Conservation

Mangroves are forested wetlands that are uniquely adapted to the intertidal zone.
These forests need to be conserved for the benefits of the local population and
environment. Among salient services offered by these majestic ecosystems are
carbon sequestration, supporting fisheries, providing timber and coastline protection
against erosion and flood protection from storm surges and tsunamis (Alongi 2008;
Van Bijsterveldt et al. 2020). Besides that, salt marshes and mangroves contribute to
coastal protection by reducing wave energy, increasing sedimentation, and/or reduc-
ing erosion and movement of sediments (as reviews in Gedan et al. 2011; Spalding
et al. 2014). Furthermore, thick vegetation cover reduces water flow velocities,
turbulent flows, and shear stress over the sea bed, promoting sediment deposition,
which can create accretion. In some cases, deposition stimulates below ground root
production (McKee and Cherry 2009), and these roots scientifically found to
improve soil cohesion and tensile strength, slowing rates of erosion at marsh
edges. Hence, the roots themselves can also present a physical barrier between the
water and soil, particularly in places where root systems extend below low tide levels
(Gedan et al. 2011; Giri et al. 2011). These ecosystems also play a role by creating
habitats for a variety of terrestrial fauna and providing various supplies for local
communities (Shrestha et al. 2019). By focussing to fauna, as birds migrate to
shallow water and muddy areas such as mangrove forests for food, resting areas,
and protection as stated in a review by Azimah and Tarmiji (2018). Thus, fauna also
needs mangroves for shelter and resting to sustain their species. Meanwhile, high
level of fauna biodiversity as over 80% of commercial fisheries and other aquatic
species spent most or part of their life cycle in the mangroves; ecologically, they play
a crucial role in fertilization, stabilization, filtration, regulation of microclimate and
462 M. D. Md Sabri et al.

acting as food chain support and as nurseries for many fish and invertebrate species
economically, they provide a wide range of timber and non-timber forest products
that support rural economies and having high ecotourism potentials (Ajonina et al.
2008).
From global climate regulation point of view, mangrove forests contribute their
roles as a part of coastal carbon sequestration. Mangroves have been estimated to
have higher amounts of carbon than the other types of forest, with a storage capacity
of between 990 and 1074 tonne C ha 1 (Donato et al. 2011). Another impressive fact
that although plant biomass in the ocean and coastal areas comprises only 0.05% of
the total plant biomass on land, it cycles a comparable amount of carbon each year
(Bouillon et al. 2008). Surprisingly, a typical hectare of mangrove has the potential
to release carbon similar up to 3–5 ha of terrestrial tropical forest (Eong 1993). This
shows that mangrove forests are having the capability to store large carbon that
useful for climate regulation and climate change mitigation. In summary, important
services provided by mangrove forests that benefit coastal communities is shown in
Fig. 20.2.

20.5 Management Action Plans for Mangrove Forest Reserves


Development

Mangroves like other natural resources require supportive and adaptive policies,
which are adequately enforced. These policies have to be accompanied by initiatives
such as plans to facilitate conflict resolution between stakeholders and encourage
consultations with administrative officials (Campredon and Cuq 2001; Feka and
Ajonina 2011). As discussed previously, mangroves provide many services to
coastline communities where the extension of the area is deemed requisite for all
mangrove countries. Meanwhile, there were so many drivers in mangrove loss and
ultimately reduce the area of these majestic ecosystems. Hence, the proposed draft
action plan is a strategic guide with a sequence of steps that forest managers could
follow in order to develop future mangrove forest reserves. However, any action
plan to be used should refer to the implemented legislation which taken locally.
Different countries may have different approach to establish or maintain their
mangrove areas. The next section proposes an action plan to develop mangrove
forests reserve that may provide some insights for decision and policy makers and
forest managers.

20.5.1 Action Plans

The action plan for development of mangrove forest reserves template follows six
main steps, viz; (i) Site selection, (ii) Scientific expedition, (iii) Stakeholder consul-
tation, (iv) Determining boundaries, (v) Gazettement, and (vi) Enforcement. Each of
the steps then leads to a number of sections which help to define the context,
operationalize the development of the plan, and implement the measures
20 Management Action Plans for Development of Mangrove Forest Reserves 463

Fig. 20.2 Actual and


potential services are offered Ecological/protective
by mangroves (Datta et al.
2012) • Sediment trapping
• Source of nutrients and organic matter
• Carbon sequestration
• Wastewater treatment
• Protection of shorelines
• Wildlife habitat

Commercial/rawmaterial/subsistence

• Plant parts as food


• Fodder
• Fuel (Charcoal)
• Construction material
• Medicine
• Honey, wax and alcohol
• Tannin and dyes
• Rayon and paper
• Industrial chemicals
• Ornamental goods
• Fishery
• Shrimp culture
• Tourism
• Butterfly farming

Aesthetic/miscellaneous

• Landscaping
• Shelter or fishing boats during storms
• Cultural/religious importance

(Fig. 20.3). This preceding section describes the steps and actions in detail. The
proposed template is defined with the circular approach to foster the continuous
improvement of the developed action plan.

20.5.2 Site Selection

The first step focuses on the context of identifying suitable sites which is fundamen-
tal to understand the main features and characteristic of the forest, in terms of its
situation and function. They are described in the preceeding subsections.
464 M. D. Md Sabri et al.

Fig. 20.3 Proposed action


plan overview
Site
selection

Scientific
Enforcement
expedition

Action plan for


development
mangrove forest
reserves

Stakeholders
Gazettement
consultation

Determining
boundaries

20.5.2.1 Windstorm Barrier and Coastal Protection


The aim is to analyse the frequency of the windstorms and cyclone events on
particular coastal areas and their damaging effects to the local communities. In
fact, more than 1.4 billion people live in coastal areas at a risk for tropical cyclone
(Dilley et al. 2005). Cyclone and windstorm events are expected to increase as a
result of more frequent high-intensity storms created by climate change and
increased exposure created by the ongoing movement of people and assets to
high-risk coastal areas (Del Valle et al. 2020). As an alternative coastal defence,
conservation and restoration of natural habitats that can provide protection and act as
a barrier against windstorm, mangroves forest may be considered as their plant’s
aerial root and canopy structure makes them capable of reducing wave action, wind
velocity, and storm surge (Massel et al. 1999; Mazda et al. 2006; Barbier et al. 2008,
2011; Krauss et al. 2009; Zhang et al. 2012; Das and Crépin 2013; Liu et al. 2013;
Horstman et al. 2014). Thus, the function of this forest as a protector of the coastline
should be considered on priority to conserve the areas.

20.5.2.2 Act as Carbon Storage for Climate Regulation


Mangroves forests play an important role as a substantial carbon sink and there are
categorized among the most carbon-rich forests in the tropics (Laffoley and
Grimsditch 2009; Hamdan et al. 2013). The degradation of this forest will be
possibly intensifying decomposition of stored carbon and finally, emitting large
amounts to serve greenhouse gases in the atmosphere. In terms of ecological view,
20 Management Action Plans for Development of Mangrove Forest Reserves 465

carbon from the atmosphere is absorbed by trees and plants through photosynthesis
process and stored the carbon in the form of food within their stems and other organs.
By this process, mangrove trees and plants also play role in climate regulation and
climate change mitigation. Conservation of the mangrove forests would be vital in
any efforts to address climate change. Responsible authorities may identify large
mangrove areas in their region through modern equipment such as remote sensing.
The targeted large area will probably store large amount of carbon.

20.5.2.3 Coastal Demographic Challenges


The coastal demographic context in which the mangrove lies is crucial both with
reference to present and future scenarios. This information would contribute signifi-
cantly to the understanding of future suburban and territorial development. Man-
grove ecosystems provide numerous benefits to local communities and adjacent
environments. They represent great direct economic importance to the livelihoods
of millions of coastal residents throughout the world who harvest marine resources;
extract timber for construction, firewood, and charcoal production; and cultivate
mangrove honey (Suman 2019). Mangrove ecosystems must compete with powerful
economic interests supported by national and local elites, such as urban residential
developments, airports, power plants, construction and expansion of ports, aquacul-
ture activities, and coastal tourism projects (Dale et al. 2014). This action aims to
provide demographic challenge resulted from local communities within proximity to
the mangrove areas in which can be useful in decision makings in development of
mangrove forest reserves. Mangrove areas which are close to these challenging areas
may not be worth considered as mangrove forest reserves. However, this idea is
totally depending on how those natural habitats are being used.

20.5.2.4 Ecotourism and Recreation Prospects


The selection of future mangrove forest reserves may consist of ecotourism and
recreation values. A previous study by Friess (2017) and numerous local case studies
show that tourist activities can have several important benefits for mangrove educa-
tion, protection, and conservation. However, the benefits of ecotourism must be
balanced against its direct and indirect negative impacts on the mangrove ecosystem
and the local communities that use them. However, strategies are needed to minimize
the impacts from both activities, for example, mitigating the direct environmental
impacts of increased numbers of tourist and boat traffic, and the indirect impacts that
a sudden and large source of tourism income can have on local communities and
economic enterprises. Hence, this action aimed to the authorities to identify areas
which may allow public to enjoy the recreation activities within the forest areas and
ultimately create awareness on the importance of mangrove conservation among
them and provide attractive economic returns to the local communities.
466 M. D. Md Sabri et al.

20.5.3 Scientific Expedition

This section describes the next step for development of the mangrove forest reserve
where the scientific information may be useful to provide scientific figures, supports
and advise for all stakeholders.

20.5.3.1 Speeding up Scientific Research Activities and Collaborative


Research
This action aimed to record all biological diversity, scientific findings, and social
studies within the selected mangrove areas. Scientific research by one agency or
collaboration with other agencies such as universities and research institutes may
bring compilation of observations, knowledge, invent solutions, and create problem
solving within those mangrove areas. It is important for scientific research to occur at
a local level because research from one area may not be applicable to the context and
needs of another region. For example, not all flora and fauna exist in all areas,
sometime, they are endemic to some areas. An endemic species is important because
they are in the habitats restricted to a particular area due to climate change, urban
development or other occurrences.

20.5.3.2 Strengthen International Cooperation in Research, Protection,


and Restoration
Related ministerial and government authorities which own the natural resources may
encourage international involvement and cooperation to achieve conservation
efforts. Engagement of international bodies, for example, the United Nations
Forum on Forests (UNFF), Convention of International Trade in Endangered Spe-
cies of Wild Fauna and Flora (CITES), Convention on Biological Diversity (CBD),
United Nations Framework Convention on Climate Change (UNFCCC), and other
international authorities could gain result in the protection of large areas of man-
grove forests locally and globally. They may also encourage global private sectors to
support through monetary values for research, conservations, forest restoration, and
climate change mitigations efforts through a number of useful mechanisms.

20.5.4 Stakeholders’ Consultation

This step is fundamental for the recognition of the importance of seeking feedback
and understanding the views of those who affected from any alteration of policies,
enforcement, and way of life. Public services have embraced the approach, seeking
the involvement of the public in the development and shaping future services to
particular communities. Therefore, stakeholder consultation becomes a requirement
in the successful development of public policy and services. Following are the
actions under the third step of proposed action plan in a way to develop mangrove
forest reserves.
20 Management Action Plans for Development of Mangrove Forest Reserves 467

20.5.4.1 Reinforce Knowledge the Role, Value of Mangroves


Management and Sustainability
This action will promote sustainable management by developing knowledge on
effective protection and rehabilitation of mangrove forests to all stakeholders espe-
cially to local communities. Authorities should increase public awareness and
education on the benefits of mangroves forests and the importance to sustain these
areas. A study in Guyana has described if community members do not have a very
good knowledge of the roles of mangroves, it could severely impact the way that the
community values the mangrove resources and hence affect their willingness to be
involved in conservation efforts (Da Silva 2015). Participation from local
communities together with authorities will probably provide a chance for mangrove
areas being protected from severe anthropogenic activities.

20.5.4.2 Transform Perception of Managers in Localities on the Role


and Values of Mangrove
The objective of this action is to train and facilitate all managers from government or
private agencies in localities on the importance of mangrove forest protection. In
addressing government policies to the local communities, they are responsible to
have a comprehensive view of the services from these forest areas provided to
humankind, especially on income generation, poverty alleviation, and the provision
of rural food security. For this action, ministerial level should provide them with
necessary trainings to establish the administrative capacity for the management of
mangroves and ability to balance between human needs and the limit of nature.

20.5.4.3 Socialize Forestry and Upgrade Living Standard for Local


People
Local people basically are using mangroves as sources of economic well-being such
as establishment of aquaculture ponds, timber products, source of medicine, infra-
structure developments, and more. However, lack of basic knowledge on the impor-
tance of conservation can lead them to think that conservation efforts may be curbing
their daily activities. In addition, conservation actions are underfunded in many
areas, especially in Southeast Asia, particularly for the marginalized ecosystems
such as mangrove forests (Friess 2016). This situation may drive away the initiative
of local communities to conserve mangrove areas. Action from relevant authorities
to socialize forestry and efforts to upgrade living standards for local communities is
required. Effort of preserving mangrove areas may provide many benefits through
future collaboration approaches between government and private sectors such as
corporate social responsibility. This effort will boost economic values of these
mangrove areas to be an education centres and ecotourism hotspots and ultimately
increase the income of the local communities.

20.5.4.4 Consolidation of Mangrove Management Systems at all Levels


In this action, all relevant authorities from different agencies, which are responsible
in management of mangrove areas, should be gathered with local communities to
discuss and review current policies in aspect of forestry production, ecological, and
468 M. D. Md Sabri et al.

Decision maker: users, politicians, scientists

Definition of requested Comparison of


conditions (e.g. protection of achieved & requested
mangrove areas, preservation situations (e.g. with
of social & economic sustai- similar mangrove areas
nability) in natural conditions)

Definition of possible
management actions (e.g.
controlled wood utilization,
restablishment of natural Implementation of desired
hydrological regime in management actions (e.g.
degraded areas) enacting licence for
supervised wood extraction)

Evaluation of consequences (e.g. assessment of ecological


impact of forest dynamic and on internal (mangrove
communities and external (trade with the city) social structures

Fig. 20.4 Example of a decision support scheme, applied to a specific situation combining
mangrove protection and use (Lara et al. 2010)

economic sustainability. Feedback from all these intersectoral entities should be


consolidated prior the establishment of mangroves forest reserves. For example, any
decision making have to include of many various parties, for example, resource
users, politicians, and scientists who should jointly define the requested systems
conditions, such as the protection of mangrove areas and the preservation of the
social and economic sustainability of a low-income population (Lara et al. 2010)
(Fig. 20.4).

20.5.4.5 Brief about the Proposed Action Plan to all Stakeholders


This section gives some insight to all stakeholders about measures and implementa-
tion of the action plan to their areas. These management action plans for develop-
ment of mangrove forest reserves should be discussed through a participative
process and through agreements among all stakeholders in the particular mangrove’s
areas. This is to ensure successful implementation of the action plan and lessen for
future public resistance.
20 Management Action Plans for Development of Mangrove Forest Reserves 469

20.5.5 Determining Boundaries

The next step is determining the boundaries of selected mangrove areas. Boundaries
are important components of spatially heterogeneous areas. Boundaries are the zones
of contact that arise whenever areas are partitioned into patches (Cadenasso et al.
2003). Natural geographic features should be selected to define the boundaries of a
forest management unit, including of rivers, streams, shorelines, ridges, and spurs.
Besides that, permanent and clearly defined roads, railways, and tracks may also be
used to mark boundaries (Armitage 1998).

20.5.5.1 Marking Selected Mangrove Areas on the Map


This section is to mark on paper using geographic information system (GIS) and
remote sensing about the proposed mangrove areas that will be submitted for
gazetting process. Maps fulfil a wide range of function in forest management
depending on the type, the amount of detail features represented and their scale. A
map can be used to portray all of the information consist of a “picture” which can be
drawn into the mind of a planner or manager about the shapes and slopes of the
ground, the pattern of streams and rivers, the vegetation cover, and the location and
nature of man-made features of a targeted area (Armitage 1998). The map is
important to be used as supporting documents in the decision makings.

20.5.5.2 Mangrove Areas Demarcation


This action is to determine the internal boundaries for selected mangrove areas that
proposed as forest reserves and the boundaries should be demarcated clearly at the
field. For demarcation process, roads, cut lines, pillars, painted standing trees, and
poles can be used to define the internal boundaries. All trees within internal
boundaries are suggested to be inventoried and mapped. Notices should be erected
showing boundaries of mangrove forest reserves to avoid encroachment by local
communities.

20.5.6 Gazettement

This section describes the next step to develop mangrove forest reserves through
gazetting the selected areas. This gazettement is crucial to ensure all activities within
proposed mangrove forest reserves will be protected by local legislation. The process
gazettement an area needs to be undergoing the step according to available policies,
legislations, and jurisdiction of the countries. A previous study explained that
gazetting is a legal act that integrates the forest in the Permanent Forests Estate as
a part of pre-requisite for a state’s land title. Thus, the gazetted forestland can be
allocated depending to the consent of private domain of other entities, such as a local
municipality (commune) for public interest, social and economic activities or a
private company for commercial logging (Ongolo and Karsenty 2015).
470 M. D. Md Sabri et al.

20.5.6.1 Preparing Authorized Documents for Gazettement


of Mangrove Areas as Forest Reserves
This action aimed to prepare the authorized documents for proposing the mangrove
areas to the national or state forest councils’ consideration as the decision makers. At
this stage, all information gathered through research expedition, study on feedback
of local communities and map is required as supportive documents to be submitted
prior the council meeting.

20.5.6.2 Approval of Gazettement of Mangroves Areas


After getting approval from the council, the next action involves relevant authorities
where in this case, forest manager is to set up the management plan to be
implemented development of the mangroves forest reserve. Land use zoning in the
forest reserves may sound crucial for monitoring purposes. The zoning approach
provides important information for potential stakeholders to identify suitable zones
for the optimal allocation of resources and minimization of conflicts among users.
Besides that, forest managers need to inform the stakeholders about the gazettement
areas through advertisement, workshops, and convention. At this stage, forest
manager and other authorities need to plan activities that may be useful for local
communities.

20.5.7 Enforcements

This section is the final step which is related to the enforcement to take place at the
forest reserves to control illegal usage of the areas. Such a management measure
would aim to avoid a system shift towards an ecologically unsustainable situation.
The national or state might confront elites (e.g., tourist entrepreneurs, land
speculators, and logging companies) and take action against legislative abuses by
these powerful groups. The motivation for strict law enforcement should be needed,
perhaps to warn other mangrove users to respect the law. This may support the
ecological sustainability of the system. Forest managers need to set up enforcement
unit to monitor the forest reserve areas.

20.6 Conclusion

There are many ways to protect mangrove forests from being degraded by anthro-
pogenic activities. However, despite the enforcement with strict law is implemented
to protect these forests, many previous studies shown above described that the
number of mangrove areas is declining every year globally. The loss of these
mangroves should be replaced with available mangroves forest. However, no
standardized plans are found to be referred by forest managers to develop mangroves
forest reserves. This important gap in the literature has been identified for this
purpose. The proposed action plan is considered to be a comprehensive guideline
for the development of mangrove forest reserves as a part of the management plan.
20 Management Action Plans for Development of Mangrove Forest Reserves 471

The step-by-step action plans will facilitate forest managers and all stakeholders to
address the technical, economic, social, and environmental sustainability that are
useful for sustainable mangrove forest management.

References
Ajonina G, Diamé A, Kairo J (2008) Current status and conservation of mangroves in Africa: An
overview. World Rainforest Movement Bulletin No 133.
Alongi DM (2008) Mangrove forests: Resilience, protection from tsunamis, and responses to global
climate change. Estuar Coast Shelf Sci 76:1–13. https://doi.org/10.1016/j.ecss.2007.08.024
Appeltans W, Ahyong ST, Anderson G, et al (2012) The magnitude of global marine species
diversity. Curr Biol 22:2189–2202. https://doi.org/10.1016/j.cub.2012.09.036
Armitage I (1998) Guidelines for the management of tropical forests 1. The production of wood.
FAO, Rome
Arumugam M, Niyomugabo R, Dahdouh-Guebas F, Hugé J (2020) The perceptions of stakeholders
on current management of mangroves in the Sine-Saloum Delta, Senegal. Estuar Coast Shelf
Sci. https://doi.org/10.1016/j.ecss.2020.106751
Azimah AR, Tarmiji M (2018) Habitat requirements of migratory birds in the Matang Mangrove
Forest Reserve, Perak. J Trop For Sci 30:304–311. https://doi.org/10.26525/jtfs2018.30.3.
304311
Barbier EB (2016) The protective service of mangrove ecosystems: A review of valuation methods.
Mar Pollut Bull 109:676–681. https://doi.org/10.1016/j.marpolbul.2016.01.033
Barbier EB, Hacker SD, Kennedy C, et al (2011) The value of estuarine and coastal ecosystem
services. Ecol Monogr 81(2):169–193
Barbier EB, Koch EW, Silliman BR, et al (2008) Coastal ecosystem-based management with
nonlinear ecological functions and values. Science (80–) 319:321–323. https://doi.org/10.
1126/science.1150349
Biswas SR, Biswas PL, Limon SH, et al (2018) Plant invasion in mangrove forests worldwide. For
Ecol Manage 429:480–492. https://doi.org/10.1016/j.foreco.2018.07.046
Blasco F, Aizpuru M, Gers C (2001) Depletion of the mangroves of Continental Asia. Wetl Ecol
Manag 9:245–256. https://doi.org/10.1023/A:1011169025815
Bouillon S, Borges A V., Castañeda-Moya E, et al (2008) Mangrove production and carbon sinks:
A revision of global budget estimates. Global Biogeochem Cycles 22:1–12. https://doi.org/10.
1029/2007GB003052
Bunting P, Rosenqvist A, Lucas RM, et al (2018) The global mangrove watch—A new 2010 global
baseline of mangrove extent. Remote Sens 10. https://doi.org/10.3390/rs10101669
Cadenasso ML, Pickett STA, Weathers KC, Jones CG (2003) A framework for a theory of
ecological boundaries. Bioscience 53:750–758. https://doi.org/10.1641/0006-3568(2003)053[
0750:AFFATO]2.0.CO;2
Campredon P, Cuq F (2001) Artisanal fishing and coastal conservation in West Africa. J Coast
Conserv 7:91–100. https://doi.org/10.1007/BF02742471
Carugati L, Gatto B, Rastelli E, et al (2018) Impact of mangrove forests degradation on biodiversity
and ecosystem functioning. Sci Rep 8:1–11. https://doi.org/10.1038/s41598-018-31683-0
Da Silva P (2015) Exploring a community’s knowledge and use of a coastal mangrove resource:
The case of Wellington Park, Guyana. Int J Sci Environ Technol 4:759–769
Dale PER, Knight JM, Dwyer PG (2014) Mangrove rehabilitation: a review focusing on ecological
and institutional issues. Wetl Ecol Manag 22:587–604. https://doi.org/10.1007/s11273-014-
9383-1
Das S, Crépin AS (2013) Mangroves can provide protection against wind damage during storms.
Estuar Coast Shelf Sci 134:98–107. https://doi.org/10.1016/j.ecss.2013.09.021
472 M. D. Md Sabri et al.

Datta D, Chattopadhyay RN, Guha P (2012) Community based mangrove management: A review
on status and sustainability. J Environ Manage 107:84–95. https://doi.org/10.1016/j.jenvman.
2012.04.013
Del Valle A, Eriksson M, Ishizawa OA, Miranda JJ (2020) Mangroves protect coastal economic
activity from hurricanes. Proc Natl Acad Sci U S A 117:265–270. https://doi.org/10.1073/pnas.
1911617116
Dilley M, Chen RS, Deichmann U, et al (2005) Natural Disaster Hotspots A Global Risk Analysis.
The International Bank for Reconstructio and Development/The World Bank and Columbia
University, Washington, DC
Donato DC, Kauffman JB, Murdiyarso D, et al (2011) Mangroves among the most carbon-rich
forests in the tropics. Nat Geosci 4:293–297. https://doi.org/10.1038/ngeo1123
Eong OJ (1993) Mangroves—A carbon source and sink. Chemosphere 27:1097–1107. https://doi.
org/10.1016/0045-6535(93)90070-L
FAO (2020) Global Forest Assessment Resources 2020: Main report. Rome
Farooq S, Siddiqui PJA (2020) Assessment of three mangrove forest systems for future manage-
ment through benthic community structure receiving anthropogenic influences. Ocean Coast
Manag 190:105162. https://doi.org/10.1016/j.ocecoaman.2020.105162
Feka NZ, Ajonina GN (2011) Drivers causing decline of mangrove in West-Central Africa: A
review. Int J Biodivers Sci Ecosyst Serv Manag 7:217–230. https://doi.org/10.1080/21513732.
2011.634436
Friess DA (2016) Mangrove forests. Curr. Biol. Mag. R746–R748
Friess DA (2017) Ecotourism as a Tool for Mangrove Conservation. Sumatra J Disaster, Geogr
Geogr Educ 1:24–35
Gedan KB, Kirwan ML, Wolanski E, et al (2011) The present and future role of coastal wetland
vegetation in protecting shorelines: Answering recent challenges to the paradigm. Clim Change
106:7–29. https://doi.org/10.1007/s10584-010-0003-7
Giesen W, Wulffraat S, Zieren M, Scholten L (2007) Mangrove guidebook for Southeast Asia. FAO
Regional Office for Asia and the Pacific, Bangkok
Giri C, Ochieng E, Tieszen LL, et al (2011) Status and distribution of mangrove forests of the world
using earth observation satellite data. Glob Ecol Biogeogr 20:154–159. https://doi.org/10.1111/
j.1466-8238.2010.00584.x
Hamdan O, Hasmadi IM, Aziz HK (2013) Mangrove carbon stock assessment by optical satellite
imagery. J Trop For Sci 25:554–565
Hamilton SE, Casey D (2016) Creation of a high spatio-temporal resolution global database of
continuous mangrove forest cover for the 21st century (CGMFC-21). Glob Ecol Biogeogr
25:729–738. https://doi.org/10.1111/geb.12449
Horstman EM, Dohmen-Janssen CM, Narra PMF, et al (2014) Wave attenuation in mangroves: A
quantitative approach to field observations. Coast Eng 94:47–62. https://doi.org/10.1016/j.
coastaleng.2014.08.005
Huitric M, Folke C, Kautsky N (2002) Development and government policies of the shrimp farming
industry in Thailand in relation to mangrove ecosystems. Ecol Econ 40:441–455. https://doi.
org/10.1016/S0921-8009(02)00011-3
Hutchison J, Spalding M, zu Ermgassen P (2014) The role of mangroves in fisheries enhance-
ment. The Nature Conservancy, Bethesda, Maryland and Wetlands International, Wageningen,
Netherlands. 54 pp.
ITTO (2002) ITTO Mangrove Workplan 2002–2006. Yokohama, Japan
Kathiresan K, Rajendran N (2005) Coastal mangrove forests mitigated tsunami. Estuar Coast Shelf
Sci 65:601–606. https://doi.org/10.1016/j.ecss.2005.06.022
Khan MFA, Rahman MS, Giessen L (2020) Mangrove forest policy and management: Prevailing
policy issues, actors’ public claims and informal interests in the Sundarbans of Bangladesh.
Ocean Coast Manag 186:105090. https://doi.org/10.1016/j.ocecoaman.2019.105090
20 Management Action Plans for Development of Mangrove Forest Reserves 473

Kiruba-Sankar R, Krishnan P, Dam Roy S, et al (2018) Structural complexity and tree species
composition of mangrove forests of the Andaman Islands, India. J Coast Conserv 22:217–234.
https://doi.org/10.1007/s11852-017-0588-3
Koh H., Teh S., Kh’ng X, Raja Barizan R (2018) Mangrove Forests : Protection Against and
Reselience to Coastal Disturbance. J Trop For Sci 30:446–460
Krauss KW, Doyle TW, Doyle TJ, et al (2009) Water level observations in mangrove swamps
during two hurricanes in Florida. Wetlands 29:142–149. https://doi.org/10.1672/07-232.1
Laffoley D, Grimsditch G (2009) The Management of Natural Coastal Carbon Sinks. Gland,
Switzerland, Switzerland
Lara RJ, Cohen M, Szlafsztein C (2010) Drivers of Temporal Changes in Mangrove Vegetation
Boundaries abd Consequences of Land Use. In: Caldwell MM, Heldmaier G, Jackson RB, et al.
(eds) Mangrove Dynamic and Management in North Brazil. Springer-Verlag, London,
pp. 127–141
Liu H, Zhang K, Li Y, Xie L (2013) Numerical study of the sensitivity of mangroves in reducing
storm surge and flooding to hurricane characteristics in southern Florida. Cont Shelf Res
64:51–65. https://doi.org/10.1016/j.csr.2013.05.015
Martínez-Espinosa C, Wolfs P, Vande Velde K, et al (2020) Call for a collaborative management at
Matang Mangrove Forest Reserve, Malaysia: An assessment from local stakeholders’ view
point. For Ecol Manage 458:117741. https://doi.org/10.1016/j.foreco.2019.117741
Massel SR, Furukawa K, Brinkman RM (1999) Surface wave propagation in mangrove forests.
Fluid Dyn Res 24:219–249. https://doi.org/10.1016/S0169-5983(98)00024-0
Mazda Y, Magi M, Ikeda Y, et al (2006) Wave reduction in a mangrove forest dominated by
Sonneratia sp. Wetl Ecol Manag 14:365–378. https://doi.org/10.1007/s11273-005-5388-0
McKee KL, Cherry JA (2009) Hurricane katrina sediment slowed elevation loss in subsiding
brackish marshes of the Mississippi river delta. Wetlands 29:2–15
Nguyen HQ, Tran DD, Luan PDMH, et al (2020) Socio-ecological resilience of mangrove-shrimp
models under various threats exacerbated from salinity intrusion in coastal area of the Vietnam-
ese Mekong Delta. Int J Sustain Dev World Ecol 4509:1–14. https://doi.org/10.1080/13504509.
2020.1731859
Ongolo S, Karsenty A (2015) The Politics of Forestland Use in a Cunning Government: Lessons for
Contemporary Forest Governance Reforms. Int For Rev 17:195–209. https://doi.org/10.1505/
146554815815500561
Polidoro BA, Carpenter KE, Collins L, et al (2010) The loss of species: Mangrove extinction risk
and geographic areas of global concern. PLoS One 5. https://doi.org/10.1371/journal.pone.
0010095
Queiroz L, Rossi S, Meireles J, Coelho C (2013) Shrimp aquaculture in the federal state of Ceará,
1970–2012: Trends after mangrove forest privatization in Brazil. Ocean Coast Manag 73:54–62.
https://doi.org/10.1016/j.ocecoaman.2012.11.009
Richards DR, Friess DA (2016) Rates and drivers of mangrove deforestation in Southeast Asia,
2000–2012. Proc Natl Acad Sci U S A 113:344–349. https://doi.org/10.1073/pnas.1510272113
Ricklefs RE, Latham RE (1993) Global patterns of diversity in mangrove floras. In: Species
Diversity in Ecological Communities: Historical and Geographical Perspectives. University of
Chicago Press, Chicago, pp. 215–229
Shah K, Mustafa Kamal AH, Rosli Z, et al (2016) Composition and diversity of plants in Sibuti
mangrove forest, Sarawak, Malaysia. Forest Sci Technol 12:70–76. https://doi.org/10.1080/
21580103.2015.1057619
Shrestha S, Miranda I, Kumar A, et al (2019) Identifying and forecasting potential biophysical risk
areas within a tropical mangrove ecosystem using multi-sensor data. Int J Appl Earth Obs
Geoinf 74:281–294. https://doi.org/10.1016/j.jag.2018.09.017
Spalding M, Kainuma M, Collins L (2010) World Atlas of Mangroves. Erathscan Ltd, London
Spalding MD, Ruffo S, Lacambra C, et al (2014) The role of ecosystems in coastal protection:
Adapting to climate change and coastal hazards. Ocean Coast Manag 90:50–57. https://doi.org/
10.1016/j.ocecoaman.2013.09.007
474 M. D. Md Sabri et al.

Suárez-Abelenda M, Ferreira TO, Camps-Arbestain M, et al (2014) The effect of nutrient-rich


effluents from shrimp farming on mangrove soil carbon storage and geochemistry under
semi-arid climate conditions in northern brazil. Geoderma 213:551–559. https://doi.org/10.
1016/j.geoderma.2013.08.007
Suman DO (2019) Mangrove Management: Challenges and Guidelines. In: Perillo G, Wolansk E,
Cahoon D, Hopkinson C (eds) Coastal Wetland An Integrated Ecosystem Approach, Second
edi. Elsevier, pp. 1055–1079
Thampanya U, Vermaat JE, Sinsakul S, Panapitukkul N (2006) Coastal erosion and mangrove
progradation of Southern Thailand. Estuar Coast Shelf Sci 68:75–85. https://doi.org/10.1016/j.
ecss.2006.01.011
UNEP-WCMC (2006) In the Front Line: Shoreline Protection and Other Ecosystem Services from
Mangroves and Coral Reefs. UNEP-WCMC, Cambridge
Valiela I, Bowen JL, York JK (2001) Mangrove forests: One of the world’s threatened major
tropical environments. Bioscience 51:807–815. https://doi.org/10.1641/0006-3568(2001)051[
0807:MFOOTW]2.0.CO;2
Van Bijsterveldt CEJ, Van Wesenbeeck BK, Van der Wal D, et al (2020) How to restore mangroves
for greenbelt creation along eroding coasts with abandoned aquaculture ponds. Estuar Coast
Shelf Sci 235:106576. https://doi.org/10.1016/j.ecss.2019.106576
Van Wesenbeeck BK, Balke T, van Eijk P, et al (2015) Aquaculture induced erosion of tropical
coastlines throws coastal communities back into poverty. Ocean Coast Manag 116:466–469.
https://doi.org/10.1016/j.ocecoaman.2015.09.004
Walters BB, Rönnbäck P, Kovacs JM, et al (2008) Ethnobiology, socio-economics and manage-
ment of mangrove forests: A review. Aquat Bot 89:220–236. https://doi.org/10.1016/j.aquabot.
2008.02.009
Wan Ismail WN, Wan Ahmad WJ, Salam MR, Latiff A (2018) Structural and floristic pattern in a
disturbed mangrove tropical swamp forest: A case study from the langkawi UNESCO global
geopark forest, peninsular Malaysia. Sains Malaysiana 47:861–869. https://doi.org/10.17576/
jsm-2018-4705-01
Wan Juliana WA, Damanhuri A, Razali MS, et al (2010) Mangrove flora of Langkawi. In:
Langkawi Research Centre, Institute for Environment and Development (LESTARI)
Wilkie ML, Fortuna S (2003) Status and trends in mangrove area extent worldwide. Rome
Yahya N, Idris I, Rosli NS, Bachok Z (2020) Mangrove-associated bivalves in Southeast Asia: A
review. Reg Stud Mar Sci 38:101382. https://doi.org/10.1016/j.rsma.2020.101382
Zhang K, Liu H, Li Y, et al. (2012) The role of mangroves in attenuating storm surges. Estuar Coast
Shelf Sci 102–103:11–23. https://doi.org/10.1016/j.ecss.2012.02.021
Geospatial Tools for Mapping
and Monitoring Coastal Mangroves 21
L. Gnanappazham, Kumar Arun Prasad, and V. K. Dadhwal

Abstract

Mangroves are biologically important and productive ecosystems endowed with


diverse flora and fauna. Despite their economic and ecological services,
mangroves face threats from anthropogenic as well as coastal hazards. Assess-
ment and monitoring of this critical and vulnerable habitat of the coastal ecosys-
tem could play a major role in implementing conservation and management plan
compatible with Sustainable Development Goals (SDGs). Mangroves by their
geographic confinement in coastal/marshy areas are well suited to study with RS
& GIS tools given the difficulty to conduct extensive field trials. Developments in
the field of remote sensing and Geographic Information System (GIS) in the last
three decades have substantially facilitated smart and efficient use of field surveys
for assessing different parameters of mangrove ecosystem such as mapping,
monitoring the health of mangrove cover, assessing their diversity, characteriza-
tion of their biophysical and biochemical properties, as well as monitoring the
conservation and restoration activities. Recent advancements in sensor
technologies, providing very high spatial resolution multispectral, hyperspectral,
microwave, and LiDAR data have substantially improved characterization and
monitoring of mangroves. Contemporary Data Science methods on storage,
geospatial data analytics, and advanced automated algorithms in handling the
BIG Data available (archive and real-time data) facilitate a better understanding
and assessment of spatio-temporal behaviour of the mangrove ecosystem. This

L. Gnanappazham (*) · V. K. Dadhwal


Indian Institute of Space Science and Technology (IIST), Trivandrum, India
e-mail: [email protected]; [email protected]
K. A. Prasad
Department of Geography, Central University of Tamil Nadu, Thiruvarur, India
e-mail: [email protected]

# The Author(s), under exclusive license to Springer Nature Singapore Pte 475
Ltd. 2021
R. P. Rastogi et al. (eds.), Mangroves: Ecology, Biodiversity and Management,
https://doi.org/10.1007/978-981-16-2494-0_21
476 L. Gnanappazham et al.

manuscript presents an overview of how the RS and GIS technologies being


evolved in the context of their use for scientific and quantitative studies on
mangroves.

Keywords

Geospatial tools · Remote sensing · Image processing · Classification ·


Geographic information system · Spatial modelling · Floristic mapping ·
Biophysical and biochemical · Shoreline changes

21.1 Introduction

Mangroves are salt-tolerant woody halophytes generally thrive in marshy inter-tidal


zones of tropical and sub-tropical coastlines inundated with diurnal (inundated once
a day) or semi-diurnal (inundated twice a day) tides. The mixing of fresh and saline
water as well as muddy and anaerobic soil substratum in the estuarine regions favour
its growth (Tomlinson 1994).

21.1.1 Distribution of Mangrove Ecosystem

Mangroves are distributed along the low energy, sedimentary tropical coastlines
largely between 30 N and 30 S latitudes of over 123 nations covering an area of
around 0.15 million sq.km (FAO 2007; FSI 2019), with most of its proportion seen
in South and South East Asian countries. Mangroves could survive in extreme
habitat conditions such as high temperature, strong winds, turbulent inundation,
high salinity, muddy anaerobic substratum. This is possible because they develop
some special physiological and morphological adaptations such as saline water
regulating root cell membranes, prop roots, aerial roots, and viviparous seedlings
(Lugo and Snedaker 1974). The mangrove forest structure is generally characterized
by zones of tree species depending on the inundation frequency, tidal range, resultant
salinity levels, microtopography, and coastal morphology (Woodroffe 1992;
Dalrymple et al. 1992; Ellison 2018). For instance, the world’s largest mangrove
ecosystem, Sundarbans is located in the extensive deltaic region shared by India and
Bangladesh. The region is characterized by flat terrain formed from the sedimenta-
tion of an intricate system of rivers such as the Ganges, Brahmaputra, Meghna, and
many other tidal channels (Nazrul-Islam 1993).
Recently, Global Mangrove Watch (GMW) Initiative estimated the mangrove
extent for the year 2010 as 137,600 sq. km., of which Asia holds the major share with
38.7% succeeded by Latin America and the Caribbean (20.3%), Africa (20%),
Oceania (11.9%), North America (8.4%), and the European Overseas Territories
(0.7%) (Bunting et al. 2018). Of the global total, three-fourth of the mangrove cover
is seen along the coastlines of just 15 nations, mostly in Southeast and South Asia.
The eastern countries were found to have more species diversity with 58 species
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 477

Arctic Circle

Tropic of Cancer Atlantic Pacific


Ocean Ocean

Pacific
Equator Ocean
Indian
Ocean
Tropic of Capricorn

Antarctic Circle

Atlantic Ocean
Number of Mangrove species

Longitudinal range
Western Mangroves Eastern Mangroves

Fig. 21.1 Global distribution and diversity of mangroves. (Map Source: Tang et al. 2018 and Chart
data from Tomlinson 1994)

under 23 genera when compared to the western countries with only 12 species under
7 genera (Spadling et al. 2010) (Fig. 21.1).
As per the India State of Forest Report 2019, the total mangrove forest cover of
the country is 4,975 sq. km. which accounts for 0.70% of the nation’s forest cover.
This includes a very dense forest of 1,476 sq. km. (29.66%), moderately dense forest
of 1,479 sq. km. (29.73%), and open forest of 2,020 sq. km. (40.61%). Among states
and UT’s, West Bengal holds the major share (42.45%) of mangrove area followed
by Gujarat (23.66%), Andaman and Nicobar Islands (12.39%), Andhra Pradesh
(8.12%), Maharashtra (6.44%), Odisha (5.04%), and other states (Fig. 21.2).
Based on the geomorphologic characters, mangrove habitat in India can be classified
into three categories: East coast, West coast, and Island mangroves. Major
mangroves occur in the east coast with 57% of the total area and 88% of Indian
species count when compared to the west coast. This is mainly because the east
flowing rivers form extensive deltas with a dense network of creeks and muddy
levees favouring the growth of mangroves (Kathiresan 2018; Mitra 2020).
478 L. Gnanappazham et al.

Fig. 21.2 Distribution of mangroves along Indian coastline. (Map prepared from data source: FSI
2019)

21.1.2 Ecological Function

Mangrove is complex yet one of the most productive ecosystems in the world due to
its survival capability through self-adaptation, habitat hosting, and diverse range of
flora present in the community. (i) The ecosystem acts as a breeding and feeding
ground for a large variety of fishes, mollusks, crustaceans, and other related fauna.
Therefore, mangroves are credited for its autotrophic nature which helps to maintain
the coastal food chain (Alongi 2002). (ii) Most recently, the efficient carbon
sequestration capability of mangroves attracts global attention and often regarded
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 479

as “Carbon Sink” with the annual carbon sequestration potential of two to four times
higher than that of the terrestrial tropical forests, which is highly significant in the
perspective towards climate change mitigation and adaptation measures (Murdiyarso
et al. 2015). (iii) Another significant characteristic of mangroves is that they act as a
natural defence against coastal hazards like cyclones, storms, tsunami, and coastal
erosion. (iv) Mangroves can withstand the sea level rise through vertical accretion
under suitable conditions (Menéndez et al. 2020). Besides these, mangroves are
considered to have medicinal value for acute diseases and be a resource of household
material, fuel and fodder for the coastal community (Mitra 2020).

21.1.3 Threats on Mangroves

Despite its ecological, biological, and economic relevance, mangroves face threats
from natural and anthropogenic factors. During the period between 1980 and 2005,
approximately 35,600 sq. km of mangrove area had been lost which is much higher
than the global forest cover loss (FAO 2007). Between the years 2000 and 2012, the
global annual average mangrove deforestation rate was ranging between 0.16% and
0.39%, while it was much higher (between 3.58% and 8.08%) in Southeast Asian
and South Asian countries. The loss was mainly due to the conversion of mangrove
area into aquaculture ponds, palm plantations, rice cultivation, timber logging, and
urban development (Sahu et al. 2015; Hamilton and Casey 2016; Jayanthi et al.
2018; Friess et al. 2019) which also reduced the species richness (Duke et al. 2007).
The intense deforestation also causes severe impacts on carbon stored in mangrove
soil and sizable below ground pools of dead roots. It is estimated that the loss of
carbon storage due to mangrove deforestation would be ranging between 90 and
970 Million Ton/year, which is much higher than their annual carbon sequestration
rate (Alongi 2014).
There exist indirect threats in the form of reduction in upland fresh water
available to the mangroves by the construction of water storage reservoirs and
associated with polluted water carried away by the discharge from upland agriculture
practices with the usage of fertilizers and insecticides.

21.1.4 Conservation and Management

Several international conservation policies and global rehabilitation targets framed


for the conservation of mangroves became successful to a certain extent. For
example, the intergovernmental treaty “Ramsar Convention of Wetlands” intends
to conserve global wetlands, including mangroves and their resources through strict
regulatory policies, participatory management, and international cooperations
(Kuenzer et al. 2011). Furthermore, “Blue Carbon Initiative” project funded by the
United Nations Environmental Programme (UNEP) aims to restore and protect the
vegetated coastal habitats to improve carbon sequestration through wetlands
(Nellemann et al. 2009; Ellenbogen 2012).
480 L. Gnanappazham et al.

Conservation and management of mangrove are highly relevant with several of


the United Nations Sustainable Development Goals (SDGs) including Goal 13: Take
urgent action to combat climate change and its impact; Goal 14: Conserve and
sustainably use the oceans, seas, and marine resources for sustainable development;
and Goal 15: Sustainably manage forests, combat desertification, halt and reverse
land degradation, halt biodiversity loss (Chow 2018). However, it is known that the
no-net-loss restoration of mangroves at global level requires mass restoration
strategies such as site selection, species selection, artificial regeneration, etc.,
through the involvement of local stakeholders and institutions (Bosire et al. 2008).
Mapping the spatial distribution and species composition of mangrove ecosystem is
a prerequisite to generate baseline information for planning effective management
strategies.
Remote sensing has replaced traditional field survey methods by providing cost-
effective, continuous data for assessing different parameters of wetlands such as
their spatial distribution, species identification, health status, invasive species iden-
tification, water quality, biomass characterization, anthropogenic impacts as well as
monitoring and management of conservation and restoration activities (Kuenzer
et al. 2014). Moreover, the availability of archived and real-time acquisition data
facilitates the evaluation of seasonal and long-term changes in the coastal wetland
ecosystem and its causative factors.

21.1.5 Information Needs for Monitoring Mangrove Ecosystems

The information that plays a crucial role in the decision making process and devising
efficient conservation and management practices of mangrove ecosystem are listed
below. And eventually, most of them could be derived through remote sensing and
GIS tools.

a. Spatial Extent and the dynamics of mangroves and associated land cover features.
b. Interrelations with changing coastal landscapes.
c. Threats on mangroves in terms of either degradation or conversion due to
anthropogenic and natural causes.
d. Species Distribution/dominance and its temporal variation.
e. Mangrove phenology (associations and species variation).
f. Mangrove biophysical characterization (height, biomass, leaf areas, etc.).
g. Mangrove Functional Attributes (Net Primary Productivity, Net Carbon
Exchange, Evapotranspiration, etc.).
h. Mangrove Ecosystem Services (e.g., protection from the tsunami).

This chapter intends to give a brief overview covering the fundamentals of remote
sensing, Geographic Information System (GIS) and their potential and demonstrated
applications in assessing and monitoring the mangroves, as well as for devising
efficient conservation and management strategies.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 481

21.2 Concepts of Geoinformatics

The current field of Geoinformatics or Geospatial Science and technology is a


synthesis of concepts from traditional and/or individually developed fields such as
Geography, Cartography, aerial photogrammetry, remote sensing, database, com-
puter science, global positioning system, navigation, etc. (Fig. 21.3).
Practically there are a few basic types of information products expected from the
application of geoinformatics:

• Map of a particular theme or set of themes. For example, thematic maps like
Mangrove distribution, vegetation map with mangrove types/association/species.
• Resultant map(s) by the combination of more than one theme of our interests. For
example, mangrove density classes falling under different age groups or the range
of soil salinity in which Avicennia species are dominant.
• Resultant statistics/tabulations of the attributes and associated attributes of the
above two.
• With the advantage of developments in the contemporary computing and hard-
ware techniques, visualization of above outputs with best possible 2D and 3D
capabilities.
• Visualization on mobile devices leading to Virtual Reality maps.

Fig. 21.3 Evolution of the field of geoinformatics or geospatial technology or simply remote
sensing and GIS
482 L. Gnanappazham et al.

21.2.1 Basics of Maps

A map is defined as the abstract of a real-world scenario and the degree of abstraction
depends on the

• Reduction: Intended application and area of interest (mainly the type and the size
of the region) will decide the scale of the map (we will discuss the scale later in
this section).
• Simplification: Level of position and their attributes to be included in the map
with the use of technologies and instruments.

However, one should find the optimum reduction and simplification requirements of
the application.
Cartography, a branch of geography, is defined as the art and science of making and
study of maps in all aspects. All maps can be characterized by two basic elements:
location of the feature/object in two-/three-dimensional space (x, y, z) and its
attributes, the qualities or magnitudes of mapped feature, for example, type of land
or man-made structures like road, buildings, etc., in mapped location (Fig. 21.4).
Depending on the objective, only the features to be represented are (i) filtered out for
data collection and further processing, and (ii) transformed/reduced to the scale.

21.2.1.1 Scale
While representing the earth’s feature on the map plane, the dimension of the feature
on map is necessarily smaller than the mapped area on the ground. Hence, the
amount of reduction from the measurements on real world to the map needs to be
established using a linear metric and this could be achieved by map scale. Theoreti-
cally, scale is the ratio between the distance represented in map to the actual distance
on the reality/ground and the scale is unit less. Map scale can be represented in three
ways namely,

Fig. 21.4 Survey of India topographic maps of Pichavaram mangroves: Left: 1:250,000 scale map
surveyed in 1935 showing over view of the wetland and Right: 1:50,000 scale map surveyed in
1970 showing intricacies of wetland including drainage details
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 483

a. Representative Fraction (RF) is represented as the ratio of the map distance to the
ground distance in which both distances are represented in the same unit. For
example, if the scale is 1:5000, then one unit in the map is equal to 5000 units in
ground, i.e., 1 cm length in the map is 5000 cm (50 m) of length in reality or 1 mm
in the map is equal to 5000 mm (5 m) in reality.
b. Statement Scale or verbal notation of expressing the scale as a simple statement
for representing the number of map units as a fraction corresponding to earth
units. The scale 1:50000 may be defined 1 cm to 500 m.
c. Graphical Scale is the method of expressing the scale by bar graph calibrated to
express RF or statement scale visually. Representing scale as Bar is ideal to avoid
any error occurring due to the enlargement or reduction of the map. However, just
enlarging a small scale map to a bigger size does not mean that the resultant map
gives all the positional details of an equivalent large scale map. Figure 21.4 (Left)
shows a small scale map enlarged to match with the large scale map given in the
right of the same area and extent. However, the details are more on the right-side
map than the left-side map.

With this one can understand that the resizing of the map does not affect the
physical measurements, but the details given in the map are specific to a representa-
tive fraction or statement scale.
In general, the map scale indicates what details a reader could expect from
the map and to estimate distance. The selection of map scale, primarily depends
on the size of the area to be mapped, the extent of details to be shown and the size of
the paper. Based on the cartographic scale, maps can be categorized into:

a. Large scale map—where specific features of the landscape in a small area will be
shown on a map and it provides detailed information. For example, cadastral
maps, town plans, etc. try to show every detail of the small area selected.
b. Small scale maps—where more earth area is covered in each map unit, and hence
the details are generally generalized or limited in these maps. For example, wall
maps and maps in school atlas are small scale maps.

Currently with the use of GIS technology, maps can be displayed at various scale
(zoom) levels and thus scale and maps have become digital which also facilitates
their viewing on web.

21.2.1.2 Coordinate System and Map Projections


A place or location is identified on the earth surface by either two- or three-
dimensional coordinates. It may be either (i) geographical or spherical coordinates
in terms of latitude, longitude, and altitude or (ii) Cartesian or planar coordinate
system using any of projection methods (http://geokov.com/education/map-
projection.aspx).
Projection is a mathematical concept of transferring the locations of the features
on the 3-dimensional global surface to a 2-dimensional map plane surface. The
transformation from the surface of the earth (near ellipsoid) to the flat surface always
484 L. Gnanappazham et al.

involves distortion and no map projection is perfect (Orange peel cannot be flattened
perfectly even after tearing). Every map projection tries to preserve one of the spatial
properties of earth features and the neighbourhood relation such as true direction,
true distance, true area, and true shape. A map projection is called conformal when
the shape of the features is preserved. Equal-area (Equivalent) projection tries to
represent the true area of the earth features on the map. Equidistant projections
preserve the true distance between points in a particular direction, while True-
directional or Azimuthal projection represents the true direction from the centre or
particular point of the mapped area.
There is a long history behind the evolution of projections (Snyder and Maling
1993). As far as India is concerned, till 2000 the Survey of India and other agencies
were using Polyconic projection. The reason could have been that the widely used
reference map prepared by the Survey of India till 2005 used Polyconic projection
aiming equivalent area upon projection (Later SOI starts releasing digital Open
Series Maps in UTM projection). Similar norms were prevailing in other countries
also with varied types of projection, which are suitable to represent the respective
region. However, Universal Transverse Mercator (UTM) projection is being univer-
sally accepted by remote sensing and GIS community because of less error upon
projection.

21.2.1.3 Surveying
Technically, surveying is the measurement of the location of a feature or object in an
area or on the earth surface in terms of Cartesian coordinates or projection
coordinates of a location having known coordinate (Bench Mark or Control
point). Traditional survey methods include measuring using chain, compass or
Theodoloite while modern methods include the use of instruments such as Total
Station, Global Positioning System (GPS), and advanced GPS instruments that help
the surveyors to make measurements at sub-cm accuracy. As such remote sensing
technology is also called as one of the methods of surveying.

21.2.1.4 Global Positioning System and Geo-location


In recent decades, Global Positioning System has substantially reduced the laborious
process of conventional survey methods and analytics. The user of a GPS receiver/
instrument can get his three-dimensional position on the earth’s coordinate system
(latitude, longitude, and altitude) and the time in a precise manner with the help of
signals received from a constellation of satellites. Hence, it is been useful for
navigational purpose in the land, air as well as on sea. The positional accuracy of
such GPS varies from a few centimetres to metres depending on the specifications of
instruments used and the availability of open space to receive the satellite signal.
However, Assisted GPS (AGPS) which is installed in smart mobile phones help
us to precisely get our location using GPS first and increase its accuracy, using the
positions of connected cell phone towers thus enabling us to get the location even
inside our home or under the trees (as long as the connectivity exists). Such AGPS
gives us the traffic information, finding the nearest facility, etc. through wireless
network.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 485

21.3 Remote Sensing: An Introduction

In earlier times, conventional field survey methods were adopted for mapping and
recording forest characteristics which is time consuming, labour intensive, expen-
sive, and practically difficult in the case of temporal data collection in the inhospita-
ble environment of the mangrove ecosystem. Remote sensing can be defined as “the
science and art of obtaining information about an object, area, or phenomenon
through the analysis of data acquired by a device that is not in contact with the
object, area, or phenomenon under investigation” (Lillesand et al. 2004) (Fig. 21.5).
Evelyn L. Pruitt of the U.S. Office of Naval Research has first introduced the term
“remote sensing” in the 1950s, to denote the technology which is observing the
earth’s landscape from the distance without any physical contact.

21.3.1 Electromagnetic Radiation

The understanding of the electromagnetic radiation (EMR) characteristics and its


interaction with the physical environment is much needed before looking into the
details as EMR is the principal source of remote sensing. Light travels at a speed of
2.99  108 m/sec and attains the form of electromagnetic radiation having an electric
and a magnetic field that are right angle to each other and perpendicular to the
direction of travel (Rees 2013).
EMR is specified in terms of either wavelength (μm or nm) or frequency hertz
(Hz). The relationship between wavelength and frequency of electromagnetic radia-
tion could be derived as c ¼ ν λ, where c is the speed of light. Based on the
wavelength, different regions of the EMR spectrum, including Gamma rays,
X-rays, Ultraviolet, Visible, Infrared, microwaves, and radio waves have been
recognized (Fig. 21.6).

• Optical spectrum includes


i. Visible spectrum (0.30–0.7μm) comprising of UV (0.3–0.4μm) and visible
light (0.4–0.7μm) which is a narrow waveband yet very important in photo-
grammetry and remote sensing and
ii. Reflective Infrared region (0.7–3.0μm) including Near-Infrared (0.7–1.3μm)
and Mid-Infrared (1.3–3.0μm) are used along with the visible region in optical
remote sensing.
• The thermal Infrared region (3–5μm and 8–14μm) involves the measurement of
emitted long wave radiation from the earth’s surface. Understanding the black
body radiation concept and associated laws will give more insight on this
(Lillesand et al. 2004; Jensen 2014).
• The Microwave region (1 mm–1 m) of the electromagnetic spectrum is used in
passive and active modes. The longer wavelength of microwave can penetrate
through clouds and thick atmosphere hence acquire images in all weather
conditions (Awange and Kyalo Kiema 2013).
486

Ground
Source of light Satellite/ receiver
(Electro Magnetic Sensor
Radiation – EMR)

Interaction with Atmosphere


(Absorption and Scattering)

Remote sensing
Interpretation/
Analysis
Interactions with earth’s features
(Reflectance, transmittance and
Absorption)

Various Applications
‐Landuse/ Land cover
‐ Geology
‐ Water resources
‐ Vegetation
‐ Urban

Fig. 21.5 Components of remote sensing process


L. Gnanappazham et al.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 487

Radio Waves

Visible
Gamma Rays X- Rays Ultraviolet Infrared Microwave FM AM Long Radio Waves

10-16 10-14 10-12 10-10 10-8 10-6 10-4 10-2 100 102 104 106 108

wavelength (λ) in metres

400 nm 500 nm 600 nm 700 nm

wavelength (λ)

Fig. 21.6 Regions in the electromagnetic spectrum categorized based on their wavelength. The
region between 400 nm and 700 nm is the only region visible to human eye. (Source: Jensen 2014)

21.3.2 Interaction of EMR

While EMR enters from the vacuum space to the atmosphere it interacts with the gas
molecules and particles in the atmosphere including thin clouds and rain drops it gets
redirected from its original path called Scattering. Absorption occurs when certain
gases in the atmosphere absorbs certain wavelengths of EMR blocking them to reach
the earth surface. The wavelength region that could pass through the atmosphere
without much attenuation is called atmospheric windows and these regions could be
effectively used in remote sensing. When the incoming radiation hits a target on the
earth surface, it either gets reflected, transmitted or absorbed by the target based on
the properties of the target and the incident wavelength (Jensen 2014; Lillesand et al.
2004).

21.3.3 Spectral Signatures

The remote sensing sensors measure the reflected energy of the incident solar
radiation from various objects of the earth’s surface depending on the texture,
physical and chemical properties of the incident target, and the prevailing atmo-
sphere. Different objects reflect and absorb differently at different wavelengths, and
when we plot this we get a continuum curve with crests and troughs implying the
reflectance peaks and absorption troughs. This is called a Spectral Signature, which
is unique to an object like a fingerprint to every individual (Fig. 21.7). This premise
provides the basis for multispectral remote sensing (Lillesand et al. 2004).

21.3.4 Evolution of Remote Sensing

In 1863, Maxwell proposed the electromagnetic theory followed by some of the


significant findings in understanding the electromagnetic radiation beyond the visi-
ble spectrum. The first aerial photograph was reportedly made by Gaspard-Felix
Tournachan (also called Nadar) from a balloon tethered over the Bievre Valley,
488 L. Gnanappazham et al.

Fig. 21.7 Spectral signatures of selected natural and man-made features for the wavelength range
of 0.4μm–2.5μm. (Data Source: USGS Spectral Library Version 7)

France in 1820. By the time of the Second World War in the 1940’s, significant
developments in the field of photographic reconnaissance were observed including
the development of infrared and radar sensors (Rees 2013).
Zhou (2001) characterized the development of remote sensing satellites broadly
into four stages: The first generation space borne (1960–1972) satellites acquired
images primarily for reconnaissance survey (CORONA, ARGON, and LAN-
YARD). Second generation satellites are successful for various applications using
multispectral images (series of Landsat, SPOT, and IRS). Third generation satellites
have been launched with high end sensor technology to acquire very high resolution
images (SPOT HRV, IRS L4) and all weather capability with active sensors like
ERS-1, JERS-1, and RISAT-2. The development in very high spatial (sub-metre)
and spectral resolution (hundreds of bands) sensors denotes the fourth generation
with the additional possibility of multiangle and three-dimensional observations
with in-built GPS. The next generation technology expects to focus on an integrated
system for a real-time earth observation requiring sophisticated imaging sensors,
storage, communication hardware, and automated algorithms (Fu et al. 2020).
In the context of ecological and environmental mapping and monitoring the
following categorization of sensors may be useful for a beginner to start working
with the satellite data.

• Coarse resolution optical (VIS-NIR-SWIR) sensors having high repetivity (more


frequent acquisition) are more suitable for environmental monitoring of large
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 489

patches (e.g., NOAA-AVHRR, MODIS, MERIS, NPP having > 250 m


resolution)
• Moderate resolution optical multispectral sensors are generally used for man-
grove separation from other vegetation as well as discriminating associated
vegetation and land cover features (~25 to 250 m). Most of such data is free
and open, thus very popular among users.
• High (~5–25) and very high resolution (0.25 to 5 m) used for discrimination,
mapping inventory, and change detection. High resolution sensors can provide
very accurate information, but have a low swath which increases the gap between
revisits to the same region. However, a new class of small satellites in large
constellation (e.g., Planet Labs) is capable of providing daily visits.
• Active sensors (Sect. 21.3.4.1.1) SAR in C, X, and L band obtain texture,
dielectric and overcome cloud and thick atmosphere which makes it operable in
all weather conditions.
• Hyperspectral sensors with >50 bands and 10 nm spectral resolution are available
for large area mapping as archive though not available for monitoring as they are
not operational. Future missions like EnMAP (DLR, Germany) may overcome
this limitation.

The main properties of the sensors widely used for earth observation are summarized
in Annexure 1.

21.3.4.1 Types of Remote Sensing


Remote Sensing can be categorized into different types based on different criteria
such as the source of EMR, platforms on which the sensor is being mounted, and its
orbit.1 Low Earth Orbit, Medium Earth, Geostationary, Sun Synchronous are the
types of RS based on orbit.

Based on the Source of EMR


Based on the source of energy used by the sensor, remote sensing can be of two
types. Sensors recording the reflected or emitted EMR from the object which is
irradiated from artificially generated energy sources are Active sensors and from a
natural source (Sun) are Passive sensors similar to the camera used with and without
flash light. Most of the earth observation multispectral sensors are of Passive type
and LiDAR, RADAR fall in Active type

Based on the Platform


Platforms are structures or vehicles on which remote sensing instruments are being
mounted providing varied scale range.

a. Ground based platforms include portable hydraulic platforms and masts, and
non-portable towers and weather surveillance stations.

1
Orbit is the circular or elliptical path of the satellite above the earth for imaging.
490 L. Gnanappazham et al.

b. Airborne platforms include hot air balloons, Unmanned Air Vehicles (UAV),
drones, and aircrafts.
c. Spaceborne platforms included rockets, space shuttles, and satellites on which the
sensors are mounted. These sensors provide global and periodical coverage of the
earth’s surface.

21.3.5 Resolutions in Remote Sensing Data

The data collected by the sensor can be characterized in terms of its resolution and
that can be of four types: spatial, radiometric, spectral, and temporal resolution.

21.3.5.1 Spatial Resolution


The spatial resolution, analogous to map scale, referring to the size of the smallest
possible object that can be identified in the image (Picture Element or Pixels) and it
depends on the Instantaneous Field of View2 (IFOV) of the sensor/camera and
height of the satellite from the ground surface. Usually, geostationary weather
satellites provide data with very coarse spatial resolution in a kilometre scale while
polar orbiting earth observation satellites often provide finer resolution in metres.
Examples for such satellites include Landsat 7 ETM+ (30 m), IRS LISS-III (23.5 m),
Resourcesat (5.6 m), etc. Due to the recent advancements in sensor and imaging
technology, remote sensing data with very high spatial resolution less than 5 m are
also available but at a higher cost.

21.3.5.2 Radiometric Resolution


It refers to the sensitivity of a sensor to the differences in signal strength while
recording the radiant flux reflected, emitted, or back-scattered off the target. Radio-
metric resolution is measured in bits to refer to the number of brightness levels
(black–grey–white) used to record the reflected energy or capture the image. The
higher the bit value, the more details can be obtained from the recorded image. For
example, Landsat 7 ETM+ image has the radiometric resolution of 8 bits which
means objects within the image can be differentiated using 28 ¼ 256 brightness/grey
levels (0 to 255) depending on the reflectance of the objects at the particular
wavelength.

21.3.5.3 Spectral Resolution


The spectral resolution refers to the ability of the sensor to acquire the image using
more number of bands at finer bandwidth. Based on the number of spectral bands
used and the spectral interval (band width) for which a sensor is sensitive, remote
sensing images are categorized into,

2
IFOV is the angular cone of sensor’s/ camera’s visibility to earth features (encompassing
pixel size).
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 491

a. Panchromatic image is the acquisition of an image in a wider spectral range


generally encompassing visible and or NIR spectrum and usually displayed in
grey scale image.
b. Multispectral image is the most commonly used multi-band image with a mini-
mum number of bands (usually less than 10), comparatively broader bandwidth,
with higher spatial resolution.
c. Hyperspectral image is also a multi-band image having more than hundreds of
spectral bands with relatively smaller bandwidth to record continuous spectrum.
Analysing hyperspectral data requires a solid understanding of the spectral
signatures of the objects under study.

21.3.5.4 Temporal Resolution


The temporal resolution of a remote sensing system refers to the frequency of image
acquisition of the same region, i.e., repetivity of the satellite which is dependent on
the orbital3 parameters of the satellite, swath4 width, and latitude of the region. High
altitude orbit, wider swath, and higher latitudes allow more frequency than low
altitude, narrow swath, and lower latitudes. The analysis of multi-date data can be
used to study the temporal behaviour of the object or phenomenon. Geostationary
satellites have a continuous view of the target (weather satellites), while in case of
polar orbiting satellites for earth observation, satellites are required to have high
temporal resolution for natural resource monitoring for example, Landsat 7 ETM+
sensor provides data with the temporal resolution of 16 days.
Some satellites are designed to acquire multiple images of the same area in a
shorter interval of time either in the same orbit (Cartosat-2 series) or on the
subsequent orbits (SPOT, IRS series) using a steering mechanism which is known
as the revisit capability of the satellite. Current days with the entry of private sectors
in space technology, a constellation of small satellites launched can accommodate
the revisit capability by acquiring images on required time (Planet Labs series of
satellites).

21.3.6 Multispectral Remote Sensing

Advancement in space technology and frequent availability of medium resolution


data from satellites like Landsat, SPOT, IRS, ASTER, etc. (Annexure 1) and
developments in computer algorithms for image preprocessing and classification
have made the regional mapping of earth resources including mangroves an easy job
and helped in monitoring and management activities in the last three decades.

3
Orbital parameters include the altitude, inclination, and duration of satellite orbit around the earth.
4
Swath is the width of the ground imaged by the satellite during the orbital pass.
492 L. Gnanappazham et al.

21.3.6.1 Thematic Mapping


Mapping is the process of converting the image to information readable by a
common user. The image collected by a remote sensor would be used either in
analog or digital formats depending on the user’s preference to map using visual
interpretation or digital mapping.

Visual Interpretation
Image interpretation could be defined as the examination of images to identify the
objects and judging them with the inherent skill of the human eye-brain system
(McGlone 2004) by applying multi-concepts like multispectral, multi-temporal,
multi-scale, multi-source, etc. Usually, visual interpretation is performed on the
digitally pre-processed satellite data to map or derive the best results out of that. In
general, satellite data is viewed either as a black and white image if a single band or
True Colour Composite (TCC) or False Colour Composite (FCC) for multispectral
bands while the latter is very common among ecologists and RS experts. Visual
interpretation constitutes the fundamental principles of the object identification or
discrimination (Jensen 2014; Konecny 2014) and the interpretation elements include
(Please refer Fig. 21.8).

a. Location: It gives the precise location information of the image through x, y


coordinates, and this could be obtained through traditional survey methods or
GPS receivers.
b. Colour: The amount of energy reflected at different wavelength regions varies for
different materials on the earth’s surface. In Fig. 21.8, the mangrove vegetation
appears red in colour because the vegetation reflects more NIR radiation, while
sand and other concrete features (buildings) appear as white as it reflects almost
equally in all bands. The water reflects comparatively more in blue and green
bands than red and NIR bands and so appears blue.
c. Tone: the variation in the same colour depending on the amount of energy
reflected in a particular wavelength. That is within vegetation itself, depending
on its types the tone of red colour varies from bright red to dull red similarly from
bright blue to dull blue for water bodies.
d. Size: The size of an object in the image is scale dependent. The determination of
relative size of the target to that of nearby objects aids in the identification of the
target.
e. Shape: It represents the structure or outline of the features and is an important
element while determining any distinctive feature. Shape may be regular, for
example, in the case of man-made features like aquaculture ponds and canals, and
irregular in case of natural rivers and mangrove vegetation.
f. Texture: It indicates the characteristic arrangement or repetitions of colour in an
image, and is scale dependent. Figure 21.8 represents heterogeneous vegetation
cover that could be identified by the difference in its texture.
g. Pattern: It is the spatial arrangement of objects in the area, either randomly (sparse
vegetation) or systematically arranged (aquaculture ponds).
21

Size and Shape: Regular


S shape in Blue shows man
Color: Red shows the made water bodies
vegetation, Blue is water (Aquaculture(AQ), Tank
and white is sand or any (T) & Canal(CL)) while
building/ concrete feature Tank can be
differentiated from
Aquaculture ponds with
its smaller dimension
Tone: Different shades/ tone CL
of red color shows the Pattern: Regular rectangular
difference in vegetation pattern in blue shows the
type A: Agriculture, M: J presence of series of
Mangroves and C: Casuarina B aquaculture ponds
(AQ)while similar pattern in
Red shows the agricultural
area (A).

A
Association: Rectangular
Texture: Smoothness in the white pattern shows the
color. The settlement (S) M C
AQ bunds (B) of Aquaculture
seems to be coarse in ponds where the linear
texture. M (mangroves) white structure in side the
T
Geospatial Tools for Mapping and Monitoring Coastal Mangroves

seems to be smoother than river is Jetty (J). Similarly


C (Casuarina) showing the regular water body
different type of associated with agriculture
vegetation. is Tank (T) while aside of
wetlands is aquaculture
ponds (AQ)

Fig. 21.8 Elements of visual image interpretation as explained through IKONOS (MS + PAN) 1 m resolution image in FCC (Red: Band 4; Green: Band 3;
Blue: Band 2) of Pichavaram, Tamil Nadu India
493
494 L. Gnanappazham et al.

h. Shadow: Remote sensing images are acquired mostly in time when the sun is in
near nadir position (between 10 AM and 2 PM local time) to avoid shadows.
Shadows may hinder the interpretation, but at specific instances, they can give
distinctive clues on the target object.
i. Association: Association refers to the circumstance when a certain feature or
object along with other related features or activities.

Minimum Mapping Unit


When satellite remote sensing data is used either by visual interpretation method of
digital classification discussed below, a minimum size of object or feature on the
earth which can be mapped depends on the spatial resolution of the sensor. For a map
being prepared by visual interpretation, an object of 3  3 pixel size can be mapped
as a polygon and for a digital classification group of 3  3 or 2  2 pixels is the
minimum size of a feature to be classified which is called Minimum Mapping Unit
(MMU). MMU is at least twice the size of resolution of the data, i.e., if the resolution
of satellite data is 30 m, then MMU will be at least 3x3 times 30  30 ¼ 90  90 m
and thus minimum size of an object of 90  90 m can be mapped and pixels below
this size are not mapped due to inherent error in the data.

21.3.6.2 Digital Image Analysis


While using visual interpretation technique, spectral characteristics beyond visible
region of EMR spectrum could not be fully utilized because of the limited ability of
the eye to distinguish the tonal difference and associated spectral reflectance.
Therefore, over the years, scientists developed advanced algorithms and methods
to process the satellite remote sensing data in digital image format for information
extraction with little to no user intervention make it time and cost efficient. Generally
visual and digital image interpretation techniques could complement each other and
the combination approach is widely accepted as the best for remote sensing applica-
tion. A detailed description of many of the digital image analysis methods could be
found in literature in recent times (Lillesand et al. 2004; Reddy 2008; Jensen 2016).
During satellite data acquisition, the radiated (reflected) light energy from the
earth surface features received by the sensor is stored as Digital Number (DN) of
pixels (picture element) of the imaged area which is a function of spatial scale,
spectral band used and radiometric resolution (grey scale). Digital analysis involves
the computation of DN value of the satellite image by considering it as a matrix
(rows  column) of pixels.

Image Preprocessing
Images collected through space borne sensors are often attributed with radiometric
and geometric errors and hence corrected/pre-processed to convert them to a usable
format before visual interpretation or further digital analysis. In fact, it can be
compared with the correction techniques available in any handheld or mobile
cameras which are similar but substantially at a lower level. Preprocessing could
be categorized into the following three broad operations.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 495

i. Radiometric Correction: Radiometric correction involves the removal or reduc-


tion of noise introduced into the data due to the electronic noise of the sensor
system and attenuation due to atmospheric conditions. Radiometric corrections
could be done through simple image normalization or advanced radiometric
calibration to radiance or scaled surface reflectance.
ii. Geometric Correction: The geometric errors may arise due to the platform
instability resulting in skewing or distortions. Geometric correction of an
image is important to precisely locate the point of interest in the geographic or
projected coordinate system (Sect. 21.2.1.2), to perform overlay analysis with
other thematic and temporal maps of the same area and to match with the
adjacent images.
iii. Atmospheric Correction: Topographic attenuation and atmospheric conditions in
the target area highly influence the image quality. However, the correction of
atmospheric error is not mandatory for all the applications unless the remotely
sensed is used for comparing the output maps/parameters with that of a different
area at different time scales.

Though the above preprocessing techniques are to be applied on raw satellite data,
when it is received by the user, these corrections are already carried out by the
respective space agency unless the raw data is needed by a researcher to work with.
Hence, the space agency provides the data from raw level to pre-processed level
depending upon the user’s requirement. It is advisable to use pre-processed data for
any ecological applications (Sect. 21.3.10).

Image Classification
When the image is corrected for errors, the image is directly used for thematic
mapping using classification methods or improved by enhancement, filtering, and
transformation algorithms based on the theme of mapping (Please refer to Lillesand
et al. 2004; Jensen 2016 for more details). Image classification involves the identifi-
cation of spatial and spectral patterns in the remote sensing data corresponding to
different land use/land cover and presents the results as thematic information
through the analytical process called pattern recognition. Several image classifica-
tion algorithms have been developed based on parametric, non-parametric, and
non-metric methods. Parametric methods are used to analyse ratio and interval
type of data (Sect. 21.4.4) while Non-Parametric methods could also be used to
analyse the nominal data type. The classification algorithms can be broadly classified
into unsupervised and supervised classification methods based on the underlying
principle of user intervention in training sample selection.
In unsupervised classification, the image pixels are grouped into a specified
number of classes based on the spectral reflectance of resultant classes. In this
case, the analysts need only to specify the number of classes and the algorithm itself
clusters the pixels based on a statistically determined measure on the pixels’
DN/radiance values. The analyst could then reiterate the number of classes if the
result is found unsatisfactory and repeat the classification until it clusters into
496 L. Gnanappazham et al.

informative classes. ISODATA and K-means are two popularly used unsupervised
classification methods.
In Supervised classification, the analyst would be able to select training sites for
each of the class within the image to train the classification algorithm. The analyst
could utilize elements of visual interpretation and field knowledge to select such
training samples representing known classes of interest. The analyst should collect
at least N + 1 training pixels for each of the class when N numbers of bands are used
in multispectral or hyperspectral image analysis to derive reliable output of
supervised classification. This is called Hughes Phenomenon (Richards and Jia
2005). So to obtain satisfactory results it is advisable to identify the maximum
number of training samples throughout the image on which the RS analyst is
confident. Classification results can be represented as thematic maps, tables, and
digital data files. Following are some prominent supervised classification algorithms
prevalently used in digital image analysis.

i. Minimum Distance Classifier (MDC) calculates the mean of each training class
in spectral space5 and then it measures the spectral distance between each of the
pixels in the image to that of the mean of each training class. Euclidean distance
is the most common distance measure used in this method. Then, the input pixel
will be assigned to the respective spectral class for which the measured distance
is minimum (Fig. 21.9a).
ii. Maximum Likelihood Classification (MLC) works based on the assumption that
the statistics for each class in each band are normally distributed. The classifier
calculates the probability that a given pixel belongs to a specific class and each
pixel is assigned to the class that has the highest probability, i.e. maximum
likelihood (Fig. 21.9b).
iii. Spectral Angle Mapper (SAM) calculates the similarity of the spectrum of
unknown pixel to the spectrum of the reference pixel in the spectral space of
dimensions corresponding to the number of spectral bands used. It calculates the
angular difference between the reference and target pixels in vector space and is
usually smaller the angle closer to the classes (Fig. 21.9c).
iv. Support Vector Machine (SVM) is a non-parametric supervised classification
method derived from statistical learning theory suitable for complex and noisy
data. It employs optimization algorithms to locate optimal boundaries with the
least errors among all possible boundaries separating classes (Fig. 21.9d linear
SVM). When the training samples are not linearly separable, the samples are
mapped using “non-linear SVM” (Schölkopf and Smola 2001; Tso and Mather
2009).

Accuracy Assessment
Regardless of visual or digital mapping, the classification output needs to be
assessed for its accuracy to allow a degree of confidence to be attached to the result.

5
Spectral/feature space is n dimensional space where the DN/spectral values of n bands are plotted.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 497

Fig. 21.9 Graphical representation of (a) Maximum Likelihood Classifier, (b) Minimum Distance
Classifier, (c) Spectral Angle Mapper, and (d) Support Vector Machine supervised classification
algorithms. (Image Source: Lillesand et al. 2004; Richards and Jia 2005)

This could be done through the preparation of classification error matrix or confu-
sion matrix. Confusion matrix (Table 21.1) compares the relationship between
known reference data (ground truth) and corresponding classification results on
class-by-class basis. Commission or inclusion error indicates the incorrect assign-
ment of class (A) to pixels that belong to classes other than A and the corresponding
accuracy is called User’s accuracy (in user’s perspective how many pixels of other
classes are included in a particular class) while omission or exclusion error occurs
when the pixels belonging to a particular class (A) are missed to get assigned to that
class (A) and the corresponding accuracy is Producer’s accuracy (in map maker’s
perspective how many pixels are omitted from a class).
498 L. Gnanappazham et al.

Table 21.1 Confusion matrix to assess the accuracy of classification for a total of 95 pixels/
samples of 3 classes: Mangrove, agriculture, and water
Reference Mangrove Agriculture Water Classified Total User accuracy
Classified
Mangrove 21 6 0 27 21/27 ¼ 0.78
Agriculture 5 31 1 37 31/37 ¼ 0.84
Water 7 2 22 31 22/31 ¼ 0.71
Reference 33 39 23 95
total
Producer 21/ 31/ 22/ OA ¼ 74/
Accuracy 33 ¼ 0.64 39 ¼ 0.8 23 ¼ 0.96 95 ¼ 0.78
Kappa ¼ 0.67

N  ΣCii  ΣðCTi  RTiÞ ΣCii


Kappa ¼ ; Overall accuracy ðOAÞ ¼
1  ΣðCTi  RTiÞ N

Where N—total number of pixels, Cii—number of correctly classified pixels/


samples, CTi—total number of the classified pixels for ith class, and RTi—total
number of reference pixels for the ith class.
The overall accuracy is computed by dividing the total number of pixels correctly
classified by the total number of reference pixels. Kappa coefficient calculation, a
discrete multivariate technique used in accuracy assessment, is a measure of the
difference between the actual agreement between reference data and an automated
classifier and the chance agreement between the reference data and a random
classifier (Lillesand et al. 2004).

21.3.6.3 Vegetation Indices (VI)


Regardless of classification or mapping, the information about the vegetation traits
could be studied using Vegetation indices (VI) which are used by application
scientist from the basic understanding of vegetation to multiple analyses, which
uses the spectral reflectance characteristics of different vegetation groups in different
spectral bands. VI indicates the vegetation phenology associated with the health
condition, stress and biophysical characteristics and is an important parameter
widely used in studying agricultural and vegetation monitoring from local to global
scale. Scientists have developed generally applicable, vegetation specific, and spec-
tral band specific vegetation indices for effective monitoring of global vegetation
(Richards and Jia 2005; Jensen 2016). The most widely used VI is Normalized
Difference Vegetation Index (NDVI) which is derived using Red and Near-Infrared
bands of multispectral data (Townshend and Justice 1986; Tucker and Sellers 1986).
The values of NDVI varied between 1 and +1. Higher NDVI values represent the
higher vegetation vigour and health (Fig. 21.10). Specific to wetlands, Normalized
Difference Wetland Vegetation Index (NDWVI) was derived (Kumar et al. 2019) by
making use of the sensitivity of wetlands to shortwave infrared which helps in
mapping the mangroves very accurately.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 499

Fig. 21.10 NDVI image derived from NIR and Red band of Pichavaram mangroves

21.3.7 High Resolution Imaging and analysis

Recent advances like high resolution multispectral data such as IKONOS, Quick
Bird, and Worldview and recent classification algorithms such as Object-Based
Image Analysis (OBIA) have refined the classification accuracy of mangroves
further. OBIA, unlike pixel-based classification, the images are first segmented
into “objects” representing groups of pixels belonging to ground patches, entities
or their elements (primitives) which can be then classified by unsupervised,
supervised, or rule-based algorithms. Some of the important properties of OBIA
method over pixel-based classification methods include (1) ability to include object-
level shape, texture, and relevant contextual properties of an image in addition to its
spectral reflectance values into the classification framework, (2) ability to reduce the
salt-and-pepper noise and enhancement of classification accuracy through smooth-
ening of local variation within objects, and (3) ability to analyse with multiple object
layers nested within each other at different spatial scales to identify landscape patch,
cover type, and ecosystem structure at multiple hierarchies. This estimation of
ground objects through image objects finds them more ecologically relevant and
potentially more resilient to minor geospatial positioning subject to image registra-
tion error (Dronova 2015).

21.3.8 Hyperspectral Remote Sensing

Hyperspectral remote sensing is otherwise called as imaging spectrometry as it


combines two sensing modalities, imaging and spectrometry. Normally the spectral
wavelength range of 400–2500 nm is used in imaging spectrometry technique which
typically acquires images in more spectral bands in a narrow bandwidth approxi-
mately ranging from 0.5 to 25 nm. This difference in reflectance properties among
500 L. Gnanappazham et al.

materials has made hyperspectral remote sensing, a potential technique for


identifying the intricacies in micro-level in earth observation.
The reflectance acquired in narrow contiguous bands of hyperspectral data (which
is lacking in multispectral) provide the analyst with very peculiar information in
relation to the leaf biochemical and biophysical characters which could be utilized to
characterize the mangrove ecosystem such as species discrimination, plant health
monitoring, nutrient intake characterization, invasive species monitoring, etc.
(Fig. 21.11). Differences in spectral reflectance among different plant species
provided by the contiguous bands can be used for (a) species level classification,
(b) biochemical (chlorophyll and carotenoid content estimation), and (c) biophysical
characterization (Leaf Area Index, Biomass, etc.). For instance, we take the example
of spectral signatures of eight mangrove species from the same family
Rhizophoraceae at multispectral bandwidth of Landsat 8 OLI sensor (left) and
hyperspectral bandwidth of ASD Fieldspec 3 radiometer (right) (Fig. 21.11).
Though hyperspectral data has an enormous amount of spectral information, its
analysis has some technical difficulties in deriving useful information out of such
voluminous data. Apart from this, hyperspectral data also possesses multi-
collinearity problem due to the redundant information from more number of
bands. Appropriate dimensionality reduction methods and advanced classification
algorithms must be chosen for developing an efficient framework for information
extraction (More details on HS image processing methods can be referred in
Richards and Jia 2005).

21.3.9 Microwave Remote Sensing

Mangroves and flooded vegetation usually have distinct microwave signatures when
compared to associate land use land cover. In microwave remote sensing, the signal
received is measured for its intensity and is termed as backscatter coefficient in
decibels (dB) unit. Differences in wavelength and polarization6 of transmitted and
received signals, and incidence angle on the vegetated surface can exhibit varying
backscatter coefficients based on the different transmission capability of microwaves
under various configurations. Also, internal properties of plant such as moisture
content, cell structure, biochemical content, etc. and biophysical properties such as
size, geometry, leaf, and branch orientation results in unique backscatter signal
value) (Kuenzer et al. 2011). The most important property of microwave data is
that it is not prone to cloud cover, haze, and other atmospheric disturbances and this
property makes it suitable for mapping mangroves as they locate in tropical and
sub-tropical areas. Microwave data were used for mangrove cover mapping, bio-
physical parameters retrieval, health status monitoring, and biomass estimation

6
Electromagnetic waves consist of an electric and a magnetic field vibrating at right angles to each
other and it is necessary to adopt a convention to determine the polarisation of the signal. For this
purpose, the plane of the electric field is used as plane of Polarization.
21
Geospatial Tools for Mapping and Monitoring Coastal Mangroves

Fig. 21.11 Spectral reflectance of eight mangrove species from Rhizophoraceae family at the spectral resolution of Landsat 8 OLI multispectral sensor (left)
and ASD Fieldspec 3 hyperspectral sensor (right). (Source: Prasad and Gnanappazham 2016)
501
502 L. Gnanappazham et al.

21.3.10 Evolution of RS Data Access

Remote sensing satellite data has been used for wide spread applications in India,
after the launch of first ever EO satellite IRS 1A launched in 1988 and National
Remote Sensing Centre (NRSC), Indian Space Research Organization (ISRO) was
identified as the authorized distribution agency of Indian and foreign satellite data
and the same continues. Since then the satellite data are accessed in the form of either
hard copy or in digital mode. Images in hard copy form were supplied either to match
with the Survey of India topographic maps (1; 250,000 or 1:50000) or in terms of
scenes7 wise images. Such images are mostly used in False Colour Composite
(FCC—Please refer Fig. 21.8) format. Digital mode of data was supplied in digital
media such as tapes/floppies/CD/DVDs. In both cases, the user was able to get either
pre-processed data (corrected for geometric and radiometric errors) or raw data.
Later with the development of technology and the huge amount of archive on
earth observation data, Analysis Ready Data (ARD) was made available by the data
providers including USGS, BHUVAN that pave the way to the access of Data Cube
(spatio-temporal data with space on 2 dimensions and time as the third dimension
(Fig. 21.12). Such data set are not limited to just remote sensing data but also,
derived maps such as landuse/land cover, water resources, infrastructure, hazard
maps, etc. Development of data cube involves data preprocessing, integration and
optimization of data and its storage and make it ready for sharing and analysis (Kopp
et al. 2019) (See also 5 Open Geospatial data).
The Open Geospatial Consortium (OGC), which was established in 1994,
brought together different academic, private and public sector parties and soon
focused on the standardization of interfaces for accessing geospatial data resources
and analytical procedures (Reichardt 2017). It enabled web based services for data
search, access, analysis, and output generation with or without downloading the
original data (Guo and Onstein 2020)

21.4 Geographic Information System

Theoretically, GIS could be defined as the organized system of hardware, software,


data, and people to enter the data, edit it, analyse or manipulate and model the data to
derive useful output in terms of maps, statistics or charts. Initially, GIS had been
widely used along with remote sensing technology mainly for the earth resource
observation, assessment, and management. However, the application of GIS has
diversified growth in different sectors from administrative management to stock
market and business sector. In this section, we will see the basics of different
components of GIS.

7
Scenes are regular division of continuous image acquired for the specified swath along the track of
the satellite’s orbit.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 503

Fig. 21.12 Data cube: Representation of spatio—temporal data showing the dynamics in
mangroves and coastal geomorphology of part of Krishna delta

21.4.1 GIS/Spatial Data Model and Data

There are two types of spatial data models, namely raster and vector depending on
the type of data handled. Raster is the data in image format, i.e., regular arrangement
of pixels in continuous rows and columns which are usually used to represent
continuous data or gradually varying data (mostly numeric data) and vector data
possess the data in the form of points and collection of points as lines or polygons
which are called discrete data (Fig. 21.13) which are of categories or nominal data.

21.4.1.1 Selection of Data Type


The choice of GIS data types either raster or vector varies across the various stages of
GIS project discussed earlier such as the data collection, entry, editing, analysis, and
output generation. For example, secondary data source of wetland map can be either
in raster form prepared using digital classification or in vector form interpreted
visually from Remote Sensing data (Fig. 21.14). Suppose the collected wetland
504 L. Gnanappazham et al.

Fig. 21.13 Representation of four different land cover features in raster format with 1 m pixel size
(left) and vector format (right) represented by connected points with coordinates

map is in raster form, the GIS analyst may like to convert into vector format for
editing or analysing the data with other maps. Finally the output of the analysed data
may be represented in vector or raster format depending on the requirements. In
general it is advisable to handle the continuous spatial data (e.g., soil or water
quality, rainfall, temperature) in raster format and discrete features/data (e.g., wet-
land classes, administrative boundaries) in vector format as discussed in the begin-
ning of this section.

21.4.1.2 GIS/Spatial Data


In the course of GIS development, many forms of spatial data are used from
secondary source and one should make sure the spatial data is from an authentic
source. For example, Survey of India (SOI) topographical maps8 are used as
reference maps for Reconnaissance or preliminary survey of any ecological study.
SOI maps are mostly available in 1:250,000 (small scale), 1: 50,000, and 1:25,000
(large scale) and the details provided are in vector form. Other maps such as village
map, cadastral map, and tourist maps are also used depending on the need. Data
acquired in raster or image format such as aerial photographs, remote sensing
satellite images are used as primary data source. As we could see from the techno-
logical development, we have authorized/reference maps and time series global
images from different sources in digital plat form (computers and mobile) such as
Google Earth in raster form, Google map provides many features including trans-
portation, land cover features, and infrastructure in vector format, most of the others
like Open street maps, Bing map, Wikimaps, Open layers, etc. provide in vector

8
SOI Maps are series of maps covering the entire country providing details about general geo-
graphic and topographic features with proper projection coordinates and scale along with elevation
contours and necessary Bench Marks and Control points (www.soinakshe.uk.gov.in).
21
Geospatial Tools for Mapping and Monitoring Coastal Mangroves

Fig. 21.14 Godavari mangroves: Mangrove community map prepared in raster format using digital classification of satellite image (left) and mangrove density
505

mapped in vector format using visual interpretation of satellite (right). (Source: Ravishankar et al. 2004)
506 L. Gnanappazham et al.

Table 21.2 General characteristics/Attributes of sample wetland features


Wetland Average Area Average Elevation from
Id class salinity (ppt) pH (sq.m) Suitability Mean Sea Level (m)
A Mangroves 28 7.0 200 High 1.0
B Mudflat 32 6.0 340 Moderate 0.8
C Sand dune 9 6.5 289 Low 2.3
D Water 30 7.3 443 Low 0.0
body

format. All these open data makes the initial step of GIS development of any
ecological survey a hassle-free process.

21.4.2 Attribute Data

Attributes are nothing but the properties and characters of a particular spatial feature
or object represented in GIS. For example, A, B, C, and D (Identifier or symbol)
given in Fig. 21.13 (a) may have an attribute of wetland classes, namely mangroves,
mudflat, sand dune, and water body. Other attributes may include its properties or
characteristics, such as average salinity, pH, area, perimeter of each feature, etc.
Collection of such attributes along with the polygonal feature is called Geospatial or
GIS database (Table 21.2, General characteristics/Attributes of sample wetland
features). Spatial and attribute data can be either a (i) Primary data which is collected
or prepared directly by the GIS developer or (ii) Secondary data which is collected
from other sources as second hand data.

21.4.3 Classification Scheme and Database

As we had seen a few of the attributes of mangrove ecosystem in Sect. 21.4.2, it is


important to know the basis behind categorizing the wetland classes. The map given
in Fig. 21.13 represents four different land cover features in raster format with 1 m
pixel size (left) and vector format (right) represented by connected points having the
coordinates(a) has four wetland classes where each class is exclusive and has its
unique name (Table 21.2) and will have a clear definition which could form a part of
wetland land classification scheme (Please refer to the well-structured hierarchical
classification scheme defined for wetland classes in Sect. 21.4.3). Defining or
adopting such a classification scheme will be the first step in any map preparation
before survey or data collection. Other attributes (except suitability) about the
classes given in Table 21.2 General characteristics/Attributes of sample wetland
features are numeric values which are directly entered in the database. Such type of
numeric data can be classified based on either (i) equal interval of values or
(ii) natural break among values or having (iii) equal number of records depending
the analytical or output requirements. Once, classification scheme is finalized, spatial
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 507

database (collection and compilation of spatial and attribute data in an organized or


structured format) is designed to make the GIS input, editing and analytical pro-
cesses in a systematic manner.

21.4.4 Data Input and Editing

From the above description, GIS data possess two components. (i) Spatial data
which containing unique geographic coordinates or other spatial identifiers and
(ii) Aspatial or Attribute data which explains the properties of the spatial informa-
tion. Spatial data is entered either by on-screen digitization (vector format) of points
(x, y); line (x1:y1, x2:y2, . . . .xn:yn) and polygon (x1:y1, x2:y2,. . . .,xn:yn, x1:y1) or
secondary source images (scanning/photographs/satellite images). The data may be
either 2 dimensional (x, y) or 3 dimensional (x, y, z). Input of data includes necessary
geometric correction as detailed in Remote sensing (Sect. 21.3.6.2.1) to match the
entered data to a preferred coordinate system either geographic or projected (see
Sect. 21.2.1.2).
Attributes can be entered either in tabular form using different data types, namely
Nominal/Characters (Id and Wetland class), Ratio (“0” has no value—Area), Inter-
val (“0” has a value—Elevation) (integer—Area; float—pH), and Ordinal (Suitabil-
ity) as given in Table 21.2. Similar data can also be prepared as text format or in
Excel/Office work sheet separately and then integrated with the GIS Database of
vector data.

21.4.5 Analysis

Handling more than one layer being the major advantage of GIS, one can make a
query or analyse one data or combination of multiple data. For example, an area up
to 500 m from either side of the river (Fig. 21.15b) can be estimated using buffer
analysis. Similarly, the extent of mangroves, mudflat, and other categories falling

Fig. 21.15 (a) Wetland map, (b) river (blue) map with buffer zone of 500 m (light red), (c) overlay
of buffer zone of river on wetland to estimate the area of the wetland feature under the buffer zone
508 L. Gnanappazham et al.

under the buffer zone (Fig. 21.15c) can be extracted using overlay analysis of both A
and B maps as highlighted by red around blue over the wetland features in
Fig. 21.15c. Varied types of analyses can be performed to derive useful information
from the GIS which includes

i. Overlay analysis: To know and estimate the extent of features from one map
which are falling in or out of the features of another map. Similar to the second
example given above (Fig. 21.15c)
ii. Buffer analysis: To find out and measure the area extending for a particular
distance from a point or line or polygon feature as given in the first example
given above (Fig. 21.15b).
iii. Neighbourhood analysis: Extension of the previous analysis. For example, to
know how many soil samples are taken within the buffer zone if samples are
collected evenly throughout the wetland features.
iv. Topographic analysis: We can combine any number of features along with
elevation data to carryout 3D analysis. For example, if elevation profile of the
region (Fig. 21.15a) is available, the extent of mangroves inundated by tidal flux
can be estimated.
v. Interpolation: Different algorithms are used to generate maps of continuous data
(raster) such as elevation, soil salinity, organic carbon, pH, temperature, rainfall,
etc. using data collected from distributed sample locations.
vi. Network analysis: Analysing and optimizing the travelling path and the object
or resource or people or any commodity travelling through a network such as
road, river, canals, electrical, telecommunication network, etc. If the drainage
network of a mangrove environment is available as GIS layer, we will be able to
map the region getting inundated through the drainage canals. Addition of
topographic data will fetch more accurate results.
vii. Visualization: Interactive way of visualizing the map individually or in combi-
nation with other maps in a two-dimensional or three-dimensional
representations. Tremendous development in computer hardware and software
technology paves way to visualize any ecosystem or environment virtually in it
otherwise called as Virtual reality in 3 Dimension.

21.4.6 Three-Dimensional (3D) Data and Analysis

As specified earlier GIS handles spatial data either in 2- or 3-dimensional mode (x, y,
z). Mostly 3D spatial data are represented in raster format and is widely known as
Digital Elevation Model (DEM) which is nothing but the raster image wherein the
pixel values represent the altitude of that pixel location. For example, if the wetland
map/LULC map is displayed along with its topography i.e., the altitudinal is added
with the spatial data we call it as 2.5D data and not a 3D data because wetland data
does not vary with altitude for a particular location (x, y). However, DEM is
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 509

Fig. 21.16 False Colour Composite of Pichavaram mangrove wetland displayed over the
integrated terrestrial elevation and sea/water depth (bathymetry) data derived from GPS and
bathymetric chart (Insert). (Source: Gnanappazham L 2008)

generally accepted as 3D GIS when integrated along with other thematic9 maps and
helps to model the area of our interest. Basically there are three ways to generate
topography (i) Modelling a region using 3D will add more value to the GIS
developed and provides crucial details to solve various decision making problems
where terrain topography/microtopography plays major role (Fig. 21.16)

21.4.7 Spatial Modelling

Geospatial modelling in the context of GIS can be defined as a system to simulate the
real-world scenario for a particular time or over a period of time. It can vary from
simple evaluation to prediction of some feature or phenomenon of interest
(Goodchild 2005).
GIS modelling can be categorized into (i) Simple data modelling is a descriptive
representation of real-world patterns in a database schema, (ii) Process modelling is
the simulation of processes in the real world either static (input and output are of

9
Thematic map is specifically designed to show a particular theme namely land use or water or
geology of a particular region.
510 L. Gnanappazham et al.

User Process of
request

Sends Request Requests data


Web Web server Database
Sends Sends data

Rendering the Post process: graphical


result of query environment of query

Fig. 21.17 Information flow from User—WebServer—User through WebGIS query

same time—for example extent of inundation of mangrove wetland due to tidal flux
in a day) or dynamic models (output time is different from input time, for example,
the future extent of mangroves if the tidal flux is going to reduce over a period of
time) (https://clarklabs.org/terrset/land-change-modeler/), (iii) Space modelling
which is a conceptualization of geographic space in 3 dimension using
DEM/Triangulated Irregular Network/LiDAR.

21.4.8 Web GIS

As on when there is a development in computer engineering and technology, it had


been incorporated simultaneously in GIS field also. Hence, the development of
World Wide Web/Internet in sharing the information across the globe in no waiting
time allowed the field of GIS also to share the geospatial data and the results of
spatial analytics across the stakeholders from different parts of the globe. Today’s
world greatly depends on WebGIS and its application in every hour of human life as
it has become an integral part of mobile. For example, transportation assistance,
routing facility, home delivery of commodities, tracking the status, etc. completely
rely on the position of the user and supplier/delivery person by his/her mobile’s
location.
A standalone GIS is accessible to the person handling the data from the computer
on which he or she is handling. If the same is accessed or analysed through a wired
network within an organization it becomes an “Enterprise GIS”. If the same is
performed through World Wide Web/Internet we call it as “WebGIS”. Simply it is
a combination of integrated or independent GIS data server and web server and the
client get information out of it using a web browser installed in a computer or mobile
(Fig. 21.17). The main focus of WebGIS is the dissemination of spatial data and the
functionalities to the end-user with easy go approach, and its major applications
include in e-governance, public utility services, real-time analysis, and mitigation
measures during disasters and hazards, crowd sourcing (data input from localized
resources) and reporting, etc. In general WebGIS portals can view derived
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 511

information, a few can perform spatial analysis and few else can access the original
data depending on the level of permission assigned. Open data portals which provide
satellite data and its derived products are discussed in Sect. 21.5.

21.4.9 Spatial Data Quality

Quality of the data is the fitness of the data for the intended application. Hence, the
measure of the data quality varies according to the requirements of the application
and is measured in terms of Accuracy, Precision, Error, and Uncertainty. Accuracy is
the degree to which the data matches with reality. Analysing country wide popula-
tion data may need an accuracy of ten thousand scales whereas on district level
would need the actual figure. Similarly the position of a city may be accurate to a
kilometre in a country map but on a district map it must be between 10 and 50 m.
However, there exists a linear relation between accuracy and the cost of data or data
collection. Precision is another measure of geospatial data quality specifying the
level of details and intricacies. For example, the temperature of a place is given as
27.55  C is more precise than 27  C. Error is the inverse of accuracy while
Uncertainty is the degree to which the accuracy of data is not known.
In general the following components are defined to assess the quality of
geospatial data.

1. Positional accuracy (PA): The degree to which the position or coordinate of the
spatial data either a pixel in an image or a point in vector data is matching with
real coordinates on the ground. Hence, the accuracy is measured in terms of the
degree of mismatch between coordinates of data or map point to coordinate on
ground point (such points are called as Ground Control Points—GCPs) also
called as Root Mean Squared Error (RMSE).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2  2  2
xm  xg ym  yg zm  zg
RMSE x ¼ , RMSE y ¼ , RMSE z ¼
N N N

xm, ym, and zm are the coordinates on map and xg, yg, and zg are coordinates
measured on Ground or GCPs. The planimetric accuracy is given
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
by RMSE x 2 þ RMSE y 2 . RMSEz is elevation accuracy.
As we have seen the accuracy depends on the intended application. Hence, the PA
is given by 0.5 mm  scale. For example, PA of 1:50000 scale is 25 m (0.5 mm  1/
50000) and 1:10000 scale is 5 m (0.5 mm  1/10000). As now all geospatial data are
handled in digital mode, with the maximum possible enlargement, 0.5 mm accuracy
threshold is raised to 0.25 mm and accordingly accuracy is estimated.
512 L. Gnanappazham et al.

2. Attribute Accuracy (AA): Geospatial data is encompassed with attributes of the


spatial data represented in raster or vector format. And we had seen that attributes
can be entered in numerical and categorical formats. Accuracy of numerical
attributes is assessed using RMSE similar to positional where instead of position,
the attribute values are compared with ground reality or sample data collected
from field. For example, the accuracy of satellite product showing temperature of
a region can be assessed by collecting representative sample points from the field.
Categorical attributes given in text form can be assessed using error matrix or
confusion matrix as discussed in classification accuracy (Sect. 21.3.6.2.2).
3. Other than these two measures, the spatial data should be verified for the
following major components which are inevitable for the potentiality of any
geospatial data.
a. Metadata: This is nothing but data about the data collection and preparation
process such as (i) survey method, (ii) accuracy of instruments, (iii) projection
and transformation parameters, (iv) time of acquisition.
b. Completeness of the data to certify if the data generated has any miss outs.
c. Consistency of spatial data in its data types, positional accuracy and attribute
accuracy of a map and list of maps stored in GIS. If village is chosen to be a
point data, across the map and spatial database, it has to be point. For PA, scale
of maps used or prepared has to be same and AA should be maintained across
the map(s).
d. Semantic accuracy: This deals with the naming convention of features or
objects. In general, location or region specific names are assigned to the
features. Sometimes according to the mapmaker. For example, one person
could identify the marshy region of coastal wetland as “Mudflat” while other
can name it as “Tidal flat”. Such discrepancies must be avoided across the
geospatial database from an organization level to national or global level.

21.5 Open Geospatial Data and Software Availability

Availability of archived and concurrent satellite and associated spatial data, and
processing methods are quintessential for effective spatio-temporal monitoring of
the earth’s environment. In the last two decades, the trend of making the satellite data
and geospatial data available for the end-user by different governmental and research
agencies is widely seen, and the way in which geospatial data are collected,
processed, analysed, and visualized are changed. Collaboratively contributed,
authoritative and scientific geospatial data are now available for registered
non-commercial users through collaborative effort of global geospatial community.
Open scientific geospatial data tends to follow the FAIR (findable, accessible,
interoperable, and reusable) principles in the collection, development, and dissemi-
nation of spatial and satellite data with global standards as framed by international
standards development organizations such as the International Hydrographic
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 513

Organization (IHO), the International Organization for Standardization (ISO), and


the Open Geospatial Consortium (OGC) (Coetzee et al. 2020).
Open source geospatial software includes diverse code libraries, algorithms,
tools, applications, and platforms developed and made available under Open Source
Initiative (OSI) license. Open Source Geospatial Foundation (OSGeo) software
ecosystem (www.osgeo.org), founded as a not-for-profit organization with the vision
of collaborative development of open geospatial technologies, data, and education
for widespread use. The OSGeo projects acknowledge the principles of inclusive-
ness, fostering, openness, and responsibility in all their projects with an objective of
open source geospatial software for community focusing on earth monitoring.
OSGeo project family includes geospatial libraries (Actinia, GeoTools, Orfeo Tool-
box, GDAL, OWSlib), web mapping GIS (Map Server, GeoWebCache, PyWPS,
GeoServer), spatial database (PostGIS, Open Data Cube), and desktop GIS
(QGIS, gvSIG, GRASS GIS, Marble). Other than OSGeo projects, contemporary
platforms such as R (R spatial tools), Python (GeoPython), Javascript (Leaflet, Map
Box), and Blender for GIS have developed open source geospatial tools (Coetzee
et al. 2020).
Open geospatial data involves the collection of geospatial data and satellite
images, and disseminating the data through modern technologies such as the Inter-
net, the Internet of Things (IoT) through hosting portals. Another kind of open data
follows the principle of sharing the information to user community by the authorities
through dedicated data portals. Since the last decade, the recent trend of Open Spatial
Data Infrastructure (SDI) emerged with the standard of transparency and collabora-
tion with government and research organizations. Google Maps, Wikimapia, and
Open Street Map (OSM) are examples of collaboratively contributed open geospatial
data that follows the principle of citizen science through crowd sourcing—paid or
volunteered, collaboratively maintained, and continuously updated through the
Internet.
Authoritative open geospatial data includes the geospatial vector data such as
administrative boundaries, point data, street lines, and associated attributes collected
and maintained by government agencies are made public to user community with
minimum to no restriction. Satellite data collected have also been made public
through open license policy, for example, Bhuvan Data Discovery and Metadata
Portal of Indian Space Research Organization (ISRO) provides thematic data and
satellite imagery of global coverage. Similar portals include Copernicus: Earth
Observation Program of European Union and USGS Earth Explorer and Earth
Data of NASA. Cloud based platforms such as Google Earth Engine maintain the
catalogue of earth observation satellite imagery, geospatial database of global
coverage at multiple scales, and also JAVA based programming portal
encompassing geospatial analytical tools enabling free access to scientists,
researchers, and developers for a variety of analyses including monitoring the
changes, trends, and quantify Earth’s surface phenomenon at varied scales.
514 L. Gnanappazham et al.

21.6 Applications of Earth Observation Techniques


for Mangrove Ecosystem Monitoring

During the last four decades, monitoring of the inhospitable mangrove environment
using field survey has been replaced by multi-source, multi-platform, multi-scale,
multi-temporal, and multi-resolution remote sensing data. Aerial photographs, mul-
tispectral, hyperspectral, and microwave remote sensing data, LiDAR, and field
spectrometry data are considered feasible, cost-effective, and time-efficient
alternatives for systematic mangrove management. Using these datasets, various
characteristics of mangroves such as spatial distribution, change detection, health
status monitoring, field survey planning, biomass estimation, tree crown delineation,
invasive species identification, anthropogenic impact assessment, ecosystem evalu-
ation, monitoring, and management of conservation and restoration activities could
be studied. In this section, we discuss different aspects of applications of remote
sensing and techniques involved for effective mangrove management.

21.6.1 Mapping and Monitoring of Mangrove Wetlands

Since the development of aerial/satellite remote sensing multispectral data, they have
been widely used for mapping and monitoring any feature of interest on earth. The
following describes the general procedures followed while using multispectral
remote sensing data as a source for an environmental or ecological survey. There
are typically two modes of analysing remote sensing satellite data resulting into
useful maps and statistics, namely (i) visual analysis, which completely relies on the
interpreter’s knowledge on the region and the field of study and (ii) digital analysis
involves series of steps in converting raw satellite data in the form of pixel values to
outputs. However, in practice, the digital analysis requires visual interpretation skills
to validate the result, thus the hybrid of both analyses is used without which the
output will be unrealistic.

21.6.1.1 Wetland Mapping Scheme


Whatever be the method of mapping one adopts, it is very much important to decide
what features are to be mapped before start mapping. Generally a classification
scheme is designed to identify or map any discrete or categorical features. For
example, forests can be categorized into (i) Evergreen, (ii) Deciduous, (iii) Scrub,
(iv) Littoral, (v) swamp, (vi) moist deciduous, etc. based on the vegetation types.
Similarly, there exists different classification schemes for Landuse/Land cover
(LULC),10 wetlands, agriculture, soil, geology, etc. Tables 21.3 and 21.4 show the
classification schemes of LULC and wetlands adopted in India. Visually interpreted

10
Landuse is the nature of use the particular land is put into by human for example agriculture or
industry while the land cover is the natural feature/phenomenon that covers the land for example
water or vegetation or snow.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 515

Table 21.3 LULC 1:50,000 Classification scheme LI—8, LII 3 & LIII—54 classes (NRSC 2017)
Level I Level II Level III
Built Up Urban Built up Built up -Compact (Continuous)
Built up -Sparse (Discontinuous)
Vegetated/Open Area
Rural Rural
Industrial Industrial area
Ash/Cooling Pond/effluent and other waste
Mining/Quarry Mining—Active
Mining—Abandoned
Quarry
Agricultural Cropland Kharif
land Rabi
Zaid
Cropped in 2 seasons
Cropped in >2 seasons
Fallow land Fallow land
Agriculture Plantation Agriculture Plantation
Aquaculture Aquaculture
Forest Evergreen/Semi Dense/Closed
evergreen Open
Deciduous (Dry/Moist/ Dense/Closed
Thorn) Open
Forest Plantation Forest Plantation
Scrub Forest Scrub Forest
Swamp/Mangroves Dense/Closed
Open
Tree Clad Area Dense/Closed
Open
Grass/Grazing Alpine/Sub-Alpine Alpine/Sub-Alpine
Temperate/Sub-Tropical Temperate/Sub-Tropical
Tropical/Desertic Tropical/Desertic
Wetlands Inland Natural (Ox-bow lake, cut-off meander,
waterlogged, etc.)
Man-made (Water logged, saltpans, etc.)
Coastal Lagoon, creeks, mud flats, etc.
Saltpans
Water bodies River Perennial
Non Perennial
Canal/drain Canal/drain
Lake/Ponds Permanent
Seasonal
Reservoir/Tank Permanent
Seasonal
(continued)
516 L. Gnanappazham et al.

Table 21.3 (continued)


Level I Level II Level III
Wastelands Salt Affected Land Salt Affected Land
Gullied/Ravinous land Gullied
Ravinous
Scrub land Dense/closed
Open
Sandy area Desertic
Coastal
Riverine
Barren rocky Barren rocky

maps of Pichavaram mangroves in a large scale using high resolution (IKONOS MS


+ PAN; 1 m) and small scale using coarse resolution (IRS L3; 23 m) satellite data are
given in Fig. 21.18 for reference.

21.6.1.2 Limitations
Even though remote sensing has enabled the user to study the natural environment in
multiple ranges of spatio-temporal scale, there are few decisive factors, including
cloud cover, sensor specifications, the difference in mapping scale, lack of real-time
data, confusion in selecting appropriate datasets, choice of image processing, and
classification techniques involved while mapping mangroves at higher floristic
hierarchical level (Green et al. 1998). Recent developments in remote sensing
technology have made it possible to overcome a few of these obstacles except the
following.

• Since mangroves thrive in the inter-tidal region, the image scene would have been
highly influenced by seasonal and diurnal tides, which makes the radiometric
correction a challenging task.
• Also, the spectral reflectance obtained from the image pixel covering the fringing
and open mangroves would be a mixture of vegetation, soil and water. Selection
of high resolution data or suitable spectral unmixing algorithms would be needed
to overcome misclassification.
• Due to the heterogeneous distribution of mangrove species in many parts of the
ecosystem, it is difficult to obtain a pure pixel for unique species or community
(Kuenzer et al. 2011).

21.6.2 Floristic Discrimination of Mangroves

As mentioned earlier, the distribution of mangroves is highly heterogeneous in a


dense mangrove ecosystem. Conservation and management practices in such man-
grove forests require baseline information of the spatial distribution of species in the
forest. Traditionally the field survey methods, for example, the Point Centred
Quadrant (PCQ) method records the distribution of species and the density of
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 517

Table 21.4 Classification system for mapping mangrove community zones (Nayak and Bahuguna
2001)
Level I Level II Level III Level IV
Onshore Beach Muddy Sandy Fringe tidal mangroves
areas No mangroves
Estuary Inter-tidal mudflat (i) Rhizophora, Sonneratia, Avicennia,
(sandy clay) Bruguiera (pure/mixed communities)
(ii) Saline blanks
Inter-tidal mudflat (i) Kandelia, Avicennia, Rhizophora,
(silty clay) Aegiceras, Sonneratia (pure/mixed
communities)
(ii) Salt marsh vegetation
(iii) Saline blanks
High-tidal mudflat (i) Salt marsh vegetation
(ii) Saline blanks
Transitional areas Grass
Deltaic Seaward margin of A. marina, A. alba, A. officinalis (pure)
complexes inter-tidal mudflats
Inter-tidal mudflats (i) Pure communities of Rhizophora,
Sonneratia, Avicennia, Bruguiera, etc.
(ii) Mixed mangrove
(iii) Salt marsh vegetation
(iv) Saline blanks
High-tidal mudflats (i) Salt marsh vegetation
(ii) Saline blanks
River mouths Sonneratia, Bruguiera, Excoecaria, Heritiera,
Lumnitzera (pure/mixed communities)
Along creeks (i) Sonneratia, Bruguiera, Excoecaria,
(upper end) Heritiera, Lumnitzera (pure/mixed
communities)
(ii) Freshwater species
Transitional areas (i) Saline blanks
(ii) Grassy banks
(iii) Trees/shrubs
Gulf Inter-tidal mudflat (i) Pure/mixed communities of Rhizophora,
region (sandy clay) Sonneratia, Avicennia, Bruguiera, etc.
High-tidal mudflats (i) Kandelia, Avicennia, Rhizophora,
Aegiceras, Sonneratia (pure/mixed
communities)
(ii) Salt marsh vegetation
(iii) Saline blanks
Transitional area Grass
Offshore Islands Sub-tidal area Algae/seaweeds
areas Inter-tidal (i) Mangroves
(ii) Sand vegetation
Coral reefs (i) Algae/seaweeds/seagrass
(ii) Mangroves
518 L. Gnanappazham et al.

Fig. 21.18 Mangrove wetland maps of Pichavaram: Level 4 classification—large scale map
prepared using high resolution data IKONOS MS + PAN data with intricacies in wetland (Left)
and Level 2 classification small scale map using coarse resolution satellite data IRS L3 with less
details (Right). (Source: Gnanappazham and Selvam 2011; Gnanappazham L 2008)

plant distribution by recording the number of trees in different species found at each
quadrant of the ground plot of fixed dimension. This method, similar to the other
field survey method requires more time, labour and financial support, which is often
difficult when planning for a temporal study with short time intervals. At this
instance, archived and real-time multi-source remote sensing could be an asset for
spatio-temporal mapping of species distribution at multiple scales (Annexure 3).

21.6.2.1 Traditional Remote Sensing Methods


At the times of development of aerial photography, they were used for the mangrove
discrimination using visual interpretation techniques. Secondary information from
topographic sheets, field surveys, and colour infrared video imagery were also used
for discrimination of mangroves from other associated land use/land cover. These
aerial imageries were collected as grey scale image (panchromatic) or colour image
with a limited number of bands often in low spatial and temporal resolutions. The
visual interpretation of these products abled the user only to discriminate mangrove
from other land uses and to assess damaged or loss of mangroves rather than
vegetation community discrimination (Fig. 21.19).

21.6.2.2 Image Processing of Multispectral, Hyperspectral


and Microwave Data
Considering mangroves, scientists used multispectral images for community level
discrimination using different approaches and vegetation indices are the foremost
among them. Several vegetation indices have been developed using different spec-
tral bands of which a few of them, for example NDVI and NDWVI, are correlated
with the plant canopy closure and health status of the vegetation, therefore popularly
used along with other indices such as NDWI, SAVI, and SR (Jensen et al. 1991;
Gupta et al. 2018). Next to VI, classification methods are widely used to map the
extent. For example, unsupervised classification algorithms use only the spectral
reflectance recorded in the image pixels, while the other category, supervised
classification algorithms requires the visual interpretation skill of the analyst to
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 519

Fig. 21.19 Community level map of Bhitarkanika National Park, Odisha, India derived from the
supervised classification of IRS P6 LISS-III data

extract samples to train the classifier for image classification (Figs. 21.8 and 21.9). In
recent times, with the development of high spatial resolution optical imaging
sensors, earth observation data at sub-metre level become available. Object-Based
Image Analysis (OBIA) on such high resolution satellite and airborne data
(Quickbird, Geoeye, IKONOS, Worldview-2, Cartosat, etc. (Please refer Annexure
1 for sensor specifications) have refined the classification accuracy of mangroves
further. OBIA, unlike pixel-based classification, the images are first segmented into
“objects” representing groups of pixels belonging ground patches, entities or their
elements (primitives) which can be then classified into categories of interest by
unsupervised, supervised or rule-based algorithms (Fig. 21.20). Thakur et al. (2020)
made a systematic review on the application of multispectral remote sensing satellite
data for the floristic discrimination and change detection of mangrove ecosystem.
At present, the availability of temporally dense, global coverage imaging spec-
trometry data is limited. There are a few space borne hyperspectral sensors that
provide data at a moderate spatial resolution. Few airborne hyperspectral sensors are
currently operational which could be flown at a lower altitude and provides data at a
higher spatial resolution, but lacks global coverage. Specifications of prominent and
globally used hyper spectral sensors are given in Annexure 1.
520 L. Gnanappazham et al.

Fig. 21.20 False Colour Composite of mangroves of Krishna Delta (A); B and C show two types
of segmented images. Vegetation density classes of mangroves using Pixel and Object-based
classification are shown in D and E

Ground truth data is vital information for any type of satellite remote sensing for
its accuracy. While categorical and qualitative ground truth data is used for multi-
spectral data analysis, the accuracy levels of hyperspectral image analysis are
increased by in-situ spectral data collected using spectroradiometer for the discrimi-
nation of features having a similar response to the light energy (spectral
characteristics) such as minerals, vegetation types, soil, and water quality
parameters. Field spectrometry is the technique which is used to quantify the
radiance, irradiance, transmission/reflectance from various earth surface features in
field condition (Jackson et al. 1980; ASD 2001). It is being actively used in the last
two decades in forestry and vegetation sciences for species identification, classifica-
tion, health status monitoring, nutrient intake estimation, invasive species monitor-
ing etc. Several site-specific spectral libraries were collected so far for various
species, including, non-native species (Underwood 2003; Aneece and Epstein
2017), wetland species (Zomer et al. 2009), Mediterranean species (Manakos et al.
2010), shrubland species (Jiménez and Díaz-Delgado 2015), coral reefs (Kutser et al.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 521

2006), agricultural crops (Datt et al. 2003; Rao et al. 2007) and Indian mangrove
species (Prasad et al. 2015).

21.6.3 Biophysical Characterization of Mangroves

Tropical forests play a critical role in ensuring the stability of the climate change
through their efficiency in capturing greenhouse gases (including atmospheric car-
bon) and sequestrate in their biomass (Fig. 21.21).

(i). Biomass is one of the structural properties of the vegetation and its estimation
is one of the essential baseline information needed for climate change mitiga-
tion programme. The change in biomass is an indicator of the stress in the
vegetation induced by natural and anthropogenic disturbances. This leads to
the reduction of its carbon sequestration capability and so it needs to be
monitored temporally (Klemas 2013). The accurate estimation of the above
ground biomass and carbon stored in the vegetation is one of the important
objectives in the resolutions of the Bali Action Plan (2007) approved by the
United Nations Framework Convention on Climate Change (UNFCCC) and
the Kyoto Protocol (2005). The traditional field survey method was exten-
sively used for collecting forest biophysical parameters which is usually time
consuming, cost and labour intensive. Field measurement of above ground
biomass (AGB) is obtained through the destructive sampling of all the plants
within the ground plots of fixed dimension, for example 5 m  5 m. The
collected plant materials are then segregated into different components
(leaves, stem, branches, roots, etc.) and weighed separately to determine
their wet weight. These components are then dried and weighed to obtain
respective dry weight to compute biomass in grams per square metre.
(ii). Plant LAI is the measure of canopy foliage content which estimates the
total area of one side of photosynthetic leaf tissue per unit area of ground
surface (m2/m2). LAI is an important parameter to study land-surface interac-
tion processes, plant productivity, plant-atmosphere interaction, and
parameterizations in climate models. Considering LAI measurements, hemi-
spherical photographs and quantum sensors are being used for field measure-
ment. LAI could be estimated indirectly from the remotely sensed images
through image derivatives and indices. LAI estimated from optical remote
sensing data could serve as a key parameter in calculating above ground
biomass of vegetation (Bréda 2008; Zheng and Moskal 2009). Furthermore,
field survey for biophysical parameter collection in mangrove forest is practi-
cally much difficult because of its hostile and swampy environment. Remote
sensing is found to be a cost-effective and potential alternative for spatio-
temporal monitoring of mangrove biophysical parameters. Multi-source,
multi-sensor, multi-resolution, multi-scale, archived and real-time remote
sensing data from airborne and EO satellites provides a strong database for
the retrieval of biophysical properties such as biomass and Leaf Area Index
522

Fig. 21.21 Carbon flux pathways in mangrove ecosystem


L. Gnanappazham et al.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 523

(LAI), biochemical characteristics of mangroves and other wetland vegetation.


Henceforth, the estimation of mangrove biomass by combining field survey
and remote sensing is recommended for accurate biomass estimation
(Heumann 2011).
(iii). Remote sensing based biomass estimation usually involves field data collec-
tion about biophysical properties such as diameter at breast height (DBH), tree
height, number of individuals of a particular species, leaf area index (LAI), etc.
The application of remote sensing data for biomass and productivity estima-
tion in different ecosystems has been reviewed in detail (Kale et al. 2002;
Klemas 2013; Lu et al. 2014). They mentioned that the unique characteristic of
plants is displayed by its reflectance in the red and infrared regions of
electromagnetic radiation and have a strong relationship with the biophysical
parameters of plants. Empirical and statistical models and advanced data
analytics of high resolution, multispectral, hyperspectral imagers, LIDAR
and microwave sensors using allometric equations provide plethora of data
for biophysical characterization of natural vegetation.
(iv). High resolution multispectral images are often used for the modelling of AGB
and LAI for mangrove forests at a large scale. Vegetation indices such as
NDVI, a simple ratio of multispectral bands, and specific image transforma-
tion methods such as Fourier-based Textural Ordination (FOTO), Grey Level
Co-occurrence Matrices (GLCM), Artificial Neural Network (ANN) methods
supported with in-situ plot biomass measurements have been proven to be
successful in modelling AGB and LAI of natural vegetation (Annexure 4) VIs
derived from WorldView-2 images are more efficient than other bands in
predicting the biomass of high-density mangrove forests. However, using
hyperspectral remote sensing data, the vegetation indices derived out of this
data are found to be outperforming multispectral vegetation indices (Marshall
and Thenkabail 2015). Additionally, mangrove species distribution mapped
using hyperspectral image could be utilized to improve the biomass assess-
ment (Marshall and Thenkabail 2015; Anand et al. 2020).
(v). Recent studies have proven that active remote sensing techniques such as SAR
and LiDAR sensing have distinct advantages over optical remote sensing for
mangrove biophysical parameters retrieval. The time independent and weather
independent operational capability of active SAR sensors is suitable for
sensing of tropical coasts including mangroves where cloud condition is
prevalent. Moreover, active SAR sensing has the potential to record canopy
structure and biomass (Sect. 21.6.3) while passive microwave remote sensing
data can be used to indirectly infer these characteristics from their spectral
response. From recent studies, it is understood that the biomass dependence of
radar backscatter varies based on radar wavelength, polarization and incidence
angle. Comparatively, the longer wavelength is found to be more sensitive
524 L. Gnanappazham et al.

than shorter wavelengths, and cross polarization11 (HV) is more sensitive than
like polarizations11 HH and VV towards plant biophysical parameters. The
reason is that the sensitivity of radar backscatter intensity to variations of
biomass saturates after a certain level of biomass is reached, and the saturation
level is higher for longer wavelengths. Availability of global coverage open
source high resolution SAR data opens a new opportunity to estimate the
biophysical structure parameters of mangrove ecosystems (Mougin et al.
1999; Fatoyinbo and Armstrong 2010; Pham et al. 2019). Furthermore, the
advent of high performance, parallel and cloud computing resources facilitates
the GIS based geospatial analysis for the global estimation of biomass and
carbon sequestration of mangrove ecosystems in an efficient and cost-effective
manner (Tang et al. 2018).

21.6.4 Characterizing Foliar Biochemistry of Mangroves

Assessing the biochemical and biophysical properties of wetland vegetation is vital


to monitor plant primary productivity, carbon cycling, and nutrient allocation within
the ecosystem (Mutanga and Skidmore 2004; Adam et al. 2009). The varying levels
of the biochemical constituents of the vegetation such as carbon, hydrogen, oxygen,
nitrogen, magnesium, etc. in the form of plant cells, tissues and pigments regulate the
photosynthesis and consequently the plant primary productivity(Muñoz-Huerta et al.
2013; Al-Naimi et al. 2016). Some vegetation biochemical constituents are detect-
able at canopy level through spectral reflectance. Visible region of the spectrum
(400–700 nm) is highly sensitive to leaf pigments such as chlorophyll, while
reflectance at infrared region (700–2500 nm) is sensitive to non-pigment constituents
such as proteins (Kokaly et al. 2009). However, spectral analysis for biochemical
constituents of fresh leaves at the canopy scale would be a complex process as the
absorption features of leaf water obscures the absorption features of the biochemical
constituents (Kokaly and Clark 1999). Moreover, the supplementary effects of
vegetation canopy structure, solar illumination effects, atmospheric influence,
signal-to-noise ratio, and the reflectance anomalies from understorey vegetation,
soil, roots, and branches must be considered while using remote sensing data for
biochemical characterization (Asner and Martin 2008). RS based biochemical stud-
ies on Indian mangroves are very limited (Annexure 4) and have great scope further
research.

11
Cross Polarization in which Transmitted and received signals are of different polarization (HV or
VH) and in Like Polarization they are of same polarization (HH or VV).
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 525

21.7 GIS in Mangrove Monitoring and Management

Since the inception of remote sensing technology, GIS became an integral compo-
nent of remote sensing. Later, satellite remote sensing data and their products
became the source for GIS inputs and analysis. In this section, we will see the
application of different components of GIS in mangrove mapping, monitoring and
its potential applications in mangrove management in the context of Indian
mangroves.

21.7.1 Developing GIS of Mangrove Ecosystem

Developing a versatile GIS database will be very beneficial for the effective man-
agement of mangrove ecosystem. GIS should be developed so that, the system will
help to identify the source of the problem, find the optimum solution leading to an
efficient decision making process (Fig. 21.22).

21.7.1.1 Define the Study Region


Identify the boundary of the region, including mangroves and associated land cover
features. It may be either reserved forest boundary or Revenue village boundary
encompassing the complete mangroves under consideration from the coast line. This
will help 70% of the decision making process to finalize the scale of a GIS database
to be developed. In general, the preferred scale for mangrove management is
1:50,000. However, depending upon the intended activities, the larger scales from
1:25,000 to 1:5000 are adopted.

21.7.1.2 Shortlisting the Required Layers/Maps


Independent of the objectives of the study, a critical map is the base map of the study
area which could be from either topographic maps by Survey of India (SOI) or
village map from revenue village office which are authenticated source of base map
details. Such details include High tide/Low tide line, reserved forest boundary,
village boundary (if needed), nearby settlements, water bodies such as creeks,
lagoons, rivers, streams and man-made canals, any reference points or benchmarks,
etc. along with their essential attributes. More than one base map layers can be
included in GIS depending upon the requirements.
Layers other than base map details generated are decided based on the objective
of the study. For example, following list may be needed to include in geospatial data
for Mangroves.

(i) Mangrove wetland map for the period(s) of study considered. So, depending on
the requirements, one or more wetland maps, mostly derived using satellite
remote sensing data are included.
(ii) Vegetation health status: NDVI would also become a supplementary layer of
GIS for the selected period(s) of study.
(iii) Corresponding satellite images for regular reference.
526

Fig. 21.22 Representation of GIS database of Pichavaram Mangrove ecosystem. (Image Source: Gnanappazham 2008)
L. Gnanappazham et al.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 527

(iv) Species composition map (derived using satellite data and verified using
ground truth samples).
(v) Locations of sample data on mangroves, its diversity, health, environment
including soil, water, and atmosphere, etc. collected from the field as per the
requirement of the study.
(vi) Digital Elevation Model of the study region: Mangroves lie along the coast
within an elevation of approximately 10 m above mean sea level depending on
the tidal amplitude of the particular region. Hence, mapping the
microtopography in the range of 0–10 m can be achieved by very high resolu-
tion stereo images or LiDAR from spaceborne or airborne sensors. Field
surveys, including DGPS, are of limited use in a marshy environment where
mangroves are very dense.

21.7.1.3 GIS Input


Categorize the GIS input data types as (i) Primary or Secondary, (ii) point or line or
polygon or raster data, (iii) If raster input, then is it thematic or continuous or satellite
image. Accordingly the spatial data are entered in GIS software along with attributes.
Spatial and attribute data need to be prepared in the format as per the specification of
the software through the process of editing and quality checking.

21.7.1.4 Analysis
There are varied types of spatial analysis carried out for mangrove related studies
including mangrove change detection, monitoring, shoreline changes, time series
analysis, and hydrological and environmental modelling.

Change Detection and Time Series Analysis


Application of satellite data for mangrove studies triggered by Blasco et al. (1986)
for assessing the extent of Pichavaram mangroves and followed by many studies to
map and monitor the mangroves of Indian coast using Landsat and IRS series of
satellite data. Change detection involves overlay analysis of mangrove wetland maps
of two time period under consideration and deriving the extent of (i) unchanged
categories, (ii) change that had happened, especially the extent of mangrove vegeta-
tion converted to other land cover types, and (iii) from other land cover features to
mangroves. Numerous change detection studies had been conducted for most of the
mangroves of India and Fig. 21.23 depicts one such change detection map of
mangroves of Pichavaram (Tamil Nadu). GIS analysis becomes an integral compo-
nent of RS based analysis and relevant literature are included in detail (Annexure 2)
marked witha.

Shoreline Changes
Generally, the difference in positions between High Tide Line (HTL) or High Water
Line (HWL) or Wet/dry line (otherwise called as shoreline proxies) is mapped
between two timestamps to estimate the shoreline changes. However, there is no
standard method available using remote sensing data due to the variability between
tidal range and satellite data acquisition and inequalities in data resolution. ArcGIS, a
528 L. Gnanappazham et al.

Fig. 21.23 Change in mangroves of Pichavaram between 1994 and 2002

popular commercial software has an extension developed by Unites States Geologi-


cal Survey (USGS) called “Digital Shoreline Analysis System (DSAS)” that helps
the user to estimate the rate of change of erosion and accretion using automated
transects along the multi-temporal shorelines provided by the user (https://www.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 529

usgs.gov/centers/whcmsc/science/digital-shoreline-analysis-system-dsas?qt-sci
ence_center_objects¼0#qt-science_center_objects). For Indian coasts, Integrated
Coastal and Marine Area Management (ICMAM) Programme developed the shore-
line proxies depending on the varied and complex shoreline such as Wet/dry line for
sandy coast, vegetative boundary along shoreline, seaward boundary of any artificial
structures along the coast and the shore edge of the cliff’s base. Depending on the
nature of mangroves along the coast, the erosion and accretion process has either a
huge impact (Fig. 21.24; Ramasubramanian et al. 2006) or no impact (Das 2020) on
the mangroves and their extent (Annexure 5).

21.7.2 Environmental Parameters of Mangrove Ecosystem

Overlay analysis of more than one map helps in the decision making process and
increase the efficiency of mangrove management activities as well. Researchers have
worked on assessing the health of mangroves either deriving vegetation indices,
environmental (soil and hydrology), and climatic parameters from time series multi-
source satellite data or by integrating them with field samples on vegetation,
environmental, and climate parameters and the details on literature are explained
in Annexure 6. Doyle and Girod (1997) developed a high end simulation model at
the landscape level, SELVA and at stand level, MANGRO to predict changes in the
mangroves for unit ha and for individual trees for sea level rise and inundation levels.
However, hydrological and environmental modelling of mangroves of India is very
less. Modelling the impact of driving parameters on mangroves and ecosystem and
vice versa, including modelling the wave dynamics has great scope in the context of
Indian mangroves (Amma and Bhaskaran 2020). Nevertheless, the integration of
environmental, hydrological, and climatological data as a Mangrove spatial database
is recognized as the most crucial source of information for mangrove management,
which is taken by the authorized organizations which are discussed in Sects. 21.7.3
and 21.7.4.

21.7.3 Mangrove Conservation and Coastal Regulation Zone (CRZ)


Mapping

Indian mangroves are categorized into Reserved/Protected forests (RF/PF), National


Parks (NP), Community reserves, Wild Life Sanctuary (WLS), and also as Marine
and Coastal Protected Areas (MCPA) based on their ecological importance declared
by Indian Forest Conservation Act, 1980 & the Wildlife (Protection) Act, 1972
(DasGupta and Shaw 2013). MCPA is further categorized into four, namely Cate-
gory: I (C:I) as complete area under inter-tidal zones of NPs and WLS; C:II as a
marine ecosystem of the islands; C:III A and B as sandy beaches beyond seawater
with occasional interaction with evergreen forests (MoEF 2008). CRZ had the initial
set up to protect the 500 m from the high tide line (HTL) from development activities
when implemented in 1981 and later in 1991 the CRZ notification was implemented
530 L. Gnanappazham et al.

Fig. 21.24 Shoreline change along Kakinada Bay endowed by mangroves from 1935 to 2001
(Source: Ramasubramanian et al. 2006)

along with the Environmental Protection Act. Recent CRZ notification has identified
4 CRZ zones with additional measures to regulate harmful activities for coastal
community and the environment (Chandramohan 2019). Since the beginning of the
coastal regulation processes, Mangroves have been found among the first prioritized
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 531

coastal habitats and are covered under CRZ 1A as they are identified as Ecologically
Sensitive areas. An association of national and regional organizations and academic
institutions is authorized to identify and prepare the maps of all four CRZ zones.
Preparation of Hazard line, one of the inputs of CRZ mapping itself is the composite
of the extents of the flooding on the land area due to water level fluctuations, sea
level rise, and shoreline changes (erosion or accretion) occurring over a period of
time which has been prepared by Survey of India and further classification of CRZ
and mapping at 1:25,000 scale involve the development of exhaustive GIS layers
and various kinds of spatial analysis discussed in Sect. 21.7.1.4 (NCSM 2020).

21.7.4 Integrated Coastal Zone Management

In 1979, the Indian National Mangrove Committee recommended areas of research


and development in mangrove mapping, quantitative survey of seasonal variations
environmental parameters, vegetation growth climatological variability for manage-
ment of the mangroves which initiated national level application of remote sensing
derived maps and field survey data in an integrated manner (Anonymous 2020; FSI
2017). Later the concept of ICZM was evolved during the Earth Summit of Rio de
Janeiro in 1992 aiming to achieve the sustainable management of the coastal zone to
ensure optimal utilization of coastal resources, maintaining the biodiversity and
conserving the critical habitat in an integrated manner regardless of the ecological
and political boundaries (FAO 2020). In India, the resource assessment activity of
ICZM was carried out for the entire coast by developing a geospatial database using
remote sensing and GIS of field surveyed data and is being used to monitor the
conservation and restoration of coastal resources including mangroves (https://www.
ncscm.res.in/). National level GIS based Critical Habitat Information System (CHIS)
of the coastal wetlands including mangrove wetlands was developed to design
appropriate management solutions to address the coastal problems (ICMAM 2000).
ENVironment Information System (ENVIS) was the initiative by MoEF in
1982–83 to provide scientific, technical, and semi-technical information on the
environment of national natural ecosystems. There are about 10 ENVIS centres
providing data on various aspects of ecological systems and environments including
mangrove ecosystem, its flora and fauna. The major centres that deal with mangrove
ecosystems are the one attached to the Institute of Ocean Management, Anna
University (http://www.iomenvis.in/) and the other to Centre for Advanced Studies
in Marine Biology (http://www.casmbenvis.nic.in/).

21.8 Streamlining the Geospatial Techniques for Mangrove

A large variation in assessment of the area of each mangrove forest in different


publications for the same period can be noticed in the literature. Such variation in
assessments is mainly due to, (i) consideration of reserved or protected forest
boundary in some studies, (ii) the varied extent of study area boundary depending
532 L. Gnanappazham et al.

on the area of interest of researchers (i.e., bounding box defined for the study area),
and (iii) from RS technology perspective, the resolution or scale of the data, i.e.,
satellite data or aerial photograph used for mapping. Thus standardization of
procedures with a possible variation due to resolution of data and minimum mapping
unit needs to be undertaken. Standardization needs to be extended to the classifica-
tion scheme, scale of mapping, species mapping (inclusion or exclusion of
associated mangroves, etc.). It is also important to define standard protocols for
ground truth, field survey, sampling, and validation process. Few recent global
studies have helped in the adoption of such an approach and would be a strong
base for the sustainable conservation and management of mangroves and also
become the base for future research activities.

21.9 Conclusions

Mangroves are an important class of natural vegetation serving diverse ecological


and economical functions at the interface of ocean and land. Given the large spread
and diversity as well as being not so amenable for assessment and monitoring, form a
focused target for its study by remote sensing technology. This chapter presents the
basic concepts of maps, survey methods, and space-based observations and Geo-
graphic Information System for understanding the various aspects of mangroves by
various Geospatial tools and techniques. A list of annexures provides the various
types of remote sensors for land observation and a compilation of literature on
studying the various aspects of mangroves using geospatial techniques. Previous
studies on Indian mangroves have been summarized for their data source, objectives,
study area, etc. The remote sensing techniques and capabilities are even advancing,
which allows for improved mapping, monitoring, and characterization. Use of
microwave and hyperspectral data for assessing different biophysical and biochemi-
cal characterization has been discussed and studies focusing on such fields are listed.
Actions on conservation and management of mangroves require integration of
space-based maps with field observations as well as integrated inputs from various
stakeholders. Such a spatial integration is ideally realized in a GIS framework and
basic details, as well as application of GIS is presented. The chapter also brings out
the crucial requirements for streamlining the Geospatial techniques for its effective
utilization for the better management of mangrove forests in general and specific to
Indian scenario.
21

Annexure 1: List of Land Observation Remote Sensors

No. of Spatial
Altitude Spectral spectral Resolution
SN Sensor Agency Platform (km) Range (nm) bands (m)
Low/Medium Multispectral Remote Sensors
1 Multispectral Scanner (MSS) (1972–1983) NASA Landsat 1–3 915 500–1100 4 60
2 Thematic Mapper (TM) (1982–2013) NASA Landsat 4–5 705 450–2350 7 30
3 LISS (Linear Imagine Self Scanning System)- ISRO-India IRS-1A, IRS-1B & 904 460–860 4 36.25
2 (1988–1997) IRS-P2
4 Enhanced Thematic Mapper (ETM+) (1999) NASA Landsat 7 705 450–2350 7 30
5 Advanced Spaceborne Thermal Emission And NASA-JPL TERRA Satellite 705 520–11650 14 15–90
Reflection Radiometer (ASTER) (1999)
6 ASTER & MERIS (2002) ESA ENVISAT Satellite 783 555–12000 7, 15 1000, 300
& 412–900
7 Advanced Visible and Near-Infrared JAXA ALOS Satellite 702 420–890 4 10
Radiometer (AVNIR-2) (2006)
8 Operational Land Imager (2013)a NASA Landsat 8 705 430–2290 9 30
9 Multispectral Imager (MSI) (2015)a ESA Sentinel-2A 786 443–2190 12 10–60
Geospatial Tools for Mapping and Monitoring Coastal Mangroves

10 LISS-3 (2016)a ISRO-India IRS-1C, IRS-1D & 817 520–1750 4 23.5


RESOURCESAT-
2A
High Resolution Multispectral Remote Sensors
11 Optical Sensor Assembly (OSA) (1999) DigitalGlobe IKONOS 681 450–900 4 4
12 Quickbird (2001) DigitalGlobe Quickbird 450 450–900 4 2.44
13 Kompsat-MSC (2006)a KARI-Korea KOMPSAT-2 685 450–900 4 4
14 REIS (2008) Planet-USA RapidEye 630 440–850 5 5
533

(continued)
No. of Spatial
534

Altitude Spectral spectral Resolution


SN Sensor Agency Platform (km) Range (nm) bands (m)
15 WV110 (2009) & WV-3 MSS (2014)a DigitalGlobe Worldview-2 & 3 770 450–1040 8 1.84
16 High Resolution Optical Imager (HiRI) Airbus PLEIADES 1A/1B 694 450–890 5 2.8
(2012)a Defence and
Space
17 AEISS-A (Advanced Earth Imaging Sensor KARI-Korea KOMPSAT-3A 528 450–900 5 2.8
System-A) (2015)a
18 LISS-IV) (2016)a ISRO-India RESOURCESAT 2A 817 520–860 3 5.8
19 SpaceView 110 Imaging System (2016) DigitalGlobe Worldview-4 617 450–920 4 1.24
20 High Resolution Multi-Spectral (HRMX) ISRO-India Cartosat-2 Series 505 & 470–830 & 4 2 & 1.14
(2016–2018)a, 2019a (2C, 2D, 2E, 2F) & 3 509 450–860
21 SkySat (2018)a Planet-USA SKYSAT 14–15 505 450–900 4 1
a
Currently operational
L. Gnanappazham et al.
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 535

Microwave Remote Sensors

Temporal Band
resolution frequency Spatial
SN Sensor Agency (Days) information resolution
1 ERS-1 SAR (1991), ERS-2 ESA 3, 35, 176 C Band: 30 m
SAR (1995), ENVISAT 5.3 GHz
ASAR (2002)
2 JERS-1 SAR (1992), ALOS JAXA 44, 46 & L Band: 18 m, 10 m
PALSAR (2006) & 6–12 1.275 GHz & 3–10 m
PALSAR 2 (2014)a
3 RADARSAT-1 (1995) & CSA-Canada 24 C Band: 10–100 m
2 (2007)a 5.3 & &
5.405 GHz 3–100 m
4 Radar Imaging Satellite- ISRO-India 14 & 25 X Band: 1–8 m,
2 (RISAT)-2 (2009)a, 9.59 GHz 1–50 m &
RISAT-1 (2012) & RISAT- & 1–8 m
2B (2019)a C Band:
5.350 GHz
5 TerraSAR-X (2007)a & DLR 2 to 4 X Band: 1–18 m
TanDEM-X (2010)a 9.65 GHz
6 Aquarius SAC-D (2011) NASA 7 Active: 76–156 km
1.26 GHz
Passive:
1.413 GHz
7 Sentinel- 1A (2014) and ESA 6 to 12 C Band: 5–100 m
Sentinel 1B (2016)a 5.405 GHz
a
Currently operational

Hyperspectral Remote Sensors (Imaging Spectrometers)

Number
of Spectral Spatial
Altitude Spectral spectral Resolution Resolution
SN Sensor Agency Platform (km) range (nm) bands (nm) (m)
8 AVIRIS NASA— Airborne 20 400–2500 224 10 4 to 20
(1986)a JPL
9 Hyperion NASA EO-1 705 400–2500 220 10 30
(2000) Satellite
10 Compact High ESA Proba-1 556 415–1050 63 1.3–12 18–36
Resolution Satellite
Imaging
Spectrometer
(CHRIS)
(2001)a
11 HySIS (2018)a ISRO-India IMS-2 636.6 400–2400 326 10 30
12 PRISMA ASI-Italy Satellite 614 400–2505 238 10 5–30
(2019)a
(continued)
536 L. Gnanappazham et al.

13 HSI (2020) DLR-ESA EnMAP 653 420–2450 232 6.5–10 30


Satellite
(Upcoming)
14 HISUI (2020) JAXA Satellite 618 400–2505 185 10–12.5 30
(Upcoming)
15 HypsIRI NASA— Satellite 626 380–2500; 217 4–12 60
(2022) JPL (Upcoming) 7500–12000
a
Currently operational

Annexure 2: Mapping and Monitoring of Mangrove Wetlands


Using Remote Sensing and GIS

Sl Study
No location Datasets Techniques Research findings References
1 India Coast IRS 1C L3 K-means Mangroves classified Nayak et al. (1996)
clustering as dense, shrub and
fringe mangroves
2 India coast IRS-1A/1B LISS- Vis. Int and Mangroves area Nayak and
I, IRS-1C/1D L3 digital analysis (1:250,000, 1:50,000 Bahuguna (2001)
(1986–1993) and 1:25,000 maps)
3 Andaman SOI Toposheet, Supervised Loss of 21 sq.km Roy et al. (1991)
Islands, AP, and Landsat (MLC) mangrove lost
TM
4 Tamil Nadu Landsat TM, Vis. Int & Fuzzy Mangrove loss Ramachandran
and SPOT, and IRS Class (1989–1996): ~ 94 sq. et al. (1998)
Andaman L2 km.
Islands,
5 Pichavaram, SOI, Landsat Vis. Int & OnS
Degraded mangrove Selvam et al.
Indiaa 5 TM, and IRS Dig forest cover has (2003)
1D L3 increased by about
90%
6 Mumbaia SPOT-2, IRS IC Vis. Int & OnS 92.94 sq. km. (1990); Vijay et al. (2005)
and 1D Dig 56.40 sq. km. (2001)
7 Pichavaram IRS-1D LISS-III, Multiplicative, Brovey transform Shanmugam et al.
and ERS-2 SAR Brovey, IHS and outperforms other (2005)
PC transforms methods
8 Godavari SOI, Landsat Vis. Int 368 ha—mangrove Ramasubramanian
Estuary, 5 TM, IRS 1C/1D increased et al. (2006)
Indiaa L3
9 Goa Coast SPOT-1, IRS-1C/ MLC MLC (OA—88.43%) Mani Murali et al.
1D (2006)
10 Balasore, Landsat MSS, Vis. Int & Mangrove lost— Reddy and
Indiaa IRS P6 L3, SOI overlay analysis 330 ha Pattanaik (2007)
in GIS
11 Sagar Island, IRS IC L3 Supervised. 2.1 sq. km. (1998); Dinesh Kumar
India 1.3 sq. km. (1999) et al. (2007)
12 Sundarbans, Geocover MLC 588,695.5 ha (1970’s); Giri et al. (2007)
Bangladesh Landsat Mosaic 581,642.2 ha (2000’s)
and India
(continued)
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 537

13 Godavari, IRS 1B L2 and MLC 1,250 ha of mangroves Satapathy et al.


AP IRS P6 L3 lost to aquaculture, (2007)
and deforestation
14 Gulf of Resourcesat I L4, GMLC OA—85.53% Chatterjee et al.
Mannar, and IRS IC/ID L3 (pre-tsunami) and 88% (2008)
Andaman (post-tsunami)
15 Godavari Landsat MLC 194.8 sq. km. (1977); Reddy and Roy
Delta MSS/TM/ETM+, 186.0 sq. km. (2005) (2008)
and IRS P6 L3
16 Global Global Land Digital analysis Total mangrove— Giri et al. (2011)
Mangroves Survey (GLS) (mosaic) 137,760 sq.km
2000 mosaic;
Landsat
17 Mahanadi Landsat Vis. Int & OnS 2606 ha mangrove Pattanaik and
deltaa MSS/TM, and Dig area loss Narendra Prasad
IRS P6 L3 (2011)
18 Pichavarama Landsat MSS, Vis. Int & Dig. 471 ha lost Gnanappazham
TM, ETM+, IRS Analysis (1970–1991); 531 ha and Selvam (2011)
L3, SOI toposheet gain (1991–2011)
19 Pichavaram, Landsat TM, ISODATA 2.51 sq. km. area Srinivasa Kumar
TN Resourcesat-1 L4 increase (1991–2006) et al. (2012)
20 India Coast Resourcesat-1 ISODATA and Total area— Ajai et al. (2012)
L3/IV supervised 495,620 ha.
21 Kachchh IRS Resourcesat- Vis. Int, GIS Inter-tidal mudflat Mahapatra et al.
coast, 1 LISS-IV overlay area—1797.91 sq. km. (2013)
Gujarata
22 Piram Island, IRS-1D L4, Vis. Int & OnS Mangrove area loss— Bhavsar et al.
Gujarat a IRS-P6 LISS-III Dig 0.74 sq. km. to 0.23 sq (2014)
km. (2007–2013)
23 Gujarat IRS-P6 L3 Vis. Int & OnS Mangrove area— Ajay et al. (2014)
Coasta Dig 996.26 sq. km
24 Andaman IRS P6 L3 Unsupervised; Mangrove area lost— Sachithanandam
Supervised 188.02 ha et al. (2014)
25 Thuraikkadu EO-1 Hyperion SAVI, SVM OA—80.89% Vidhya et al.
RF, TN data (healthy); 86.59% (2014)
(degraded); 76.28%
(sparse)
26 Kachchha IRS-P6 L3; Vis. Int & OnS 9,894 ha (2005); Upadhyay et al.
IRS-RS2 L4 Dig 11,703 ha (2011) (2015)
(2014)
27 Sundarbans, Corona KH, Vis. Int & 6588 sq. km. Ghosh et al. (2015)
Bangladesh Landsat ISODATA (1776);1852 sq.
and India TM/ETM/OLI km. (2014)
28 Indian Landsat Supervised 406 sq. km. increase Jayanthi et al.
Mangroves TM/ETM+ classification (1998–2013) (2018)
29 Pichavaram Landsat ETM+ / MLC 11.41% mangrove Vani and Rama
OLI area increased Chandra Prasad
(2018)
OA—Overall Accuracy; Vis. Int.—Visual Interpretation; OnS Dig: On screen Digitization; MLC—
Maximum Likelihood Classification; ISODATA—Iterative Self-Organizing DATA; SVM—Sup-
port Vector Machine; SAVI—Soil Adjusted Vegetation Index; IHS—Intensitu Hue Saturation
a
RS integrated GIS analysis
538 L. Gnanappazham et al.

Annexure 3: Floristic Discrimination of Indian Mangroves

Research findings
Sl. Study (OA/No. of
No. location Datasets Techniques Species identified) References
30 Bhitarkanika IRS P6 L3 ISODATA, 12 mangrove Reddy et al.
(2004) MLC communities (2008)
31 Lothian IRS P6 L4 Unsupervised 8 mangrove Debashis and
Island (2005) classification classes (73.44%) Karmaker
(2010)
32 Sundarbans IRS 1D LISS- MLC, 5 mangrove Nandy and
III ISODATA classes identified Kushwaha
(2011)
33 Sundarbans Field spectral k-means, 17 mangrove Manjunath et al.
signatures ANOVA, SDA, species (2013)
and Factor
analysis
34 Sundarbans EO-1 Hyperion N-FINDR, 7 mangrove Chakravortty
data ATGP, and species (74.07%) and Choudhury
LSU (2013)
35 Bhitarkanika EO-1 Hyperion MD, SVM, 5 mangrove Kumar et al.
data SAM species (96.85%) (2013)
36 Sundarbans Landsat MLC 6 mangrove Giri et al.
TM/ETM+, species (2014)
EO-1 Hyperion
37 Pichavaram EO-1 Hyperion JM, SAM 2 mangrove Padma and
data classes Sanjeevi (2014)
38 Uttara Landsat GMLC 3 to 5 mangrove Mesta et al.
Kannada TM/ETM+, species (2014)
IRS-P6 L4 MX
39 Bhitarkanika Field and PCA, SDA, JM 8 species of Prasad and
Laboratory distance Rhizophoraceae Gnanappazham
spectra (2016)
40 Pichavaram EO-1 Hyperion SAM, SFF, 10 species Salghuna and
data SID, LSU (SAM + LSU) Pillutla (2017)
41 Sundarbans EO-1 Hyperion MD, SAM, 5 mangrove Kumar et al.
data SVM classes (99.08%) (2019)
42 Lothian EO-1 OBIA 7 mangrove Mondal et al.
Island Hyperion, IRS classes (2019)
Resourcesat-
2 LISS-IV
43 Lothian AVIRIS-NG SAM Lothian Island Chaube et al.
Island and (15 species); (2019)
Bhitarkanika Bhitarkanika
(7 species)
44 Bhitarkanika EO-1 Hyperion SAM 10 species (84%) Anand et al.
data (2020)
OA—Overall Accuracy; SFF—Spectral Feature Fitting; MD—Minimum Distance; MLC—Maxi-
mum Likelihood Classification; SDA—Spectral Distance Analysis; SVM—Support Vector
Machine; SAM—Spectral Angle Mapper; LSU—Linear Spectral Unmixing; JM—Jeffries–
Matusita distance; PCA—Principal Component Analysis; OBIA—Object-Based Image Analysis;
*—RS integrated GIS analysis
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 539

Annexure 4: Studies on Biophysical and Biochemical


Characterization of Indian Mangroves

Sl. Study
No location Datasets Techniques Research findings References
45 Coringa IRS-1C L3 NDVI, MLC, Strong relationship Satyanarayana
ANOVA between VI and basal area et al. (2001)
46 Sundarbans IRS- NDVI, OSAVI, Total AGB—236 metric Manna et al.
Resourcesat and TDVI; Plot tons (2014)
2 L4 biomass, CHN
analysis
47 Sundarbans Landsat MLC, AGB— Potential CO2 emission is Akhand et al.
MSS/TM/ IPCC 1567.98  551.69 Gg (2017)
ETM+ (1975–2013)
48 Bhitarkanika Worldview- GLCM texture, Texture of band ratio Prasad and
2,Plot VI’s highly correlates with plot Gnanappazham
Biomass biomass (2018)
49 Andaman EO-1 Correlates Identified bands sensitive George et al.
Hyperion; in-situ LAI and to LAI (2018)
Plot Hyperspectral
Biomass VI’s
50 Bhitarkanika EO-1 Five ML Total AG Carbon Anand et al.
Hyperion models applied 459.82 kt. C. (EVI); (2020)
data on NDVI and 514.47 kt. C. (NDVI)
EVI
51 India Coast SPOT-VGT NDVI 38% (very healthy) and Chellamani
incl. (March to Maximum 27% (healthy) et al. (2014)
Andaman May) Value
(1999–2008) Compositing
(MVC)
52 Middle EO-1 Hyperspectral Sensitive wavelengths— George et al.
Andaman Hyperion VI’s—SR, 549 nm, 559 nm (green), (2019)
Island data MSR, NDVI, 702 nm, 722 nm, 742 nm,
and NLI and 763 nm (red-edge)
53 Lothian AVIRIS-NG 9 Vegetation 56% of the mangrove Hati et al.
Island Indices, increase in stress (2020)
Red-Edge
Analysis
GLCM—Grey Level Co-occurrence Matrix; SR—Simple Ratio; NLI—Non-Linear Index; MSR—
Modified Simple Ratio; AGB—Above Ground Biomass; LAI—Leaf Area Index; TDVI—
Transformed Difference Vegetation Index; (Please refer Muhdoni et al. 2018)
540 L. Gnanappazham et al.

Annexure 5: Shoreline Changes

Sl. Study
No. location Datasets Techniques Research findings References
1 Godavari SOI Toposheet, Vis. Int and OnS Sand spit of Hope Ramasubramanian
Estuary Landsat 5 TM, Dig Island has grown et al. (2006)
IRS 1C/1D L3 nearly 2.6 km.
(1937–2001)
2 India Coast Landsat TM, Change 45.5%—erosion, SAC (2017)
ETM, AWiFS detection 35.7%—accretion, https://vedas.sac.
and L4; CSIZ (1989–91 to and 18.79%— gov.in/vedas/node/
Database, 2004–06) map stable; Net 61
on 1:25,000 loss ¼ 73 sq. km
scale
3 Karnataka Landsat DSAS in 70%—stable or ChenthamilSelvan
Coast MSS/TM/ETM ArcGIS accretion, 30%— et al. (2014)
+ erosion (varying
magnitude)
4 Sundarbans IRS 1D L3, Vis. Int and OnS Hooghly island— Raha et al. (2014)
IRS P6 Dig eroding
AWiFS, Thakuran Char, and
Resourcesat Lothian islands—
1/2 L4 accreting
5 Andhra Landsat DSAS in 275 km—erosion; Kankara et al.
Pradesh TM/ETM+, ArcGIS 417 km—accretion; (2015)
coast IRS-P5/P6 153 km—stable.
L3/IV, Cartosat
1
6 Karnataka Landsat DSAS in Ankola & Karwar— Hegde and
Coast TM/ETM+/ ArcGIS accretion; Akshaya (2015)
OLI (1991–2014) Honnavar—erosion
7 Tamil Landsat MSS, DSAS in Max. erosion: Natesan et al.
Nadu Coast TM ETM+, ArcGIS Ennore 26.4 m/yr; (2015)
OLI (1978–2014) Max. accretion:
South of Pulicat
34.3 m/yr
8 India Coast Landsat Long-term shore 33%—erosion, Kankara et al.
TM/ETM+, line change 29%-accretion, and (2018)
Cartosat (1990–2016) 38% stable http://www.nccr.
1 (PAN), (1:25000 scale) gov.in
Resourcesat—
L3, L4
9 Globally ArcGIS 10.4 or Rate-of-change Developing add-on Himmelstoss et al.
applicable higher statistics for a tools for shoreline (2018)
time series of change detection https://code.usgs.
shoreline vector DSAS v 5.0 gov/cch/dsas
data
(continued)
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 541

10 Andhra Landsat DSAS in Maximum accretion: Tyagi and Rai


Pradesh MSS/TM/ETM ArcGIS Vishakhapatnam (2020)
coast + (1973–2015) and Srikakulam
coast;
Maximum Erosion:
Southern part of
Krishna delta
11 Gujarat Landsat TM, GIS Overlay Kori Creek— Das (2020)
coast Resourcesat analysis increased erosion;
2—L3 Gulf of Kachchh—
high accretion
DSAS—Digital Shoreline Analysis System; Vis. Int.—Visual Interpretation; OnS Dig: On-screen
Digitization

Annexure 6: Coastal Zone Management

Sl. Study
No. location Datasets Techniques Research findings References
12 Global: Landsat MSS, Mapping, sample Assessment, FAO (1994)
Thailand, SPOT, SIR-A plots and overlay planning,
Sundarbans analysis conservation and
management
13 Indian IRS L3 TM and GIS and Mapping, NCCR
mangroves Environmental hydrological Monitoring, (1999–2002)
parameters—soil, modelling environment https://www.
water quality, impact assessment, nccr.gov.in/?
phyto, q¼publication/
zooplankton, etc. technical-
reports-
14 East coast Landsat TM, IRS Change detection, Identification of Selvam et al.
mangroves L3 Overlay analysis conservation and (2002)
of environmental protection zones,
parameters Joint Mangrove
Management
15 Gulf of Hydrodynamic, Numerical Priority index due Kankara and
Kutchch meteorological, modelling and to risk of Oil spill Subramanian
and spatio- GIS based identified in the (2007)
temporal Environmental order of coral reefs,
behaviour of sensitivity mangroves,
oil mass. analysis and risk mudflats and rocky
assessment coast
(continued)
542 L. Gnanappazham et al.

16 Pichavaram Landsat TM, IRS Visual 22 ha of Jayanthi et al.


L3 (1987, 1994, interpretation and mangroves (2007)
1998 and 2004) Change detection converted to
and field sample aquaculture and
data reconversion of
15 ha of aqua
farms to
agriculture
17 Godavari Landsat MSS, MCL, change Extent of Reddy and Roy
TM, IRS L3 detection encroachment (2008)
18 Chennai RTK DGPS data Spatial Inundation Satheesh
on run up level, Interpolation and Vulnerability Kumar et al.
IRS L3 and L4 Overlay analysis assessment (2008)
(Tsunami 2004)
19 Indian IRS L3 and L4, Image processing, Normalized health Ajai et al.
Mangroves field data from Weighted overlay index of (2013)
sample survey analysis mangroves to
identify hotspots
20 Gulf of LULC, Overlay analysis Suitable sites for Mahapatra et al.
Kutchch geomorphology, mangrove (2013)
wave height and plantation
tidal range
21 Mumbai Allometry, CHN Estimated Patil et al.
analysis and GIS C ¼ 34.14769 T/ha (2014)
analysis
22 Pichavaram, RS data, tidal GIS overlay, Freshwater and Gnanappazham
amplitude, time proximity analysis Tidal dynamics and Selvam
series freshwater and impacts on (2014)
data mangroves
23 Pichavaram RS data, field Numerical tidal Categories of Sathyanathan
samples simulation model mangroves based et al. (2014)
on tidal influence
24 Gujarat Coastal Coastal High to very high Mahapatra et al.
coast geomorphology, Vulnerability risk category: (2015)
slope, SLC rate, Index (CVI) for 85 km (45.67%),
Tidal range and anticipated sea Moderate to low
wave height level rise risk: 934 km
(54.33%)
25 Gulf of IRS L3 and L4 GIS Overlay Block level Upadhyay et al.
Kutchch, analysis mangrove (2015)
Gujarat change—
Restoration
26 Krishna Resourcesat L3, MS4W Web GIS of Jayakumar K
estuary Landsat TM, (MapServer) and mangroves (2019)
ETM PostgreSQL wetland maps,
changes and
27 Goa and FSI maps IRS 1A, 1B Identification of Kumar R
Andaman sites for (2020)
conservation and
plantation
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 543

References
Adam E, Mutanga O, Rugege D (2009) Multispectral and hyperspectral remote sensing for
identification and mapping of wetland vegetation: a review. Wetl Ecol Manag 18:281–296.
https://doi.org/10.1007/s11273-009-9169-z
Ajai, Bahuguna A, Chauhan HB, et al (2012) Mangrove Inventory of India at Community Level.
Natl Acad Sci Lett 36:67–77. https://doi.org/10.1007/s40009-012-0087-x
Ajai, Bahuguna A, Chauhan, H.B., Sarma, K.S., Bhattacharya S, Ashutosh S, Pandey, C.N.,
Thangaradjou T, Gnanppazham L, Selvam, V. and Nayak, S.R. (2013). Mangrove inventory
of India at community level, National Academy Science Letters, 36(1):67–77.
Ajay P, Vijay S, Mehmood K, et al (2014) Mapping and Monitoring of Mangroves in the Coastal
Districts of Gujarat State using Remote Sensing and Geo-informatics. Asian J Geoinformatics
14:15–26.
Akhand A, Mukhopadhyay A, Chanda A, et al (2017) Potential CO2Emission Due to Loss of
Above Ground Biomass from the Indian Sundarban Mangroves During the Last Four Decades. J
Indian Soc Remote Sens 45:147–154. https://doi.org/10.1007/s12524-016-0567-4
Al-Naimi N, Al-Ghouti MA, Balakrishnan P (2016) Investigating chlorophyll and nitrogen levels of
mangroves at Al-Khor, Qatar: an integrated chemical analysis and remote sensing approach.
Environ Monit Assess 188: https://doi.org/10.1007/s10661-016-5269-4
Alongi DM (2014) Carbon cycling and storage in mangrove forests. Ann Rev Mar Sci 6:195–219.
https://doi.org/10.1146/annurev-marine-010213-135020
Alongi DM (2002) Present state and future of the world’s mangrove forests. Environ Conserv
29:331–349. https://doi.org/10.1017/S0376892902000231
Amma PKG, Bhaskaran PK (2020) Role of mangroves in wind-wave climate modeling—A review.
J Coast Conserv 24:1–14. https://doi.org/10.1007/s11852-020-00740-0
Anand A, Pandey PC, Petropoulos GP, et al (2020) Use of hyperion for mangrove forest carbon
stock assessment in bhitarkanika forest reserve: A contribution towards blue carbon initiative.
Remote Sens 12:. https://doi.org/10.3390/rs12040597
Aneece I, Epstein H (2017) Identifying invasive plant species using field spectroscopy in the VNIR
region in successional systems of north-central Virginia. Int J Remote Sens 38:100–122. https://
doi.org/10.1080/01431161.2016.1259682
Anonymous (2020) Mangrove Forest in India, http://old.cwc.gov.in/CPDAC-Website/Paper_
Research_Work
ASD (2001) Field spectrometry: Techniques and instrumentation
Asner GP, Martin RE (2008) Spectral and chemical analysis of tropical forests: Scaling from leaf to
canopy levels. Remote Sens Environ 112:3958–3970. https://doi.org/10.1016/j.rse.2008.07.003
Awange JL, Kyalo Kiema JB (2013) Fundamentals of Remote Sensing. In: Awange JL, Kyalo
Kiema JB (eds) Environmental Geoinfoirmatics: Monitoring and Management. Springer, pp
111–118
Bhavsar DO, Jasrai YT, Pandya HA, et al (2014) Monitoring Mangrove Status using Remote
Sensing and Geo-informatics in Piram Island, Gulf of Khambhat, Gujarat State, India. Int J Sci
Eng Res 5:999–1005
Blasco F, Lavenu F, Baraza J (1986) Remote sensing data applied to mangrove of Kenya coast. In:
Proceedings of the 20th International Symposium on Remote Sensing of the Environment
Programme 3, pp 1465–1480.
Bosire JO, Dahdouh-Guebas F, Walton M, et al (2008) Functionality of restored mangroves: A
review. Aquat Bot 89:251–259. https://doi.org/10.1016/j.aquabot.2008.03.010
Bréda NJJ (2008) Leaf Area Index. In: Encyclopedia of Ecology. Academic Press. pp. 2148–2154
Bunting P, Rosenqvist A, Lucas RM, et al (2018) The global mangrove watch—A new 2010 global
baseline of mangrove extent. Remote Sens 10:. https://doi.org/10.3390/rs10101669
Chakravortty S, Choudhury AS (2013) Determining spatial location of sub pixels in hyperspectral
data for mangrove species identification. In: 2013 International Conference on Signal
Processing, Image Processing & Pattern Recognition. IEEE, pp 39–43
544 L. Gnanappazham et al.

Chandramohan P (2019) Quick Reference for CZ for Indian Coastline—CRZ classification 2019.
www.indomer.com
Chatterjee B, Porwal MC, Hussin YA (2008) Assessment of Tsunami Damage To Mangrove in
India Using Remote Sensing and Gis. In: The International Archives of the Photogrammetry,
Remote Sensing and Spatial Information Sciences. pp 283–288
Chaube NR, Lele N, Misra A, et al (2019) Mangrove species discrimination and health assessment
using AVIRIS-NG hyperspectral data. Curr Sci 116:1136–1142. https://doi.org/10.18520/cs/
v116/i7/1136-1142
Chellamani P, Singh CP, Panigrahy S (2014) Assessment of the health status of Indian mangrove
ecosystems using multi temporal remote sensing data. Trop Ecol 55:245–253
ChenthamilSelvan S, Kankara RS, Rajan B (2014) Assessment of shoreline changes along
Karnataka coast, India using GIS & remote sensing techniques. Indian J Geo-Marine Sci
43:1286–1291
Chow J (2018) Mangrove management for climate change adaptation and sustainable development
in coastal zones. J Sustain For 37:139–156. https://doi.org/10.1080/10549811.2017.1339615
Coetzee S, Ivánová I, Mitasova H, Brovelli MA (2020) Open geospatial software and data: A
review of the current state and a perspective into the future. ISPRS Int J Geo-Information
9:1–30. https://doi.org/10.3390/ijgi9020090
Dalrymple RW, Zaitlin BA, Boyd R (1992) Estuarine facies models: conceptual basis and strati-
graphic implications. J Sediment Petrol 62:1130–1146. https://doi.org/10.1306/D4267A69-
2B26-11D7-8648000102C1865D
Das S (2020) Does mangrove plantation reduce coastal erosion? Assessment from the west coast of
India. Reg Environ Chang 20. https://doi.org/10.1007/s10113-020-01637-2
DasGupta R, Shaw R (2013) Changing perspectives of mangrove management in India—An
analytical overview. Ocean Coast Manag 80:107–118. https://doi.org/10.1016/j.ocecoaman.
2013.04.010
Datt B, Mcvicar TR, Niel TG Van, et al (2003) Preprocessing EO-1 Hyperion Hyperspectral Data to
Support the Application of Agricultural Indexes. IEEE Trans Geosci Remote Sens
41:1246–1259
Debashis M, Karmaker S (2010) Mangrove Classification in Sundarban using High Resolution
Multi-Spectral Remote Sensing Data and GIS. Asian J Environ Disaster Manag 02:197–207.
https://doi.org/10.3850/s179392402010000268
Dinesh Kumar PK, Gopinath G, Laluraj CM, et al (2007) Change detection studies of Sagar Island,
India, using Indian remote sensing satellite 1C linear imaging self-scan sensor III data. J Coast
Res 23:1498–1502. https://doi.org/10.2112/05-0599.1
Doyle TW, Girod GF (1997) The Frequency and Intensity of Atlantic Hurricanes and Their
Influence on the Structure of South Florida Mangrove Communities. In: Diaz HF, Pulwarty
RS (eds) Hurricanes. Springer Berlin Heidelberg, pp 109–120
Dronova I (2015) Object-Based Image Analysis in Wetland Research: A Review. Remote Sens
7:6380–6413. https://doi.org/10.3390/rs70506380
Duke N, Meynecke J, Dittmann S (2007) A world without mangroves? Science (80) 317:41–43
Ellenbogen K (2012) Coastal Ecosystems: Why sound management of these key natural carbon
sinks matter for greenhouse gas emissions and climate change? The Blue Carbon Initiative
FAQs. https://www.car-spaw-rac.org/IMG/pdf/BC_FAQ_UNFCCC-2.pdf
Ellison JC (2018) Biogeomorphology of mangroves. Coast Wetl An Integr Ecosyst Approach
687–715. https://doi.org/10.1016/B978-0-444-63893-9.00020-4
FAO (1994) Mangrove forest management guidelines. FAO For Pap 117. Rome. pp 319
FAO (2007) The world’s mangroves 1980–2005. FAO For Pap 153:89. https://doi.org/978-92-5-
105856-5
FAO (2020) http://www.fao.org/forestry/mangrove/3942/en/
Fatoyinbo TE, Armstrong AH (2010) Remote Characterization of Biomass Measurements: Case
Study of Mangrove Forests. In: Momba MNB, Bux F (eds) Biomass. Sciyo, pp 65–78
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 545

Friess DA, Rogers K, Lovelock CE, et al (2019) The State of the World’s Mangrove Forests: Past,
Present, and Future. Annu Rev Environ Resour 44:1–27
FSI (2017) Mangrove Cover in India: India State of Forest Report 2017 by Forest Survey of India
FSI (2019) Mangrove Cover in India: India State of Forest Report 2019 by Forest Survey of India
Fu W, Ma J, Chen P, Chen F (2020) Remote Sensing Satellites for Digital Earth. In: Guo H,
Goodchild MF, Annoni A (eds) Manual of Digital Earth. Springer Singapore, pp 55–123
George R, Padalia H, Sinha SK, Kumar AS (2018) Evaluation of the use of hyperspectral vegetation
indices for estimating mangrove leaf area index in middle Andaman Island, India. Remote Sens
Lett 9:1099–1108. https://doi.org/10.1080/2150704X.2018.1508910
George R, Padalia H, Sinha SK, Kumar AS (2019) Evaluating sensitivity of hyperspectral indices
for estimating mangrove chlorophyll in Middle Andaman Island, India. Environ Monit Assess
191. https://doi.org/10.1007/s10661-019-7679-6
Ghosh A, Schmidt S, Fickert T, Nüsser M (2015) The Indian Sundarban mangrove forests: History,
utilization, conservation strategies and local perception. Diversity 7:149–169. https://doi.org/10.
3390/d7020149
Giri C, Ochieng E, Tieszen LL, et al (2011) Status and distribution of mangrove forests of the world
using earth observation satellite data. Glob Ecol Biogeogr 20:154–159. https://doi.org/10.1111/
j.1466-8238.2010.00584.x
Giri C, Pengra B, Zhu Z, et al (2007) Monitoring mangrove forest dynamics of the Sundarbans in
Bangladesh and India using multi-temporal satellite data from 1973 to 2000. Estuar Coast Shelf
Sci 73:91–100. https://doi.org/10.1016/j.ecss.2006.12.019
Giri S, Mukhopadhyay A, Hazra S, et al (2014) A study on abundance and distribution of mangrove
species in Indian Sundarban using remote sensing technique. J Coast Conserv 18:359–367.
https://doi.org/10.1007/s11852-014-0322-3
Gnanappazham L (2008) A remote sensing and GIS based decision support system for effective
management of Pichavaram mangrove wetland, South India. Dissertation, University of Madras
Gnanappazham L, Selvam V (2011) The dynamics in the distribution of mangrove forests in
Pichavaram, South India—perception by user community and remote sensing. Geocarto Int
26:475–490. https://doi.org/10.1080/10106049.2011.591943
Gnanappazham L and Selvam V (2014) Response of mangroves to the change in tidal and fresh
water flow–A case study in Pichavaram, South India. Ocean and Coast Manage 102, 131–138.
Goodchild M (2005) GIS and Modeling Overview. In: Maguire DJ, Batty M, Goodchild MF (eds)
GIS, Spatial Analysis, and Modeling. ESRI Press, pp 1–17.
Green EP, Clark CD, Mumby PJ, et al (1998) Remote sensing techniques for mangrove mapping.
Int J Remote Sens 19:935–956. https://doi.org/10.1080/014311698215801
Guo D, Onstein E (2020) State-of-the-art geospatial information processing in NoSQL databases.
ISPRS Int J Geo-Information 9. https://doi.org/10.3390/ijgi9050331
Gupta K, Mukhopadhyay A, Giri S, et al (2018) An index for discrimination of mangroves from
non-mangroves using LANDSAT 8 OLI imagery. MethodsX 5:1129–1139. https://doi.org/10.
1016/j.mex.2018.09.011
Hamilton SE, Casey D (2016) Creation of a high spatio-temporal resolution global database of
continuous mangrove forest cover for the 21st century (CGMFC-21). Glob Ecol Biogeogr
25:729–738. https://doi.org/10.1111/geb.12449
Hati JP, Goswami S, Samanta S, et al (2020) Estimation of vegetation stress in the mangrove forest
using AVIRIS-NG airborne hyperspectral data. Model Earth Syst Environ. https://doi.org/10.
1007/s40808-020-00916-5
Hegde AV, Akshaya BJ (2015) Shoreline Transformation Study of Karnataka Coast: Geospatial
Approach. Aquat Procedia 4:151–156. https://doi.org/10.1016/j.aqpro.2015.02.021
Heumann BW (2011) Satellite remote sensing of mangrove forests: Recent advances and future
opportunities. Prog Phys Geogr 35:87–108. https://doi.org/10.1177/0309133310385371
Himmelstoss EA, Henderson RE, Kratzmann MG, Farris AS (2018) Digital Shoreline Analysis
System (DSAS) Version 5.0 User Guide
546 L. Gnanappazham et al.

ICMAM (2000) Report on “Critical habitat Information System using GIS for Pichavaram Man-
grove South India. Available via http://www.icmam.gov.in. Accessed on August, 2020
Jackson RD, Pinter PJ, Reginato RJ, Idso SB (1980) Hand-held radiometry. In: Agriculitural
Reviews and Manuals ARM—W—19/October 1980. U.S. Deaprtment of Agriculture Report.
p 66
Jayakumar K (2019) Managing Mangrove Forests Using Open Source-Based WebGIS, Coastal
Management, Global Challenges and Innovations. Academic Press, pp 301–321.
Jayanthi M, Gnanappzhaam L and Ramachandran S (2007) Assessment of Impact of Aquaculture
on Mangrove Environments in the South East Coast of India Using Remote Sensing and
Geographical Information System (GIS). Asian Fisheries Science 20:325–338
Jayanthi M, Thirumurthy S, Nagaraj G, et al (2018) Spatial and temporal changes in mangrove
cover across the protected and unprotected forests of India. Estuar Coast Shelf Sci 213:81–91.
https://doi.org/10.1016/j.ecss.2018.08.016
Jensen JR (2014) Remote Sensing of the Environment: an Earth Resource Perspective, Second
Edition. Pearson Education Limited
Jensen JR (2016) Introductory Digital Image Processing: A Remote Sensing Perspective. Pearson
Prentice Hall, Pearson Eductaion Inc., New Jersey, USA
Jensen JR, Lin H, Yang X, et al (1991) The measurement of mangrove characteristics in Southwest
Florida using SPOT multispectral data. Geocarto Int 6:13–21. https://doi.org/10.1080/
10106049109354302
Jiménez M, Díaz-Delgado R (2015) Towards a Standard Plant Species Spectral Library Protocol for
Vegetation Mapping: A Case Study in the Shrubland of Doñana National Park. ISPRS Int J
Geo-Information 4:2472–2495. https://doi.org/10.3390/ijgi4042472
Kale MP, Singh S, Roy PS (2002) Biomass and productivity estimation using aerospace data and
geographic information system. Trop Ecol 43:123–136
Kankara RS and Subramanian BR. 2007. Oil Spill Sensitivity Analysis and Risk Assessment for
Gulf of Kachchh, India, using Integrated Modeling. Journal of Coastal Research. 23
(5) 1251–1258.
Kankara RS, Murthy MVR, Rajeevan M (2018) National Assessment of Shoreline changes along
Indian Coast: Status Report for 26 years (1990–2016). Chennai
Kankara RS, Selvan SC, Markose VJ, et al (2015) Estimation of long and short term shoreline
changes along Andhra Pradesh coast using remote sensing and GIS techniques. Procedia Eng
116:855–862. https://doi.org/10.1016/j.proeng.2015.08.374
Kathiresan K (2018) Mangrove forests of India. Curr Sci 114:976–981. https://doi.org/10.18520/cs/
v114/i05/976-981
Klemas V (2013) Remote Sensing of Coastal Wetland Biomass: An Overview. J Coast Res
290:1016–1028. https://doi.org/10.2112/JCOASTRES-D-12-00237.1
Kokaly R, Clark R (1999) Spectroscopic determination of leaf biochemistry using band-depth
analysis of absorption features and stepwise multiple linear regression. Remote Sens Environ
4257:267–287
Kokaly RF, Asner GP, Ollinger S V., et al (2009) Characterizing canopy biochemistry from
imaging spectroscopy and its application to ecosystem studies. Remote Sens Environ 113:
S78–S91. https://doi.org/10.1016/j.rse.2008.10.018
Konecny G (2014) Geoinformation: Remote Sensing, Photogrammetry, and Geographic Informa-
tion Systems, Second Edi. CRC Press, Boka Raton
Kopp S, Becker P, Doshi A, et al (2019) Achieving the full vision of earth observation data cubes.
Data 4:. https://doi.org/10.3390/data4030094
Kuenzer C, Bluemel A, Gebhardt S, et al (2011) Remote Sensing of Mangrove Ecosystems: A
Review. Remote Sens 3:878–928. https://doi.org/10.3390/rs3050878
Kuenzer C, Ottinger M, Wegmann M, Guo H (2014) Earth observation satellite sensors for
biodiversity monitoring: potentials and bottlenecks. Int J Remote Sens 35:6599–6647. https://
doi.org/10.1080/01431161.2014.964349
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 547

Kumar R (2020) Conservation and management of mangroves in India, with special reference to the
State of Goa and the Middle Andaman Islands, http://www.fao.org/3/x8080e/x8080e07.htm
Kumar T, Mandal A, Dutta D, et al (2019) Discrimination and classification of mangrove forests
using EO-1 Hyperion data: a case study of Indian Sundarbans. Geocarto Int 34:415–442. https://
doi.org/10.1080/10106049.2017.1408699
Kumar T, Panigrahy S, Kumar P, Parihar JS (2013) Classification of floristic composition of
mangrove forests using hyperspectral data: case study of Bhitarkanika National Park, India. J
Coast Conserv Plan Manag 17:121–132. https://doi.org/10.1007/s11852-012-0223-2
Kutser T, Miller I, Jupp DLB (2006) Mapping coral reef benthic substrates using hyperspectral
space-borne images and spectral libraries. Estuar Coast Shelf Sci 70:449–460. https://doi.org/
10.1016/j.ecss.2006.06.026
Lillesand TM, Kiefer RW, Chipman JW (2004) Remote Sensing and Image Interpretation, Fifth
Edit. John Wiley and Sons Ltd, Hoboken, NJ, USA
Lu D, Chen Q, Wang G, et al (2014) A survey of remote sensing-based aboveground biomass
estimation methods in forest ecosystems. Int J Digit Earth 1–43. https://doi.org/10.1080/
17538947.2014.990526
Lugo AE, Snedaker SC (1974) The Ecology of Mangroves. Annu Rev Ecol Syst 5:39–64
Mahapatra M, Ramakrishnan R, Rajawat AS (2015) Coastal vulnerability assessment of Gujarat
coast to sea level rise using GIS techniques: a preliminary study. J Coast Conserv 19:241–256.
https://doi.org/10.1007/s11852-015-0384-x
Mahapatra M, Ratheesh R, Rajawat AS (2013) Potential Site Selection for Mangrove Plantation
Along the Kachchh District, Gujarat, India Using Remote Sensing and Gis Techniques. Int J
Geol Earth Environ Sci 3:18–23.
Manakos I, Manevski K, Petropoulos GP (2010) Development of a spectral library for mediterra-
nean land cover types. Proc 30th EARSeL Symp Remote Sens Sci Educ Nat Cult Herit 663–668
Mani Murali R, Vethamony P, Saran AK, Jayakumar S (2006) Change detection studies in coastal
zone features of Goa, India by remote sensing. Curr Sci 91:816–820
Manjunath K, Kumar T, Kundu N, Panigrahy S (2013) Discrimination of mangrove species and
mudflat classes using in situ hyperspectral data: a case study of Indian Sundarbans. GIScience
Remote Sens 50:400–417
Manna S, Nandy S, Chanda A, et al (2014) Estimating aboveground biomass in Avicennia marina
plantation in Indian Sundarbans using high-resolution satellite data. J Appl Remote Sens
8:083638. https://doi.org/10.1117/1.jrs.8.083638
Marshall M, Thenkabail P (2015) Advantage of hyperspectral EO-1 Hyperion over multispectral
IKONOS, GeoEye-1, WorldView-2, Landsat ETM+, and MODIS vegetation indices in crop
biomass estimation. ISPRS J Photogramm Remote Sens 108:205–218. https://doi.org/10.1016/j.
isprsjprs.2015.08.001
McGlone JC (2004) Manual of Photogrammetry, Fifth Edit. American Society for Photogrammetry
and Remote Sensing.
Menéndez P, Losada IJ, Torres-Ortega S, et al (2020) The Global Flood Protection Benefits of
Mangroves. Sci Rep 10:1–11. https://doi.org/10.1038/s41598-020-61136-6
Mesta PN, Setturu B, Chandran MDS, et al (2014) Inventorying, Mapping and Monitoring of
Mangroves towards Sustainable Management of West Coast, India. J Geophys Remote Sens
3:1000130. https://doi.org/10.4172/2169-0049.1000130
Mitra A (2020) Ecosystem Services of Mangroves: An Overview. In: Mitra A (ed) Mangrove
Forests in India: Exploring Ecosystem Services. Springer, Cham, pp 1–32
MoEF, GoI (2008) Mangroves for the Future: National Strategy and Action Plan, India (Revised
Draft).
Mondal B, Saha AK, Roy A (2019) Mapping mangroves using LISS-IV and Hyperion data in part
of the Indian Sundarban. Int J Remote Sens 40:9380–9400. https://doi.org/10.1080/01431161.
2019.1630784
548 L. Gnanappazham et al.

Mougin E, Proisy C, Marty G, et al (1999) Multifrequency and multipolarization radar backscatter-


ing from mangrove forests. IEEE Trans Geosci Remote Sens 37:94–102. https://doi.org/10.
1109/36.739128
Muhdoni, FF, Sambah AB, Mahmudi M, Wiadnya DGR (2018) Comparison of Different Vegeta-
tion Indices for assessing Mangrove Density using Sentinel-2 Imagery. Int. J. Geomate,
14, 42–51. https://doi.org/10.21660/2018.45.7177.
Muñoz-Huerta RF, Guevara-Gonzalez RG, Contreras-Medina LM, et al (2013) A review of
methods for sensing the nitrogen status in plants: Advantages, disadvantages and recent
advances. Sensors 13:10823–10843. https://doi.org/10.3390/s130810823
Murdiyarso D, Purbopuspito J, Kauffman JB, et al (2015) The potential of Indonesian mangrove
forests for global climate change mitigation. Nat Clim Chang 8–11. https://doi.org/10.1038/
nclimate2734
Mutanga O, Skidmore AK (2004) Hyperspectral band depth analysis for a better estimation of grass
biomass (Cenchrus ciliaris) measured under controlled laboratory conditions. Int J Appl Earth
Obs Geoinf 5:87–96. https://doi.org/10.1016/j.jag.2004.01.001
Nandy S, Kushwaha SPS (2011) Study on the utility of IRS 1D LISS-III data and the classification
techniques for mapping of Sunderban mangroves. J Coast Conserv 15:123–137. https://doi.org/
10.1007/s11852-010-0126-z
Natesan U, Parthasarathy A, Vishnunath R, et al (2015) Monitoring Longterm Shoreline Changes
along Tamil Nadu, India Using Geospatial Techniques. Aquat Procedia 4:325–332. https://doi.
org/10.1016/j.aqpro.2015.02.044
Nayak S, Bahuguna A (2001) Application of remote sensing data to monitor mangroves and other
coastal vegetation of India. Indian J Mar Sci 30:195–213
Nayak S, Chauhan P, Chauhan HB, et al (1996) IR7S-1C applications for coastal zone manage-
ment. Curr. Sci. 70:614–618
Nazrul-Islam AKM (1993) Environment and vegetation of Sundarban mangrove forest. In: Lieth H,
Masoom A (eds) Towards the rational use ofhigh salinity tolerant plants. Kluwer. pp 81–88
Nellemann C, Corcoran E, Duarte CM, et al (2009) Blue carbon: A Rapid Response Assessment.
United Nations Environmental Programme, GRID-Arendal, www.grida.no
NCSM (2020) https://ncscm.res.in/pdf_docs/crz-2019.pdf
NRSC (2017) Natural Resource Census—Land Use Land Cover Database. Technical Report v.1
Padma S, Sanjeevi S (2014) Jeffries Matusita based mixed-measure for improved spectral matching
in hyperspectral image analysis. Int J Appl Earth Obs Geoinf 32:138–151. https://doi.org/10.
1016/j.jag.2014.04.001
Patil V, Singh A, Naik N, Unnikrishnan S (2014) Estimation of carbon stocks in avicennia marina
stand using allometry, CHN analysis, and GIS methods. Wetlands 34:379–391. https://doi.org/
10.1007/s13157-013-0505-y
Pattanaik C, Narendra Prasad S (2011) Assessment of aquaculture impact on mangroves of
Mahanadi delta (Orissa), East coast of India using remote sensing and GIS. Ocean Coast
Manag 54:789–795. https://doi.org/10.1016/j.ocecoaman.2011.07.013
Pham TD, Yokoya N, Bui DT, et al (2019) Remote sensing approaches for monitoring mangrove
species, structure, and biomass: Opportunities and challenges. Remote Sens 11:1–24. https://
doi.org/10.3390/rs11030230
Prasad KA, Gnanappazham L (2016) Multiple statistical approaches for the discrimination of
mangrove species of Rhizophoraceae using transformed field and laboratory hyperspectral
data. Geocarto Int 31:891–912. https://doi.org/10.1080/10106049.2015.1094521
Prasad KA, Gnanappazham L (2018) Estimation of Above Ground Biomass using High Resolution
Multispectral Worldview 2 image. In: Indian Cartographer. pp 569–577
Prasad KA, Gnanappazham L, Selvam V, et al (2015) Developing a spectral library of mangrove
species of Indian East Coast using field spectroscopy. Geocarto Int 30:580–599
Raha AK, Mishra A, Bhattacharya S, et al (2014) Sea Level Rise and Submergence of Sundarban
Islands: A Time Series Study of Estuarine Dynamics. J Ecol Environ Sci 5:114–123
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 549

Ramachandran S, Sundaramoorthy S, Krishnamoorthy R, et al (1998) Application of remote


sensing and GIS to coastal wetland ecology of Tamil Nadu and Andaman and Nicobar group
of islands with special reference to mangroves. Curr Sci 75:236–244
Ramasubramanian R, Gnanappazham L, Ravishankar T, Navamuniyammal M (2006) Mangroves
of Godavari—Analysis Through Remote Sensing Approach. Wetl Ecol Manag 14:29–37.
https://doi.org/10.1007/s11273-005-2175-x
Rao NR, Garg PK, Ghosh SK (2007) Development of an agricultural crops spectral library and
classification of crops at cultivator level using hyperspectral data. Precis Agric 8:173–185.
https://doi.org/10.1007/s11119-007-9037-x
Ravishankar T, Gnanappazham L, Ramasubramanian R, et al (2004) Atlas of Mangrove Wetlands
of India: Part 2 - Andhra Pradesh. M.S. Swaminathan Research Foundation, Chennai, India
Reddy CS, Pattanaik C (2007) Mangrove vegetation assessment and monitoring in Balasore district,
Orissa using remote sensing and GIS. Natl Acad Sci Lett 30:377–381
Reddy CS, Pattanaik C, Murthy MSR (2008) Community zonation of mangroves in Bhitarkanika
Wildlife Sanctury,Orissa, India using IRSP6LISS III data. Proc Natl Acad Sci India Sect B—
Biol Sci 78:246–252
Reddy CS, Roy A (2008) Assessment of Three Decade Vegetation Dynamics in Mangroves of
Godavari Delta, India Using Multi-Temporal Satellite Data and GIS. Res. J. Environ. Sci.
2:108–115
Reddy MA (2008) Text Book of Remote sensing and Geographical Information Systems, Third
Edit. BS Publications, Hyderabad, India
Rees WG (2013) Physical Principles of Remote Sensing: Third Edition. Cambridge University
Press.
Reichardt M (2017) Open Geospatial Consortium standards. Int Encycl Geogr People, Earth,
Environ Technol 1–8. https://doi.org/10.1002/9781118786352.wbieg0348
Richards JA, Jia X (2005) Remote Sensing Digital Image Analysis: An Introduction, 4th Edition.
Springer-Verlag, Berlin
Roy PS, Ranganath BK, Diwakar PG, et al (1991) Tropical forest typo mapping and monitoring
using remote sensing. Int J Remote Sens 12:2205–2225. https://doi.org/10.1080/
01431169108955253
SAC (2017) Shoreline Change Atlas of the Indian Coast Shoreline Change Atlas (1989–91 to
2004–06). Ahmedabad
Sachithanandam V, Mageswaran T, Ragavan P, et al (2014) Mangrove regeneration in tsunami
affected area of north and south Andaman using insitu and remote sensing techniques. Indian J
Geo-Marine Sci 43:1055–1061
Sahu SC, Suresh HS, Murthy IK, Ravindranath NH (2015) Mangrove Area Assessment in India:
Implications of Loss of Mangroves. J Earth Sci Clim Change 06:280. https://doi.org/10.4172/
2157-7617.1000280
Salghuna NN, Pillutla RCP (2017) Mapping Mangrove Species Using Hyperspectral Data: A Case
Study of Pichavaram Mangrove Ecosystem, Tamil Nadu. Earth Syst Environ 1:24. https://doi.
org/10.1007/s41748-017-0024-8
Satapathy DR, Krupadam RJ, Kumar LP, Wate SR (2007) The application of satellite data for the
quantification of mangrove loss and coastal management in the Godavari estuary, East Coast of
India. Environ Monit Assess 134:453–469. https://doi.org/10.1007/s10661-007-9636-z
Satheesh Kumar, P. Arul Murugan, R. R. Krishnamurthy, B. Prabhu Doss Batvari, M. V.
Ramanamurthy, T. Usha, and Y. Pari (2008). Inundation mapping—a study based on December
2004 Tsunami Hazard along Chennai coast, Southeast India. Nat. Hazards Earth Syst. Sci.,
8, 617–626
Sathyanathan R, Thattai D, Selvam V (2014) The Coleroon river flow and its effect on the
Pichavaram mangrove ecosystem. J Coast Conserv 18:309–322. https://doi.org/10.1007/
s11852-014-0313-4
550 L. Gnanappazham et al.

Satyanarayana B, Thierry B, Seen D Lo, et al (2001) Remote sensing in mangrove research-


relationship between vegetation indices and dendrometric parameters: A case for Coringa,
east coast of India. In: 22nd Asian Conference on Remote Sensing. Singapore, pp 5–9
Schölkopf B, Smola AJ (2001) Learning with kernel: Support Vector Machines, Regularization,
Optimization and Beyond. The MIT Press, Cambridge, Massachusetts, London, England,
Cambridge, Massachusetts, London, England
Selvam V, Gnanappazham L, Navamuniyammal M et al (2002) Atlas of mangrove wetlands of
India: part 1 Tamil Nadu, MS Swaminathan Research Foundation, Chennai, p 100.
Selvam V, Ravichandran KK, Gnanappazham L, Navamuniyammal M (2003) Assessment of
community-based restoration of Pichavaram mangrove wetland using remote sensing data.
Curr Sci 85:794–798
Shanmugam P, Manjunath AS, Ahn YH, et al (2005) Application of multisensor fusion techniques
in remote sensing of coastal mangrove wetlands. Int J Geoinformatics 1:1–17
Snyder JP, Maling DH (1993) Flattening the Earth. Nature 366:522–522
Spadling M, Kainuma M, Collins L (2010) World Atlas of Mangroves. A collaborative project of
ITTO, ISME, FAO, UNEP-WCMC, UNESCO-MAB, UNU-INWEH and TNC. Earthscan
Publications, Abingdon, Oxon OX144RN, UK, Abingdon, Oxon OX144RN, UK
Srinivasa Kumar T, Mahendra RS, Nayak S, et al (2012) Identification of hot spots and well
managed areas of Pichavaram mangrove using Landsat TM and Resourcesat-1 LISS IV: An
example of coastal resource conservation along Tamil Nadu Coast, India. J Coast Conserv
16:1–12. https://doi.org/10.1007/s11852-011-0162-3
Tang W, Zheng M, Zhao X, et al (2018) Big Geospatial Data Analytics for Global Mangrove
Biomass and Carbon Estimation. Sustainability 10:472. https://doi.org/10.3390/su10020472
Thakur S, Mondal I, Ghosh PB, et al (2020) A review of the application of multispectral remote
sensing in the study of mangrove ecosystems with special emphasis on image processing
techniques. Spat Inf Res 28:39–51. https://doi.org/10.1007/s41324-019-00268-y
Tomlinson PB (1994) The Botany of Mangroves. Cambridge University Press, Cambridge, UK
Townshend JRG, Justice CO (1986) Analysis of the dynamics of African vegetation using the
normalized difference vegetation index. Int J Remote Sens 7:1435–1445. https://doi.org/10.
1080/01431168608948946
Tso B, Mather PM (2009) Classification Methods for Remotely Sensed Data, Second Edi. CRC
Press, Boka Raton, Boca Raton
Tucker CJ, Sellers PJ (1986) Satellite remote sensing of primary production. Int J Remote Sens
7:1395–1416. https://doi.org/10.1080/01431168608948944
Tyagi S, Rai SC (2020) Monitoring shoreline changes along Andhra coast of India using remote
sensing and geographic information system. Indian J Geo-Marine Sci 49:218–224
Underwood E (2003) Mapping nonnative plants using hyperspectral imagery. Remote Sens Environ
86:150–161. https://doi.org/10.1016/S0034-4257(03)00096-8
Upadhyay R, Joshi N, Sampat AC, et al (2015) Mangrove Restoration and Regeneration Monitor-
ing in Gulf of Kachchh, Gujarat State, India, Using Remote Sensing and Geo-Informatics. Int J
Geosci 06:299–310. https://doi.org/10.4236/ijg.2015.64023
Vani M, Rama Chandra Prasad P (2018) Geospatial assessment of spatio-temporal changes in
mangrove vegetation of pichavaram region, Tamil nadu, India. In: Threats to Mangrove Forests.
Springer. pp 89–102
Vidhya R, Vijayasekaran D, Farook MA, et al (2014) Improved classification of mangroves health
status using hyperspectral remote sensing data. Int Arch Photogramm Remote Sens Spat Inf
Sci—ISPRS Arch XL–8:667–670. https://doi.org/10.5194/isprsarchives-XL-8-667-2014
Vijay V, Biradar RS, Inamdar AB, et al (2005) Mangrove mapping and change detection around
Mumbai (Bombay) using remotely sensed data. Indian J Mar Sci 34:310–315
21 Geospatial Tools for Mapping and Monitoring Coastal Mangroves 551

Woodroffe C (1992) Mangrove Sediments and Geomorphology. In: Robertson AI, Alongi DM
(eds) Tropical Mangrove Ecosystems. American Geophysical Union, Washington, DC
Zheng G, Moskal LM (2009) Retrieving Leaf Area Index (LAI) Using Remote Sensing: Theories,
Methods and Sensors. Sensors 9:2719–2745. https://doi.org/10.3390/s90402719
Zhou G (2001) Architecture of Future Intelligent Earth Observing Satellites (FIEOS) in 2010 and
Beyond: Technical report (June 01, 2001–November 31, 2001)
Zomer R, Trabucco A, Ustin S (2009) Building spectral libraries for wetlands land cover classifica-
tion and hyperspectral remote sensing. J Environ Manage 90:2170–2177. https://doi.org/10.
1016/j.jenvman.2007.06.028.

You might also like