Mass and Energy Balancing Calculations For Plant Chapter
Mass and Energy Balancing Calculations For Plant Chapter
Mass and Energy Balancing Calculations For Plant Chapter
net/publication/350469979
CITATIONS READS
0 2,040
2 authors, including:
S. Feroz
Prince Mohammad University
274 PUBLICATIONS 511 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by S. Feroz on 11 April 2021.
The right of David Pritchard and Shaik Feroz to be identified as authors of this work has been asserted by
them in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988.
Reasonable efforts have been made to publish reliable data and information, but the author and publisher
cannot assume responsibility for the validity of all materials or the consequences of their use. The authors and
publishers have attempted to trace the copyright holders of all material reproduced in this p ublication and
apologise to copyright holders if permission to publish in this form has not been obtained. If any copyright
material has not been acknowledged, please write and let us know so we may rectify in any future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted or utilised in any form by any electronic, mechanical or other means, now known or hereafter
invented, including photocopying, microfilming and recording, or in any information s torage or retrieval
system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, access www.copyright.
com or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,
978-750-8400. For works that are not available on CCC, please contact [email protected]
Trademark notice: Product or corporate names may be trademarks or registered trademarks and are used
only for identification and explanation without intent to infringe.
Typeset in Times
by codeMantra
Contents
Foreword....................................................................................................................ix
Preface.......................................................................................................................xi
Acknowledgements................................................................................................... xv
Nomenclature..........................................................................................................xvii
Authors.....................................................................................................................xix
ix
Preface
The purpose of this book is to provide an example of detailed mass and energy
balances carried out on an existing process. This sort of calculation would, typi-
cally, be required as part of design calculations. Many books that are used to teach
mass and energy balancing illustrate the various pieces of knowledge required
through a series of relevant, but not necessarily integrated, calculations. This
book will not claim to do this, and the various pieces of knowledge required will
be assumed to be in place. What the text will do is give a detailed series of pro-
cess integrated and illustrated calculations. It will attempt to guide the reader as
to how the knowledge that they should have can be realistically applied, and the
series of integrated calculations will lead to the production of a specified amount
of final product.
Some applications of prior assumed knowledge will be laid out in ‘teaching’
appendices. These will be able to be used in conjunction with the main text where
the figures calculated may need a more detailed explanation. This book would, typi-
cally, be used by first- and second-year students as a teaching aid but could give some
useful insights to other students, researchers and industry personnel.
In considering the text, the reader should be aware that they have a responsibility.
Any attempt to demonstrate mass and energy balances means that it is important
that the numbers are understood. That might appear obvious, but it is easy just to see
figures without realising where they came from; an attempt should at least be made
to try to check the genesis of the figures.
Any contact with the energy balance in the text will immediately indicate that the
enthalpy values to be used are usually calculable using a polynomial giving com-
ponent enthalpies as a function of temperature. Finding the enthalpy value can then
become almost a mechanical exercise. How exactly enthalpies are calculated for
energy balances differs according to sources of data. The important thing in this text
is that sources of enthalpy data are discussed and basic equations explored. If these
sections are properly checked, then the background given usually makes the calcula-
tion of enthalpies more than just a mechanical act. At least, unlike some mass and
energy balances, some data are available.
Usually, the data for enthalpy are given in kJ/mol for component enthalpies and
kJ for stream enthalpies. Where appropriate, the mole is sometimes used, but usually
the text deals in kmol and kJ.
It is always interesting to follow the latest convention. The text had to deal
with desulphurisation, desulfurisation or desulfurization. This text has adopted
desulphurisation.
A short examination of the book text will reveal a range of tables with a variety
of numbers in each table. Given the fact that mass and energy balances for a range of
operations are being calculated, this is inevitable. These values are presented in dif-
ferent formats, but in this text, numbers are in the so-called scientific format where
numbers are expressed in exponential form. The number of significant figures is a
xi
xii Preface
matter for debate. It is acknowledged that the number of significant figures should
not show the excesses of spreadsheets sometime seen. Large numbers are rounded
off in scientific notation.
In this text, the number of significant figures in energy balances is set at 5. This is
illustrative and a compromise for the data being used and the data that may be used
as it is available. In this text, various data sources for enthalpy are quoted and the
figures used in the main enthalpy polynomials are reproduced. Using the basis of
the mass balance quoted, relatively large numbers can be generated. The number of
significant figures can be decided by the nature of the data to be used and processed.
The basis in this text is fixed and is not the main purpose of the text. The balances to
be carried out are illustrative and can be adapted as necessary.
A text that attempts to carry out multiple calculations can be difficult to follow.
It is impossible to illustrate every step taken in a calculation, so there has to be an
element of self-instruction and individual responsibility in tracing from where and
why particular numbers are produced. This is particularly true in a text of the type
shown here. The appendices contain material that should be studied in conjunction
with the calculations being carried out.
The production of the syngas can be followed on the block diagram supplied. It is
vital that this is available as the mass and energy balances progress. The various
chapter contents and process progression can be followed using the block diagram,
and it is vital that this is used as the guide as progress is made through the chapters.
It will be observed that Chapter 5 contains fairly standard, orthodox questions
that do relate in many instances to the ammonia synthesis loop. There are usually
no ‘extra’ principles involved in solving these although the manner in which they are
laid out tries to be supportive in producing answers. In teaching, I always found that
many learners were adept at fitting the answer to the question, and it was a mistake
to supply too many details. That has not happened here, and there is a challenge as to
how an answers chapter should be used.
In many ways, this is an experimental text with possible attendant problems.
As has been pointed out, it is worth taking time to go through the calculations
carried out and try to follow the theoretical logic of what is laid out. There is enough
information supplied in the text and appendices for you to attempt to carry out the
calculation. This challenges you to reproduce your own version of the answer and not
necessarily to reproduce the answer in the text.
You should find boxes attached to each chapter which initially try to identify what
will be calculated in the chapter. The box at the end of the chapter tries to summarise
the parameters that have actually been calculated. This is a good exercise in deciding
if you understood how they were obtained and in deciding which other parameters
you might wish to be included or even omitted.
Good luck on your journey.
Preface xiii
xv
Nomenclature
a, b, c, d Polynomial coefficients
a, b, c, d Stoichiometric coefficients
ai Activity of component i
CP Specific heat capacity at constant pressure
CV Specific heat capacity at constant volume
fi Component fugacity
fiθ Standard component fugacity
( ∆G )
T f
Gibbs free energy change of formation at temperature T
∆GT Gibbs free energy change at temperature T
∆GTθ Standard Gibbs free energy change at temperature T
H Power requirement
∆H Enthalpy change
∆H F0 Standard enthalpy of formation
H 0o Standard enthalpy related to a temperature of 0°K.
KP Equilibrium constant
M component moles
N Number of components
n Polytropic index
n Number of moles
P Pressure
pi Partial pressure of component i
P* Standard pressure
q Energy transfer as heat
T Temperature
∆U Internal energy change
w Energy transfer as work
yi Gas-/vapour-phase mole fraction of component i
zi Fractional conversion of component i
γ = CP CV
xvii
Authors
Dr. David Pritchardis a retired academic who continues to work in academic areas
of chemical engineering. He obtained his first degree from the Manchester College
of Science and Technology, Faculty of Technology in the University of Manchester,
UK. He obtained his doctorate from Bath University of Technology (now University
of Bath), UK. On graduation, he initially worked in the process industries for British
Titan (subsequently Tioxide International, UK) as a research officer looking to
improve the plant engineering. From here, he joined the then Teesside Polytechnic,
Department of Chemical Engineering, UK, as a lecturer. The Polytechnic was subse-
quently incorporated as the University of Teesside and still exists as such.
At Teesside, he held every undergraduate course leadership post that then existed
(not all at the same time) and was responsible for course development and, with col-
leagues, produced a pioneering first-year course on the chemical engineering degree.
This first year obtained awards for the novelty of the teaching approach including
recognition from the RSA. He lectured chemical engineering at every level from
undergraduate diploma (HND, HNC), degree (BSc, BEng), postgraduate (MSc) and
design projects. He helped pioneer the introduction of part-time routes for degrees in
chemical engineering and managed many students studying on ERASMUS schemes
in receiving degrees under a special, externally examined, pathway.
He also carried out research into a range of areas particularly in the fields of sepa-
ration processes and phase equilibria. He produced over 30 refereed research papers
together with a number of conference papers and supervised over 20 PhD students and
a number of MSc students. Within the newly formed School of Science and Technology,
he took responsibility for the running of the Division of Chemical Engineering and
oversaw various academic initiatives. He acted as an external examiner for various UK
courses at undergraduate and postgraduate levels and was a PhD external examiner.
Dr. Shaik Feroz is currently working as Research Professor at Prince Mohammed
Bin Fahd University, Kingdom of Saudi Arabia. Dr. Feroz obtained his doctor-
ate in the field of chemical engineering from Andhra University, India, in 2004;
Postdoc Research Fellowship from Leibniz University, Germany, in 2015; M.Tech in
chemical engineering from Osmania University, India, in 1998; B.Tech in chemical
engineering from S. V. University, India, in 1992; and Post-Graduate Diploma in
Environmental Studies from Andhra University in 2003.
Dr. Feroz h as expertise in process engineering, plant design and t roubleshooting,
quality control using advance analytical equipment, wastewater treatment, solar
energy systems (PV and CSP) for desalination, hot water systems and water
t reatment, synthesis of nano-photocatalysts, simultaneous treatment of wastewa-
ter and production of hydrogen, environmental impact assessment, and design,
delivery and management of chemical and process engineering programmes and
tailor-made industrial-based training programmes related to chemical and process
engineering.
xix
xx Authors
Dr. Feroz has more than 170 publications to his credit in journals and conferences
of international repute and supervised four PhD research works with another three
ongoing. Dr. Feroz is associated as the Principal Investigator/Co-Investigator to a num-
ber of research projects and also involved as “Technology transfer agent” by Innovation
Research Center, Sultanate of Oman. He has a significant contribution as editorial
member and peer reviewer for various reputed international journals and conferences.
He was awarded “Best Researcher” by the Caledonian College of Engineering for the
year 2014. Dr. Feroz has total experience of 27 years in teaching, research and industry.
1 Introduction, Reformers
and Stream Energy
Interchange
Input = Output
Those of you who have experienced process engineering in theory and in practice
will know that it is crucial to be able to handle the masses and energies involved in
the chemical reaction as well as non-reacting systems. As usual, thermodynamics
takes us calmly by the shoulder and points out that species may change through
the reaction, but the total mass of atoms confirms that mass and energy input does
indeed equal mass and energy output. The second law of thermodynamics gives us
the concepts and tools to be able to carry out quantitative analysis of the balance
when chemical reaction is also involved.
Many operations within process engineering are carried out at steady state, and
the mass and energy balances do not vary with respect to time; for both mass and
energy, the input equals the output. Mass and energy balances have a time-variant
form if required; a time-variant term, the accumulation or depletion, can be added or
subtracted. This term will not be considered in this text. All production of ammonia
syngas is considered at steady state.
If you are relatively new to process engineering and its component parts, then
it might be useful to read the next few pages. If you know everything there is to
know about it, then the next few pages may be old hat – but still readable. A general
description of process engineering would include the following, this description
includes concepts already covered:
1
2 Mass and Energy Balancing
All engineers usually employ mathematics, physics and the ‘engineering art’ to over-
come technical problems in a safe and economical fashion; usually, it is the process
engineer alone that draws upon the powerful science of chemistry to solve a wide
range of problems.
In practice, process/chemical engineering is the application of basic sciences
(maths, chemistry, physics and biology) and engineering principles to the develop-
ment, design, operation and maintenance of processes to convert raw materials to
useful products and, at the same time, improve the human environment. Process/
chemical engineering involves specifying equipment, operating conditions, instru-
mentation and process control for all these changes. Process/chemical engineers
translate processes developed in the lab into practical applications for the produc-
tion of products such as plastics, medicines, detergents and fuels; design plants to
maximise productivity and minimise costs; and also evaluate plant operations for
performance and product quality.
Thus, as stated, the knowledge base for process/chemical engineers starts from
a basic understanding of chemical engineering principles with a strong chemistry
back ground; material and energy balance calculation, thermodynamics principles,
momentum, heat and mass transfer principles, mechanical separation principles,
manufacturing process flow diagrams; reaction engineering principles, process
instrumentation and control principles, process equipment and plant design princi-
ples, and health and safety management principles; and management and economics.
In addition to this, the knowledge base is extended to complementary areas such as
modelling and simulation principles, optimisation principles, environmental engi-
neering principles, corrosion engineering principles, and nanotechnology.
A structure of the logic used for synthesising and analysing processing schemes in
the chemical and allied industries is essential. Thus, the basic concept adopted is that
all processing schemes can be built into a series of individual, or unit, steps. Usually,
if a step involves a chemical change, it is called a unit process; if physical change,
a unit operation. The basic tools of the chemical engineer for the design, study or
improvement of a unit process are the mass balance, the energy balance, kinetic rate
of reaction and position of equilibrium (the last needs to be included only if the reac-
tion does not go to completion).
In manufacturing, a unit process is a single component part of the end-to-end
manufacturing process that transforms raw materials into finished goods. Unit pro-
cessing is the basic processing in process/chemical engineering. Together with unit
operations, it forms the main principle of the varied chemical industries. Each genre
Reformers and Stream Energy Interchange 3
of unit processing follows the same chemical law much as each genre of unit opera-
tions follows the same physical law.
Material balances are one of the most basic and useful tools in the process/
chemical engineering field. The principles involved have been set out previously in
this introduction. One of the first things that chemical engineers need to know about
processes is the many different ways in which a process may be operated. There are
three major classifications of processes:
i. In a batch process, material is placed in the vessel at the start and (only)
removed at the end – no material is exchanged with the surroundings dur-
ing the process. Examples are baking bread, fermentations and small-scale
chemical production (pharmaceuticals).
ii. In a continuous process, material flows into and out of the process during
the entire duration. Examples are pool filter and distillation processes.
iii. A semi-batch process is one that does not neatly fit into either of the other
categories and often contains elements of both (i.e. it is a catch-all classifica-
tion). Examples are a washing machine and a fermentation with purge.
Again, as previously indicated, each of the above classes of process may be further
distinguished by how they operate with respect to time. For a steady state, none of
the process variables change with time (if we ignore small, random fluctuations);
conversely, a process may be run at unsteady state if the process variables change
with time.
Energy balances are used to quantify the energy used or produced by a system.
Energy balances can be calculated on the basis of external energy used per kilogram
of product, or raw material processed, or on dry solids or some key component; a
basis must be stated in carrying out the balance.
The first law of thermodynamics enables us to understand the mechanisms
involved in energy transfer (heat and work transfer) between a system and its sur-
roundings and produces thermodynamic properties and functions that enable us to
quantify these transfers. At constant pressure, the thermodynamic property enthalpy
is defined and used. This property, at constant pressure, can be shown to quantify
the energy transfer as heat and hence is used in the energy balance. Enthalpy is con-
served and, as with the mass balance, enthalpy balances can be written round the
various items of equipment in the process, either in process stages or round the whole
plant; it is assumed that no appreciable heat is converted to other forms of energy
such as work. Enthalpy (H) is always referred to some reference level or datum, so
that the quantities are relative to this datum. Working out energy balances is then a
matter of considering the various quantities of material involved, their specific heats
and their changes in temperature or state (as quite frequently energies of vapori-
sation (latent heats) arising from phase changes are encountered). In the problems
you will deal with when ammonia syngas production mass and energy balances are
considered, enthalpy values are required, and the availability of the necessary data
is discussed in Appendix 3.
In basic consideration of the process engineering, it is important to be able to deal
with the mass and energy balances in terms of how far a chemical reaction will go
4 Mass and Energy Balancing
under fixed conditions. Again thermodynamics, through the Gibbs energy and the
equilibrium constant, can help to fix the final balances in the plant.
For other specific elements of the plant, thermodynamics can be used to do mean-
ingful calculations on various unit operations; thermodynamics can provide the
equilibrium and enthalpy data needed for design of distillation columns, absorption
columns, evaporators, condensers and other units where heat exchange is involved.
Many chemical processes are concerned with the problem of changing composi-
tion of solutions and mixtures which may or may not involve chemical reactions.
These operations are directed towards separating a substance into its components,
e.g. filtration and screening, which are entirely mechanical operations.
N 2( g) + 3H 2( g) ↔ 2NH 3( g)
In synthesis gas production, the aim is to produce a pure mixture of nitrogen and
hydrogen in the stoichiometric ratio 1:3 from an appropriate feedstock.
This text is based on the production of the synthesis gas using a process based on
the block diagram shown below (Figure 1.1):
It is worth noting that the basic units making up the block diagram are the items
on which we will be carrying out detailed mass and energy balances.
Basically, Step (a) as described above is set out as a series of operations in the
block diagram.
AMMONIA SYNTHESIS GAS PRODUCTION.
24.9 Bar
E103A 387°C V102 397°C E103B 288°C M101 V103
201°C
Converted Gas Low Temperature
Reformers and Stream Energy Interchange
220°C
21.8 Bar
V107 355°C E107 E108
70°C 40°C
Methanator Low Pressure Boiler
21.3 Bar Final Cooler 20.9 Bar
Feed Water Heater 21.1 Bar
then on the block diagram, Item C101 (where the primary feedstock, natural gas,
enters the process via a compressor) to Item E102 (where energy from the initial
production of the N2/H2 mixture is exchanged to provide energy elsewhere in the
process) cover ‘Feedstock pre-treatment and gas generation’.
Item V102 (where CO is removed via the water–gas shift reaction)
CO + H 2O ↔ CO 2 + H 2
to Item E104 (where energy from the water–gas shift converters is used to input
energy to help to raise steam) relate to the carbon monoxide conversion.
Item E105 (where energy is used to recover solvent for absorber V105 by
removing CO2) to Item V105 (where CO2 is actually removed in an appropriate
solvent) cover most of the CO2 removal. Item V107, the methanator, removes
vestiges of CO and CO2 left in the gas stream by converting to methane through
the reactions:
CO + 3H 2 ↔ CH 4 + H 2O
CO 2 + 4H 2 ↔ CH 4 + 2H 2O
With some more heat exchange, the syngas is then ready for the ammonia
synthesis loop.
The scene is set, and the chapters and sections that follow deal with this mass and
energy balancing in the production of ammonia syngas. Each section deals with one
of the sequential elements shown in the block diagram for the process. Each section
takes the reader through the balances required for the mass and energy in the par-
ticular operation and tries to point out some of the other factors that can operate when
carrying out the mass and energy balances.
The mass and energy balances can be carried out but, occasionally, it is helpful to
describe the operation being carried out in slightly more detail. No attempt is made
to show the detailed process design involved in the operation. Again, it is empha-
sised that the text assumes that the basic elements of mass and energy balancing are
familiar to the reader.
Reformers and Stream Energy Interchange 7
In carrying out the energy balancing, a reference has been made to Appendix 3 in
which elements of the approaches to calculation of enthalpy and enthalpy differences
are set out. In some ways, the calculation of enthalpy and enthalpy differences has
been ‘simplified’, by use of polynomials for component enthalpies. The basis of the
polynomials is set out and the basis of temperature explained. The calculation of the
enthalpy change from the defining equation:
T2
∆H =
∫T1
mCP dT
is illustrated by its use in the section dealing with the desulphurisation section of the
process.
Because each chapter with its sections is in itself an exercise in carrying out a
mass and energy balance, no other tutorial exercises are set in the section. Each sec-
tion schedules the calculations required and how it is proposed to carry them out.
At the beginning of each section, a box laying out the basic calculational require-
ments of the section is presented. At the end of each section, a box containing the
figures calculated is presented and a brief discussion of their significance given.
In some of the other chapter sections, there are other calculations presented; e.g.,
is the water vapour in the gas stream likely to condense at the pressure and tempera-
ture figures given? This usually requires some estimate of a gas/vapour dew point.
There is a presupposition that basic ideas related to Dalton’s law and partial pressures
are known and even some understanding of dew point is assumed. There is a brief
explanation of Dalton’s law in Appendix 1. Knowing some basic principles, dew
point calculations are presented as appropriate.
Because the main emphasis of the text is calculation, it is necessary to carry out
the numerical calculations as appropriate. In today’s calculational world, there is a
range of hardware/software available. A perusal of the text will show that things have
been kept basic and relatively straightforward. The calculations have normally been
carried out using Microsoft Excel spreadsheets. The necessary equations are usually
made clear in the text, e.g. Appendix 2. It is made quite clear that any appropriate
calculational aids can be used to carry out calculations and check numerical answers.
Most of the operations will require an element of mass balancing, but the bulk of
those scheduled in this list involves mass balancing with chemical reaction. The
absorber involves internal streams inside the mass balance envelope, and this will
need individual and overall mass balancing.
For energy balance, most of the operations within the block diagram have energy
transfers involved. In addition to the major mass and energy balancing operations
already identified, there are other operations that require balances to be carried out.
The raising of steam using the energy from the syngas stream and the use of the
syngas stream energy for heating purposes are major operations requiring energy
balances.
The main operations shown on the block diagram involving further energy
transfers are scheduled below:
1. Intercooler (E101)
2. Feed heater (E101A)
3. Reformed gas boiler (E102)
4. Converted gas boilers A and B (E103A,B)
5. Quench cooler (M101)
6. High-pressure boiler feed water (BFW) heater (E104)
7. CO2 reboiler (E105)
8. Interchanger (E106)
9. Low-pressure BFW heater (E107)
10. Final cooler (E108)
As has been stated, the calculations of the required mass and energy balances will be
laid out in sections that follow the block diagram.
Obviously, such a statement relates to the mass of each species entering and leaving.
Thus, the simplest statement is to recognise that all atoms of an element entering a
defined system must balance the number of atoms of the same element leaving the
system. This balance can be written for all elements entering and leaving the system.
We also recognise that if, for a particular species, e.g. a molecule, we express its mass
in moles, then we know that the moles have been calculated relative to the species
molar mass; thus, whilst in the input/output balance the masses will balance, the
Reformers and Stream Energy Interchange 9
1.5 DESIGN BASIS
The simplest design method is to use a basis of 100 kmol feedstock and work through
the process step by step to determine the amount of product. For a specified flow
rate of nitrogen/hydrogen product, everything can be scaled accordingly. In these
calculations, everything is related to 100 kmol of feedstock. In addressing the energy
balance, the first problem that will arise is specification of the feedstock compressor
and pumps since the efficiency depends on the flow rate.
Estimates of pressure drop will not separate pressure drops in equipment from
pressure drop in pipework. At the level that these calculations are presented, this
would not be practical.
Part of the process requires the removal of carbon dioxide after the water–gas
shift conversion. This chapter section will not be considered in detail. It will be
assumed that the process operates with CO2 removal by absorption. It is recognised
that other methods are available, e.g. pressure swing adsorption, but for mass balanc-
ing purposes, we will use the absorption technique. This can depend on licensor’s
data, and assumptions will be made. The solution from the absorption may well be
circulated by pumps driven by steam turbines; these are ignored in the steam system.
It will be implicitly assumed that the pumps are electrically driven.
Heat losses will be allowed in the energy balances, usually on a sensible, arbitrary
basis, since neither the actual capacity of the plant nor the detailed equipment design
is known in the calculations laid out.
As said, the assumption is that the feedstock is 100 kmol of a natural gas. The
compositions of such a gas can vary considerably, but the assumption in this work
10 Mass and Energy Balancing
will be made that the source is still similar to North Sea gas and the composition
taken is broadly representative of this feedstock (Table 1.1):
TABLE 1.1
Composition of Undesulphurised and Desulphurised
Natural Gas
Undesulphurised Component %v/v
CH4 94.68
C2H4 3.00
C3H8 0.50
C4H10 0.40
N2 1.40
H2S 0.02
Supply Conditions
includes a brief description of the compression and does energy balances illustrating
a number of approaches referenced in Appendix 3.
REQUIRED
Component mass balancing.
Understanding and calculation of equilibrium constants.
Calculation of partial pressures using Dalton’s law.
Definition and manipulation of reactant fractional conversions.
Setting conditions for endo- and exothermic reactions.
Calculation of fuel and energy requirements for an overall endothermic
reactor.
Solving for fractional conversions by solving simultaneous, non-linear
equations.
A nickel catalyst is employed for the reforming, and under the conditions described,
the water–gas shift reaction is considered to reach thermodynamic equilibrium. For
the methane–steam reaction, in doing this mass balance, an approach to equilibrium
is used (see Appendix 5). In this case, for the conditions and catalyst used, a 30°C
approach to equilibrium is assumed. Such information will be used in calculating on
the basis of known thermodynamic data.
The approach to equilibrium is an attempt to quantify the performance of the
catalyst being used in a reactor. The actual exit product compositions can be used to
calculate an equilibrium constant, and hence, using appropriate ln KP vs 1/T data,
an equivalent value of T, the temperature, can be calculated. The difference between
this value and the actual product temperature assumed is called the approach to
equilibrium. Thus, for the methane–steam reaction, we will assume an approach
to equilibrium of 30°C in an attempt to represent a realistic figure related to the
catalyst used. The actual operation will be illustrated in the mass balance on the
primary reformer.
The appropriate data are, at least, the equilibrium constants for Reactions 1 and 2,
and these are set out in Appendix 1 and their genesis given.
For the gaseous reactions discussed, we could write:
pCO pH32
K P1 =
pCH4 pH2O
pCO2 pH2
K P2 =
pCO pH2O
The equilibrium constant, as we should know, arises directly from the application of
the second law of thermodynamics to the achieving of equilibrium for the chemical
reaction. In this text, we write the equilibrium constant, KP, in terms of partial pres-
sures. We are assuming ideality in the gaseous phase. The partial pressures, pi, are
12 Mass and Energy Balancing
easily understood as the proportion of the total pressure exerted by any component
i. The proportionality is expressed through the value of yi, i.e. pi = yi P, where yi is a
component i mole fraction in the gaseous phase and P is the total system pressure.
In the published data on the K values for these reactions, there is usually agree-
ment on the K values for the water–gas shift reaction, but there is some variation on
the K values for the methane–steam reaction.
The values used in this work are calculated from the equations given below that
give the K values as a function of absolute temperature T. Data from other sources
quoted in Appendix 1 have similar ln KP vs 1/T equations fitted; none of the equations
fitted show a particular advantage over the others, so the stated equations, which had
been used for many of the calculations, were employed.
1.6.1 Methane–Steam Reaction
27,106
lnK P1 = 30.42 −
T
4160
lnK P2 = −3.798 +
T
There is a problem in steam reforming that, under the conditions of the reform-
ing, carbon can also be formed by various reactions; this carbon formation can lead
to deposition on the catalyst surface and reaction is inhibited at the surface. These
reactions are usually supressed by increasing the steam/methane ratio above the stoi-
chiometric ratio of 1. An inspection of the stoichiometric equation for the methane–
steam reaction should lead to an appreciation that excess steam, i.e. steam above the
required stoichiometric amount, will act as a diluent and increase the total moles at
equilibrium. In terms of the calculation of reactant and product moles at equilibrium,
this should increase the conversion of methane in this reaction. In the calculations
outlined here, the steam/carbon ratio is taken as 3.5/1. In many modern plants operat-
ing under different conditions, this can be operated at a lower ratio.
In this text, we will use a fractional conversion, z, to quantify the amount of a
given reactant that has been converted. Thus, the fractional conversion will be:
is applied to the figures quoted. Being independent of path, it is assumed that the
reactions can be considered in series, i.e. Reaction 1 followed by Reaction 2. This
produces the moles quoted in Table 1.3:
A simple atom balance on the feedstock gives the following:
TABLE 1.2
Atom Balances on Some Natural Gas Components
Component %v/v C H
CH4 94.70 94.7 94.7 × 4 = 378.8
C2H6 3.00 6.0 3 × 6 = 18
C3H8 0.50 1.5 0.5 × 8 = 4
C4H10 0.40 1.6 0.4 × 10 = 4
N2 1.40 0.0 0
Total 103.8 404.8
TABLE 1.3
Primary Reformer Analysis using Component Fractional Conversions
Initial Moles Moles after Reaction 1 Moles after Reaction 2
Component (kmol) (kmol) (kmol)
CH4 94.68 98.6 − 98.6zCH4 98.6 − 98.6zCH4
C2H4 3.00
C3H8 0.50
C4H10 0.40
N2 1.40 1.4 1.4
CO 98.6zCH4 98.6zCH4 − 98.6zCH4zCO
H2O 363.3 363.3 − 98.6zCH4 363.3 − 98.6zCH4 − 98.6zCH4zCO
CO2 98.6zCH4zCO
H2 3 × 98.6zCH4 3 × 98.6zCH4 + 98.6zCH4zCO
TOTAL 463.3 + 2 × 98.6zCH4
Thus, for a steam/carbon ratio of 3.5, we will take the steam flow rate as 103.8 × 3.5 = 363.3 kmol.
The generated values of zCH4 and zCO are 0.781 and 0.461, respectively. Thus, the
output values from the primary reformer can then be fixed.
Using the conversions calculated, we can use the analysis in Table 1.3 to produce
the output values given in Table 1.4.
TABLE 1.4
Results for Output from the Primary Converter Analysis
Initial Moles Moles after Reaction 1 Moles after Reaction 2
Component (kmol) (kmol) (kmol)
CH4 94.68 21.6 21.6
N2 1.40 1.4 1.4
CO 77.01 41.51
H2O 363.3 286.6 251.1
CO2 35.5
H2 231.02 266.5
Total 617.62
It will be noted that a significant quantity of hydrogen has been produced, but
correspondingly, there is significant unreacted methane, an unwelcome quantity of
carbon monoxide and also steam that will have to be considered within the stoichi-
ometry of subsequent processing.
given by Graham M Hampson (Private Communication) were used and the figures
calculated are used as the basis in subsequent calculations. The relevant polynomial
constants are given in Table A3.3 in Appendix 3.
To carry out a calculation for the relevant enthalpies, the input and output opera-
tional temperatures are required. Normally, if these are not quoted in the text, the
values fixed can be found on the block diagram.
Once the appropriate values of enthalpy were available from the Hampson poly-
nomials, the simple energy balance, based on the assumption that the reactor could
be considered adiabatic, was carried out (Table 1.5):
TABLE 1.5
Primary Reformer Energy Balance
Input Specific Output Specific
Input Output Enthalpy Input Enthalpy Output
(kmol) (kmol) (kJ/kmol) Enthalpy (kJ) (kJ/kmol) Enthalpy (kJ)
CH4 94.68 21.6 −3.9715 + 04 −3.7602E + 06 −9.4027E + 03 −2.0310E + 05
C2H6 3 0.000 −4.4367 + 04 1.331E + 05 – –
C3H8 0.5 0.000 −4.8422 + 04 −2.4211E + 04 – –
C4H10 0.4 0.000 −5.1880 + 04 −2.0752E + 04 – –
N2 1.4 1.400 −2.0738 + 04 2.9033E + 04 3.3563E + 04 4.6988E + 04
H2O 363.3 251.1 −2.1600E + 05 −7.8473E + 07 −1.9799 E + 05 −4.9715E + 07
CO 0 41.51 – – −7.9390E + 04 −3.2955E + 06
H2 0 266.52 – – 3.4231E04 9.1233e + 06
CO2 0 35.5 – – −3.5034E05 −1.2437E + 07
Output–input 2.5792E + 07
For the purposes of calculating the fuel requirements, the value of 2.5792E + 07 kJ
will be used.
CH 4 + 2O 2 → CO 2 + 2H 2O
C2H 6 + 7 2 O 2 → 2CO 2 + 3H 2O
C3H8 + 5O 2 → 3CO 2 + 4H 2O
C4 H10 + 13 2 O 2 → 4CO 2 + 5H 2O
H 2S + 3 2 O 2 → H 2O + SO 2
In analysing the combustion of the natural gas as an energy source, there are a num-
ber of calculated figures quoted. These figures are not difficult to obtain in them-
selves, but it is vital that each number is checked and understood so that the mass and
energy balances make sense. The learning process depends on the reader doing this;
otherwise, the labour involved in the calculations is wasted.
We can draw up the following table (Table 1.6) based on 100 kmol of natural gas
and the stoichiometry of the combustion equations based on 100% burning:
TABLE 1.6
Product Amounts from Combustion of Natural Gas assuming
100% Conversion
kmol in O2 CO2 SO2
Species Gas Required Produced H2O Produced Produced
CH 4 94.68 189.36 94.68 94.68 × 2 = 189.36
C2 H 6 3.0 10.50 2 × 3 = 6.00 3 × 3 = 9.00
C3 H 8 0.5 2.50 3 × 0.5 = 1.50 4 × 0.5 = 2.00
C4 H10 0.4 2.60 4 × 0.4 = 1.60 5 × 0.4 = 2.00
H 2S 0.02 0.03 0.02 0.02
N2 1.4 0
Total 100.00 205.0 103.78 202.38 0.02
If we supply the air at 5% excess, then the extra amount of O2 supplied will be:
Taking the composition of the air as 21% v/v O2 and 79% v/v N2, we can calculate
the amount of N2 as:
235.75 × 79/21 = 886.9 kmol + 1.4 kmol (from the fuel).
1.10.1 Input Enthalpy
TABLE 1.7
Input Enthalpies of Entering Natural Gas and Air Streams
Enthalpy of Fuel
Species kmol in Gas kJ/kmol at 25°C kJ
CH 4 94.68 −5.6719E + 04 −5.3702E + 06
C2 H 6 3.0 −5.6296E + 04 −1.6889E + 05
C3 H 8 0.5 −6.7061E + 04 −3.3531E + 04
C4 H10 0.4 −9.0014E + 04 −3.6006E + 04
H 2S 0.02 −1.9960E + 04 −3.9900E + 02
N2 1.4 8.6460E + 03 1.2104E + 04
Total 100.00 −5.5969E + 06
Check out the enthalpy values using the polynomial constant values used in Appendix
3. This will be required in the subsequent energy balances to be carried out. Although
these calculations and results are not carried out here, it might be a useful exercise
to check out enthalpy values (Table 1.7) from other sources as given in Appendix 3.
Thus, the total enthalpy of the input streams is:
1.10.2 Output Enthalpy
The amounts of CO2 and H2O produced have all been previously calculated assum-
ing complete combustion.
The amount of N2 associated with the input oxygen can also be calculated as:
(See Table 1.7; we note that1.4 kmol has come through from the fuel)
Taking the flue gas components, we can attempt to calculate an enthalpy figure for
the output gas (Table 1.8):
TABLE 1.8
Calculated Enthalpies for Output Gases from the Combustion
Species kmol in Gas kJ/kmol at 950°C kJ
CO2 103.78 −3.3754E + 05 −3.5030E + 07
H2O 202.38 1.9649E + 05 −3.9766E + 07
N2 886.87 + 1.4 = 888.27 3.6459E + 04 3.2385E + 07
O2 30.75 3.9190E + 04 1.2051E + 06
Total 1225.18 −4.1205E + 07
Assume that, with heat losses, the fuel requirement will be:
The input and output temperature and pressure stream conditions were taken as
follows:
Input
Temperature = 400°C
Pressure = 32.06 bar
Output
Temperature = 850°C
Pressure = 27.25 bar.
A reasonable pressure drop for such a reformer is assumed. This can be confirmed
with a suitable design-based calculation.
METHANE–STEAM REACTION
27,106
lnK P1 = 30.42 −
T
20 Mass and Energy Balancing
4160
lnK P2 = −3.798 +
T
The fractional conversions for the two reactions were calculated as 0.781 and
0.461, respectively. The output components kmol are summarised in Table 1.4.
The required energy input to the reformer has been calculated as:
2.5792E + 07 kJ.
The same natural gas as for the feedstock is burned to supply this energy. The
energy balance indicates that 52.0 kmol of gas must be burned to supply the
required energy.
REQUIRED
Primary reformer input to secondary reformer.
Information on water–gas shift reaction K values:
H 2 N 2 = 3 :1.
Fixing the output component masses by solving component mass balances and
equilibrium equations simultaneously.
Confirming the energy input and output streams balance.
Understanding some of the implications of the secondary reformer outputs
for further operations in the block diagram.
by using the energy released in the exothermic reaction that involves the combustion
of some of the methane in the stream from the primary reformer. The oxygen neces-
sary for the combustion is supplied by injection of combustion air. The amount of air
injected has to be fixed by the requirement that the amount of nitrogen entering the
ammonia synthesis loop is in the ratio N2/H2 = 1:3. The secondary reformer is obvi-
ously supplying the bulk of the necessary nitrogen for the syngas.
In the environment of the secondary reformer, it is obvious that as well as the
combustion reaction with oxygen, there could be further reforming of the methane
by the methane–steam reaction and the possibility of the water–gas shift reaction
occurring. It is important in carrying out a mass balance that these possibilities are
recognised and the appropriate calculations carried out.
The input to the secondary reformer has been fixed by the calculations carried out
on the primary reformer. To fix the oxygen and hence combustion air requirements,
certain assumptions can be made.
We assume that certain important components can be represented by the following:
Component Symbol
Carbon monoxide, CO M
Carbon dioxide, CO2 D
Nitrogen, N2 N
Hydrogen, H2 H
Water (steam), H2O W
We can assume that the water–gas shift reaction takes place in the secondary
reformer:
CO + H 2O → CO 2 + H 2
At the end of the synthesis gas production, we try to process to ensure that the CO
content is negligible; the implication of this is that at the exit from the secondary
reformer, the ratio ( CO + H 2 ) N 2 has to be 3/1. This enables us to set up balances
that can be used to calculate the amounts of each component in the secondary
reformer exit.
If we take the exit component amounts from the primary reformer, we can then
write the following atom/molecule balances and other relevant equations (Table 1.9):
In terms of our unknowns, we have indicated five; we need to identify five inde-
pendent equations to obtain quantities for each of these components at the reformer
exit. Thus:
TABLE 1.9
Atom Balance for Components from the Primary Reformer
Symbol and
Component kmol C O H Quantity at Exit
CH4 21.0 21.0 0 84.0
CO 41.51 41.51 41.51 0 M
CO2 35.30 35.30 70.60 0 D
H2 266.52 0 0 533.04 H
H2O 251.1 0 251.1 502.2 W
N2 1.4 0 0 0 N
21 21
Air (O2) N 0 X 2 N 0
79 79
Total 97.81 363.21 + 0.532N 1119.24
M + H = 3N (1.4)
The fifth equation in these calculations is provided by considering the KP value for
the water–gas shift reaction. As stated previously, the secondary reformer normally
operates at a temperature 150°C–250°C in excess of the exit temperature from the
primary reformer. In this case, we will assume a temperature of 1080°C for the exit
temperature. We can write KP as:
H×D
KP =
M×W
Because we have fixed an exit temperature, we can compute a value of KP from the
equation previously quoted for this reaction:
where T is in K.
At a temperature of 1080°C = 1353 K, we can compute a value of KP as:
H×D
KP = = 0.485 (1.5)
M×W
These five equations can be solved simultaneously (Appendix 4) and a value for M
obtained.
A brief description of a method using SOLVER in an Excel spreadsheet is described
in Appendix 4. A classic substitution/elimination approach is also illustrated.
Reformers and Stream Energy Interchange 23
We find the value of M to be 64.41 kmol. Thus, we know the amount of CO exiting
the secondary reformer.
The amounts of the other components exiting the reformer can then be calculated
using the appropriate equations for H, M, W and D, and this is also illustrated in
Appendix 4.
We can make an estimate of the CH4 that has slipped through the process. If
we assume that the methane–steam reaction can also take place in the reformer at
1080°C, then we can calculate the KP value for the reaction at this temperature from
the equation used previously:
1.12.2 Methane–Steam Reaction
27,106
lnK P = 30.42 −
T
With a 30°C approach for this reaction, we take the temperature as
1050°C = 1050 + 273 = 1323 K.
We can calculate a value of KP as 20572.6.
We are operating at a pressure of 25.77 bar.
We write for KP:
2
nCO (nH2 )3 P
KP =
nCH4 nH2O
∑n
where n is the number of moles for each component and ∑n is the total moles in
the equilibrium mixture. If we consider the total kmol of all the components except
methane in the reformer exit stream, we get a value of 769.82 kmols. We can then
express the number of kmols of CH4 in the outlet as nCH4 and the total moles in the
outlet as 769.82 + nCH4. If we then write an equation:
2
64.41 × 269.173 25.77
20,572.6 =
nCH4 × 290.45 nCH4 + 769.82
TABLE 1.10
Summary of Mass Balance on the Secondary Reformer
Component kmol at Input kmol at Exit
CH4 21.0 0.24
CO 41.51 64.41
CO2 35.30 33.2
H2 266.52 269.17
H2O 251.1 290.45
N2 1.4 111.19 + 1.4 = 112.59
21
Air (O2) N = 29.5
79
Total 770.06
Thus, the enthalpy of the air entering can be calculated by finding enthalpies for N2
and O2 using the polynomial data from Table A3.5. At 573 K, the enthalpy of N2 is
1.705E + 04 kJ and O2 is 1.7069E + 04 kJ.
The enthalpy for the air stream is:
The output data can be calculated using the data from Table A3.5 for an output
temperature of 1080°C.
TABLE 1.11
Enthalpy Data for Output Stream from the Secondary Reformer
Component kmol at Exit Enthalpy (kJ/kmol) Enthalpy (kJ)
CH4 0.24 8.9690E + 03 2.1526E + 03
CO 64.41 −6.8424E + 04 −4.4072E + 06
CO2 33.2 −3.3024E + 05 −1.0964E + 07
H2 269.17 4.2055E + 04 1.1320E + 07
H2O 290.45 −1.9165E + 05 −5.5665E + 07
N2 111.19 + 1.4 = 112.59 4.0152E + 04 4.5207E + 06
Total 770.06 −5.5193E + 07
Energy Input to the Secondary Reformer = Energy Out from the Secondary Reformer
Reformers and Stream Energy Interchange 25
In this case, we are essentially assuming an adiabatic operation with no heat leaks
assumed.
If we examine our figures:
−5.4317E + 07 kJ = −5.5193E + 07 kJ
CO + H 2O ↔ CO 2 + H 2 ∆H = −41kJ
Essentially, the action we require, the removal of CO, is being carried out with the
production of further hydrogen. (We tried to adjust for this in our N2/H2 ratio calcula-
tions in the secondary reformer.) The removal of the CO2 has to then be considered.
It can be seen from the reaction enthalpy figure quoted with the stoichiomet-
ric equation that the reaction is exothermic. Thus, we face the classic problem.
Thermodynamically, we should reduce the temperature to increase the thermody-
namic conversion of the CO. Kinetically, we should increase the temperature so that
the final thermodynamic equilibrium is approached quickly. The necessary com-
promise is achieved by using appropriate catalysts and arranging the temperature
conditions for the reaction. In this example, a high-temperature shift converter and a
low-temperature shift converter are used. As described previously, the shift convert-
ers use the water–gas shift reaction to remove carbon monoxide before the synthesis
gas enters the ammonia production reactor. The water–gas shift reaction is exother-
mic and thermodynamically should be carried out at a low temperature. In the cur-
rent set-up, this is carried out in a two-stage high-temperature shift reactor followed
26 Mass and Energy Balancing
−5.4317E + 07 kJ = −5.5193E + 07 kJ
REQUIRED
Calculation of the energy available from the secondary reformer output stream
to raise stream in a boiler.
The energy available requires a calculation of the input and output enthal-
pies based on input and output temperatures.
The difference between the input and output energies indicates the amount
of energy available for stream raising.
Enthalpies of vaporisation (latent heat) taken from appropriate thermodynamic
tables (steam tables) allow calculation of the amount of steam that can be raised.
The input stream to this boiler is the exit stream from the secondary reformer. The
component quantities, stream enthalpies, temperature and pressure for this stream
are known and shown in Table 1.12.
Temperature: 1080°C
Pressure: 26.1 bar
Total kmol: 770.06
Stream enthalpy: −5.5193E + 07 kJ
The energy available in this stream is used in the boiler to raise high-pressure
steam. The pressure of this steam is plant determined and in this case is generated
at a pressure of 62 bar. The exit temperature of the gas stream is determined by the
temperature required in the interchanger (Item E106). There is a stream from the
interchanger to the methanator (Item V107); the operation of this is discussed later
in these calculations. The suggested operating temperature range for this operation
is 250°C–350°C. The temperature finally used for calculation in this operation is
326°C at a pressure of 21.8 bar. The calculations carried out fix a temperature for the
input gas stream to the interchanger (the output stream from the reformed gas boiler
(E102)); this temperature is 535°C and this is used in the calculations related to the
reformed gas boiler (E102). We again look at Table 1.12.
TABLE 1.12
Input Mass to Reformed Gas Boiler and Component and Stream Enthalpies
Component kmol at Exit Enthalpy (kJ/kmol) Enthalpy (kJ)
CH4 0.24 8.9690E + 03 2.1526E + 03
CO 64.41 −6.8424E + 04 −4.4072E + 06
CO2 33.2 −3.3024E + 05 −1.0964E + 07
H2 269.17 4.2055E + 04 1.1320E + 07
H2O 290.45 −1.9165E + 05 −5.5665E + 07
N2 111.19 + 1.4 = 112.59 4.0152E + 04 4.5207E + 06
Total 770.06 −5.5193E + 07
28 Mass and Energy Balancing
We have the data on the input stream, and we now need to fix the enthalpy of the
output stream; obviously, there is no change in the mass balance.
We can use the constants given in Table A3.3 (Appendix 3) in the polynomial
equation to find component enthalpies at 535°C.
TABLE 1.13
Output Mass from Reformed Gas Boiler and Component and
Stream Enthalpies
Component kmol at Exit Enthalpy (kJ/kmol) Enthalpy (kJ)
CH4 0.24 −3.1586E + 04 −7.5810E + 03
CO 64.41 −8.9134E + 04 −5.7411E + 06
CO2 33.2 −3.6098E + 05 −1.1984E + 07
H2 269.17 2.4007E + 04 6.4620E + 06
H2O 290.45 −2.1170E + 05 −6.1487E + 07
N2 111.19 + 1.4 = 112.59 2.4193E + 04 2.7239E + 06
Total 770.06 −7.0035E + 07
This is an energy output available to the BFW. If we factor in a small heat loss to the
atmosphere – say 1% – then the energy available is:
We have stated the steam down pressure as 62 bar. Essentially, we are using the gas
stream energy to supply latent heat (enthalpy of vaporisation) to the water.
From steam tables:
1.4594E + 07
= 9436 kg
1557.2
Reformers and Stream Energy Interchange 29
REFERENCE
1. Richard M Felder, Ronald W Rousseau, December 2004, Elementary Principles of
Chemical Processes, Wiley Inc, USA.
1. ‘Steam Tables’ G F C Rogers, Y R Mayhew, Thermodynamic and Transport Properties
of Fluids, 5th Edition, Blackwell Publishing, Oxford, UK.
1. Arthur L Kohl, Richard Nielson, 1997, Gas Purification, Elsevier, Netherlands.
1. API Research Project 44, Carnegie Institute of Technology, 1953, Selected
Values of Physical and Thermodynamic Properties of Hydrocarbons
and Related Compounds, White Papers: University of Nebraska-Lincoln
Libraries, Pittsburgh.
2. The values are reproduced by A M Mearns, 1973, Chemical Engineering
Process Analysis, Oliver and Boyd, Edinburgh, UK, p238.
3. B Ruscic, 2015, Active Thermochemical Tables (ATcT) values based on ver.
1.118 of the Thermochemical Network; available at ATcT.anl.gov).
4. Edward V Thompson, William H Ceckler, 1978, Introduction to Chemical
Engineering, McGraw-Hill International Edition, New York, USA.
5. W C Gardiner, Ed, 1984, Thermochemical Data for Combustion Calculations,
Chapter 8 of Combustion Chemistry, Springer-Verlag, New York, USA.