Zhang 2018
Zhang 2018
Zhang 2018
https://doi.org/10.1007/s10409-018-0782-z
RESEARCH PAPER
Abstract
The dynamic stall problem for blades is related to the general performance of wind turbines, where a varying flow field is
introduced with a rapid change of the effective angle of attack (AOA). The objective of this work is to study the aerodynamic
performance of a sinusoidally oscillating NACA0012 airfoil. The coupled k−ω Menter’s shear stress transport (SST) tur-
bulence model and γ −Reθ transition model were used for turbulence closure. Lagrangian coherent structures (LCS) were
utilized to analyze the dynamic behavior of the flow structures. The computational results were supported by the experiments.
The results indicated that this numerical method can well describe the dynamic stall process. For the case with reduced
frequency K = 0.1, the lift and drag coefficients increase constantly with increasing angle prior to dynamic stall. When the
AOA reaches the stall angle, the lift and drag coefficients decline suddenly due to the interplay between the first leading- and
trailing-edge vortex. With further increase of the AOA, both the lift and drag coefficients experience a secondary rise and
fall process because of formation and shedding of the secondary vortex. The results also reveal that the dynamic behavior of
the flow structures can be effectively identified using the finite-time Lyapunov exponent (FTLE) field. The influence of the
reduced frequency on the flow structures and energy extraction efficiency in the dynamic stall process is further discussed.
When the reduced frequency increases, the dynamic stall is delayed and the total energy extraction efficiency is enhanced.
With K = 0.05, the amplitude of the dynamic coefficients fluctuates more significantly in the poststall process than in the
case of K = 0.1.
Keywords Oscillating foil · Dynamic stall · Lagrangian coherent structures · Computational fluid dynamics (CFD)
123
M. Zhang, et al.
is of significant importance to understand the dynamic inter- ing motion had a negative effect on the harvesting effi-
actions between transient wing motion and unsteady flow ciency.
structures. Moreover, as one of the primary flow energy con- Although much work has been carried out on the aerody-
version technologies, improved understanding of dynamic namic performance of oscillating wind turbine blades [32–
stall characteristics will help development of oscillating wing 36], significant uncertainty still exists regarding the influence
systems to enhance the energy harvesting performance of of unsteady flow features. The effect of the dynamic stall
wind or hydro energy. phenomenon on energy harvesting still cannot be explained
Many experimental and numerical studies have been car- clearly, and the influence of the reduced frequency on the flow
ried out on the aerodynamic performance of oscillating foils evolution and energy harvester efficiency still requires fur-
[12–17]. Ferreira et al. [18] used particle image velocime- ther investigation. The objective of the work presented herein
try (PIV) to visualize the transient flow in the operational is to analyze the flow vortex structures around an oscillating
regime. The results illustrated that the flow pattern is depen- foil using a Lagrangian-based numerical method, and provide
dent not only on the magnitude of the angle of attack (AOA) further insight into the interplay between the unsteady flow,
but also on the transportation and interaction of the shedding oscillatory motion of the foil, and aerodynamic performance.
vorticity. Wernert et al. [19] used laser-sheet visualization The numerical models are described in Sect. 2, followed by a
and PIV to investigate the unsteady flow around a pitching summary of the numerical setup in Sect. 3. In Sect. 4, detailed
airfoil. They found that the dynamic stall process can be analysis of the aerodynamic performance and flow structures
categorized into four stages: (1) attached flow, (2) develop- is presented. Finally, conclusions are drawn in Sect. 5.
ment of the leading-edge vortex (LEV), (3) poststall vortex
shedding, (4) flow reattachment. Lee and Gerontakos [20]
applied smoke flow visualization and hot-film sensors to
investigate the transient flow structures and dynamic stall
2 Numerical model
characteristics of an oscillating NACA0012 airfoil. They
illustrated that the boundary-layer transition and dynamic
2.1 Basic governing equations
stall point were delayed with increase of the reduced fre-
The flow field was simulated by solving the unsteady
quency. This conclusion has also been proved experimentally
Reynolds-averaged Navier–Stokes (URANS) equations, with
by Carr [21] and Ekaterinaris and Platzer [22]. Simpson
the continuity and momentum equations listed below:
et al. [23] experimentally studied the unsteady flow around
a sinusoidal heaving and pitching foil, focusing on the influ-
ence of the maximum AOA, Strouhal number, and aspect ∂u j
= 0, (1)
ratio on the energy extraction. The results revealed hydro- ∂x j
dynamic efficiency of 43% ± 3% with maximum AOA
∂u i ∂(u i u j ) ∂p ∂ ∂u i
of 34.37◦ , Strouhal number of 0.4, and aspect ratio of ρ + =− + μ , (2)
∂t ∂x j ∂ xi ∂x j ∂x j
7.9.
With the development of computing equipment and tech-
niques, much attention has been paid to computational fluid where ρ is the fluid density (all flow conditions in this study
dynamics to better investigate dynamic stall characteristics being incompressible), t is time, u is the velocity, x is the
[24–28]. Hang et al. [4] numerically investigated the per- coordinate, p is the pressure, μ is the fluid viscosity, and
formance of offshore floating vertical-axis wind turbines subscripts i and j denote the directions of the Cartesian coor-
subjected to pitch motion. The results indicated that the dinates.
power output of the turbines and the range of aerody-
namic force variations were enlarged. The transient flow 2.2 Turbulence model
of an oscillating hydrofoil at different pitching rates was
studied numerically by Huang et al. [7,29]. The results The simulation solved the URANS equations by applying the
revealed that the pitching velocity had an important effect revised k–ω SST turbulence model, which couples the k–ω
on the hydrodynamic characteristics. The energy harvest- shear stress transport (SST) turbulence model [37] and the
ing performance of a fully activated flapping foil under γ −Reθ transition model [38–40]:
wind gust conditions was studied by Chen et al. [30].
Compared with the uniform flow condition, the energy
harvesting efficiency was higher. Energy extraction effi- ∂ (ρk) ∂ ρu j k
+ = P̃k − D̃k
ciency with pitching motion was also investigated by Teng ∂t ∂x j
et al. [31]; the results showed that nonsinusoidal pitch- ∂ ∂k
+ (μ + σk μt ) , (3)
∂x j ∂x j
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
+ ( ∂ x jj ) = Cω Pω − βω ρω2
∂(ρω) ∂ ρU ω
∂t
+ ∂∂xi μ + μ t ∂ω
σk ∂ xi (4)
+ 2ρ (1 − F1 ) σω2 ω1 ∂∂kxi ∂ω
∂ xi ,
123
M. Zhang, et al.
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
1.25ea21 the mean total power extracted to the total power available
21
GC Ifine = p . (15)
r21 − 1 in the oncoming flow passing through the swept area.
To quantify the energy extraction performance of an oscil-
The averaged lift coefficient is considered to be an impor- lating foil system, the nondimensional power coefficient
tant parameter for analysis of dynamic performance and was CPower and the energy extraction efficiency η are defined as
therefore chosen as the main parameter for uncertainty analy-
sis. The relative parameters were calculated and are presented PPower
CPower = , (16)
2 ρU∞ sc
1 3
in Table 1. Based on the error analysis results, the value of the
GCI for the average lift coefficient was found to be 0.67 % PPower = M (t) wp (t) , (17)
and 3.52%. As both uncertainty estimators lie in a reason-
PPower
able range, the medium grid (2.2 × 105 ) was selected for all η= , (18)
simulations in the present work.
1
2 ρU 3
∞ sd
Additionally, numerical results obtained using different 1 T
PPower = PPower (t)dt, (19)
time step sizes are presented in Fig. 4. Compared with the T 0
experimental data, the lift coefficient predicted with t =
1 × 10−3 s cannot reflect the transient lift evolution in the where PPower and PPower are the instantaneous and time-
downstroke, while when the time step was chosen as t = averaged power, respectively, extracted from the oncoming
1 × 10−4 s or t = 1 × 10−5 s, the predicted lift coefficient flow from the pitching contribution, M(t) is the torque about
remained almost the same. Hence, a time step of t = 1 × the pitching center, wp (t) is the transient angular velocity,
10−4 s was chosen for the computations, ensuring a Courant d is the vertical extent of the oscillating foil, and T is the
number of CFL x = U∞ × t/x ≈ 1 in the streamwise period of one pitching cycle.
direction and CFL y = V∞ × t/y ≈ 1 in the y-direction.
Improved understanding of dynamic stall characteristics will 4.1 Analysis of transient aerodynamic load and flow
help development of oscillating wing systems, which repre- features
sent one of the primary flow energy conversion technologies,
to enhance the energy harvesting performance for wind or Figure 5 shows the evolution of the lift coefficient Cl and drag
hydro energy. Hence, the instantaneous power coefficient coefficient Cd (Cd = D/(0.5ρU∞ 2 sc), where D is the drag)
CPower and the total energy extraction efficiency η were eval- with the AOA. The dynamic experimental results shown in
uated to quantify the energy extraction performance of the Fig. 5 were obtained from Ref. [20]. It is shown that, in the
oscillating foil system, also representing a valuable refer- upstroke from t1 (α+ = −5◦ ) to t7 (α+ = 25◦ ), the numeri-
ence for wind turbine blade designs. The instantaneous power cal results agree well with the experimental data, showing a
extracted from the flow comes from the pitching contribution, maximum predicted Cl very close to the measured value. In
and the energy extraction efficiency is defined as the ratio of the downstroke, the predicted lift coefficient presents small
123
M. Zhang, et al.
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
123
M. Zhang, et al.
Fig. 8 FTLE field and corresponding LCS during LEV developments stage (K = 0.1): a t2 (α+ = 17.5◦ ), b t3 (α+ = 20.4◦ ), c t4 (α+ = 23.5◦ )
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
the boundary of leading-edge vortexes with reverse rotational 4.3 Influence of reduced frequency on flow
direction and the first TEV. This can also be seen from the structures in dynamic stall process
purple particle inside LCS C and the pink particle inside
LCS D. To further investigate the effect of the oscillation fre-
quency on the aerodynamic performance, the transient flow
(2) Dynamic stall stage (α+ = 23.5◦ − α− = 21.8◦ ) structures and corresponding aerodynamic characteristics at
different reduced oscillating frequencies are discussed in this
To provide insight into the vortex structures in this stage, as section.
shown in Fig. 9, the FTLE field and corresponding LCS at Figures 10 and 11 show the evolution of the predicted lift
t5 (α+ = 24.4◦ ), t6 (α+ = 24.8◦ ), t7 (α+ = 25◦ ), t8 (α− = coefficient (Cl ) and power coefficient (CPower ) for different
24.7◦ ), t10 (α− = 22.8◦ ), and t11 (α− = 21.8◦ ) are presented. reduced frequencies (K = 0.05 and 0.1). Table 2 presents
At t5 (α+ = 24.4◦ ), LCS D rolls up and becomes larger in the energy extraction efficiency at the different reduced fre-
size with the development of the first TEV, accompanied by quencies. It can be observed that the total energy extraction
shedding of the first LEV. Meanwhile, LCS C forms outside efficiency at K = 0.1 was larger than the case of K = 0.05.
LCS B, which results from shedding of the first LEV and the The z-vorticity contours and the instantaneous streamlines at
interaction between the reverse-directional vortex structures. typical angles of attack for both cases are shown in Fig. 12,
During the time interval TLE , the black and purple particles in where t1 −t7 represent typical times for the case with K = 0.1
Fig. 9a bypass the circulation region and follow its boundary. and t1 −t7 for the case with K = 0.05. Comparisons of
LCS B indicates the boundary of the secondary LEV. After the predicted pressure coefficients for the cases with differ-
the time interval TLE , the green particle in Fig. 9a moves ent reduced frequencies (K = 0.05 and 0.1) are shown in
backward on the surface. Fig. 13.
When it reaches t6 (α+ = 24.8◦ ), LCS C becomes weaker. Prior to dynamic stall, it is found that the evolution of the
This means that the visibility of the vortex boundaries and the predicted Cl and the power coefficient (CPower ) are similar
vortex strength become weaker. LCS D becomes larger and for the different reduced frequencies, as shown in Figs. 10
sheds away. LCS B and E also become larger and attached and 11. The corresponding Cl increases almost linearly with
to the surface with the development of the secondary LEV. increasing angle, while the corresponding CPower remains
Meanwhile, LCS F is newly formed inside LCS E, as shown at low amplitude then increases sharply. At K = 0.05, the
in Fig. 9b. At t7 (α+ = 25◦ ), LCS B covers three-quarters formation of the first LEV occurs at t1 (α+ = 16◦ ), being
of the suction surface. The track of the gray-brown particle more specific than at K = 0.1, as shown in Fig. 12a, g.
inside LCS E indicates that the vortex center moves down- The larger adverse pressure gradient distributions at α+ =
stream. The blue particle inside LCS F rolls up and attaches 16◦ for K = 0.05 are responsible for the advanced for-
at the surface due to the suction force of the secondary LEV. mation of the first LEV, as shown in Fig. 13a. The first
When the angle of attack reaches α− = 24.7◦ (t8 ), LCS B has LEV begins to develop with the low-pressure region at t2
covered the whole suction surface with low-pressure region, (α+ = 17.1◦ ), resulting in the sharp increase of Cl and
as shown in Fig. 8, which leads to the secondary increase of CPower , as shown in Figs. 10 and 11. When it reaches t3
the lift coefficient. (α+ = 19.5◦ ), the first LEV has developed and covered
As α decreases in the downstroke phase, the secondary the entire suction surface, as shown in Fig. 12b, c. As α
LEV sheds away and the secondary TEV is induced. At t10 increases, the low-pressure region moves to the trailing edge
(α− = 22.8◦ ), LCS G, which represents the boundary of the with the development of the first LEV, significantly enhanc-
secondary TEV, extrudes LCS B, as shown in Fig. 9e. The ing the lift and power prior to the dynamic stall process.
gray-brown particle is attracted to leave the foil by the suction The adverse pressure gradient distributions at the same angle
force. Because of the interaction of the secondary LEV and are larger for K = 0.05 than K = 0.1, as shown in
TEV, the blue and cyan particles are forced to attach to the Fig. 13a–c. This is because the pitching velocity is slow at
surface. This results in a decline of the lift coefficient again. K = 0.05, and the formation of the first LEV can develop
When it reaches t11 (α− = 21.8◦ ), LCSs weaken gradually. adequately.
As α decreases, LCS F constantly merges with the free stream During dynamic stall, the stall point is again more spe-
flow and sheds away on account of the interaction between cific and the maximum value of Cl and CPower becomes
the merged vortex and the third TEV, which is responsible low for K = 0.05, as shown in Figs. 10 and 11. This sig-
for the small-amplitude oscillating behavior in the lift and nificantly affects the total energy extraction efficiency of
drag coefficient curves. the oscillating foil. With increase of the angle of attack, as
shown in Fig. 13d, e, the adverse pressure gradient distri-
butions at the suction surface for K = 0.05 become low
with low variations instead of the complex pressure coef-
123
M. Zhang, et al.
a t5 +
) b t6 +
) c t7 +
)
FTLE
field
LCS
d t8 -
) e t10 -
) f t11 -
)
FTLE
field
LCS
Fig. 9 FTLE field and corresponding LCS during poststall vortex shedding stage (K = 0.1)
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
Fig. 10 Evolution of predicted Cl with different reduced frequencies (K = 0.05 and 0.1) for Re = 135,000: a full range, b zoomed view of
dynamic stall process
Fig. 11 Evolution of power coefficient (CPower ) with different reduced frequencies (K = 0.05 and 0.1) for Re = 135,000: a full range, b zoomed
view of dynamic stall process
ficient distributions at K = 0.1. At t4 (α+ = 21.7◦ ) in Table 2 Energy extraction efficiency at different reduced frequencies
the upstroke phase, the first LEV has already shed away (K = 0.05 and 0.1)
and three vortices align alongside the upper surface com- K 0.05 0.1
pared with t4 (α+ = 24.4◦ ), which is responsible for the
η (%) 32.18 55.46
decline of Cl and CPower , as shown in Figs. 10, 11, and 12d.
Compared with K = 0.1, the small-amplitude oscillating
behavior becomes more severe for the case with K = 0.05
from t5 (α+ = 23◦ ), as shown in Figs. 10 and 11. More- the dynamic stall process, the evolution of the dynamic per-
over, Cl and adverse pressure gradient distributions are lower formance further results in a decrease of the total energy
than at K = 0.1, as shown in Figs. 10 and 13g, h. At t5 extraction efficiency compared with the case of K = 0.1.
(α+ = 23◦ ), the vortex near the center clearly shrinks while When the angle of attack increases to t7 (α+ = 24.5◦ ),
another pair of vortices merge, as shown in Fig. 12e, which the third TEV is induced on the trailing edge, as shown in
is responsible for the increase of Cl again. However, com- Fig. 12f. As α declines in the downstroke phase, TEV for-
pared with the formation of the attached secondary LEV at t5 mation, interaction, and shedding are repeated many times,
(α− = 24.7◦ ), the merged vortex interacts with the secondary being responsible for the decreasing amplitude and high-
TEV at K = 0.05. It can be observed that no attached sec- frequency oscillating behavior of the flow. This is mainly
ondary LEV was induced at K = 0.05. This is responsible for attributed to the merged vortex and TEV having sufficient
the difference in the adverse pressure gradient distributions time to develop and interact for K = 0.05. This oscillat-
for different reduced frequencies, as shown in Fig. 13e–g. At ing behavior may result in the decrease of the total energy
t6 (α+ = 24◦ ), development and shedding of the secondary extraction efficiency of the oscillating foil. For K = 0.1,
TEV result in the decline of Cl and CPower 10 and 12f. In the pitching velocity is too fast for the vortex to develop and
interact completely.
123
M. Zhang, et al.
K=0.05 K=0.1
a t’1 (α+ = 16˚) h t1 (α+ = 17.5˚)
Fig. 12 Contours of z-vorticity superimposed on instantaneous streamlines at different angles of attack (K = 0.05 and 0.1)
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
Fig. 12 continued
123
M. Zhang, et al.
Fig. 13 Comparisons of predicted pressure coefficients at same geometric angles of attack for different reduced frequencies (K = 0.05 and 0.1)
123
Lagrangian-based numerical investigation of aerodynamic performance of an oscillating foil
(2) The reduced frequency significantly affects the flow 4. Hang, L., Zhou, D., Lu, J., et al.: The impact of pitch motion of
structures and energy extraction performance in the a platform on the aerodynamic performance of a floating vertical
axis wind turbine. Energy 119, 369–383 (2017)
dynamic stall process. At K = 0.1 and 0.05, the evolu- 5. Mckenna, R., Leye, P.O.V.D., Fichtner, W.: Key challenges and
tion of Cl and CPower is approximately similar prior to prospects for large wind turbines. Renew. Sustain. Energy Rev. 53,
the dynamic stall process. However, compared with the 1212–1221 (2016)
case of K = 0.1, the dynamic stall point is advanced 6. Lu, K., Xie, Y., Zhang, D., et al.: Systematic investigation of the
flow evolution and energy extraction performance of a flapping-
and no attached secondary LEV is generated when airfoil power generator. Energy 89, 138–147 (2015)
K = 0.05. Additionally, the small-amplitude oscillat- 7. Huang, B., Wu, Q., Wang, G.Y.: Numerical simulation of unsteady
ing behavior of the dynamic curve becomes more severe cavitating flows around a transient pitching hydrofoil. Sci. China
for the case of K = 0.05. These phenomena result in Technol. Sci. 57, 101–116 (2014)
8. Lee, T.: Effect of flap motion on unsteady aerodynamic loads. J.
the oscillating behavior of the power coefficient (CPower )
Aircr. 44, 334–338 (2015)
and affect the total energy extraction efficiency. The total 9. Birch, D.M., Lee, T.: Tip vortex behind a wing undergoing deep-
energy extraction efficiency is higher for K = 0.1 than stall oscillation. AIAA J. 43, 2081–2092 (2015)
K = 0.05. 10. Liu, T.T., Huang, B., Wang, G.Y., et al.: Experimental investigation
of the flow pattern for ventilated partial cavitating flows with effect
(3) LCSs defined by ridges of the FTLE field were utilized
of Froude number and gas entrainment. Ocean Eng. 129, 343–351
to investigate transient flow structures. Compared with (2017)
the Eulerian approach, e.g., using the z-vorticity (ωz ), 11. Wang, Y.W., Xu, C., Wu, X.C., et al.: Ventilated cloud cavitating
such Lagrangian-based analysis of flow structures in the flow around a blunt body close to the free surface. Phys. Rev. Fluids
2, 084303 (2017)
dynamic stall process can effectively avoid overpredic- 12. Long, X.P., Cheng, H.Y., Ji, B., et al.: Large eddy simulation and
tion of vortex structures. The dynamic behavior of the Euler–Lagrangian coupling investigation of the transient cavitating
flow structures was effectively identified using the FTLE turbulent flow around a twisted hydrofoil. Int. J. Multiph. Flow 100,
field. 41–56 (2018)
13. Wang, G.Y., Wu, Q., Huang, B.: Dynamics of cavitation–structure
interaction. Acta Mech. Sin. 33, 685–708 (2017)
14. Huang, B., Zhao, Y., Wang, G.Y.: Large eddy simulation of
In the future, the three-dimensional effect [51–54] and its turbulent vortex-cavitation interactions in transient sheet/cloud
interactions with turbulence and energy harvesting are wor- cavitating flows. Comput. Fluids 92, 113–124 (2014)
15. Choudhry, A., Arjomandi, M., Kelso, R.: Methods to control
thy of further investigation, so LCS and particle tracking dynamic stall for wind turbine applications. Renew. Energy 86,
techniques will be applied to three-dimensional flow fields 26–37 (2016)
to present more details of the transient flow structures. To 16. Hameed, M.S., Afaq, S.K.: Design and analysis of a straight bladed
determine the spatial and temporal variation of the turbu- vertical axis wind turbine blade using analytical and numerical
techniques. Ocean Eng. 57, 248–255 (2013)
lent structures more accurately, direct numerical simulations 17. Huang, B., Young, Y.L., Wang, G.Y., et al.: Combined experi-
(DNS) and large-eddy simulations (LES) will be added in mental and computational investigation of unsteady structure of
future work. In addition, the effect of the pitching amplitude sheet/cloud cavitation. J. Fluids Eng. Trans. ASME 135, 071301
on the flow evolution and energy harvesting performance will (2013)
18. Ferreira, C.S., Bussel, G.V., Kuik, G.V.: 2D CFD simulation of
also be discussed further in the future. dynamic stall on a vertical axis wind turbine: verification and vali-
dation with PIV measurements. In: 45th AIAA Aerospace Sciences
Acknowledgements This work was supported by the National Post- Meeting and Exhibit (2006)
doctoral Program for Innovative Talents (Grant BX201700126), the 19. Wernert, P., Geissler, W., Raffel, M., et al.: Experimental and
China Postdoctoral Science Foundation (Grant 2017M620043), the numerical investigations of dynamic stall on a pitching airfoil.
National Natural Science Foundation of China (Grants 51679005 and AIAA J. 34, 982–989 (1996)
91752105), and the National Natural Science Foundation of Beijing 20. Lee, T., Gerontakos, P.: Investigation of flow over an oscillating
(Grant 3172029). airfoil. J. Fluid Mech. 512, 313–341 (2004)
21. Carr, L.W.: Progress in analysis and prediction of dynamic stall. J.
Aircr. 25, 6–17 (1988)
22. Ekaterinaris, J.A., Platzer, M.F.: Computational prediction of air-
foil dynamic stall. Prog. Aerosp. Sci. 33, 759–846 (1997)
References 23. Simpson, B.J., Hover, F.S., Triantafyllou, M.S.: Experiments in
direct energy extraction through flapping foils. in: the Eighteenth
1. Melício, R., Mendes, V.M.F., Catalão, J.P.S.: Transient analysis International Offshore and Polar Engineering Conference, 2008
of variable-speed wind turbines at wind speed disturbances and a 24. Shehata, A.S., Xiao, Q., Saqr, K.M., et al.: Passive flow control for
pitch control malfunction. Appl. Energy 88, 1322–1330 (2011) aerodynamic performance enhancement of airfoil with its appli-
2. González, L.G., Figueres, E., Garcerá, G., et al.: Maximum-power- cation in wells turbine—under oscillating flow condition. Ocean
point tracking with reduced mechanical stress applied to wind- Eng. 136, 31–53 (2017)
energy-conversion-systems. Appl. Energy 87, 2304–2312 (2010) 25. Tseng, C.C., Hu, H.A.: Dynamic behaviors of the flow past a pitch-
3. Karbasian, H.R., Esfahani, J.A., Barati, E.: The power extraction ing foil based on Eulerian and Lagrangian viewpoints. AIAA J. 54,
by flapping foil hydrokinetic turbine in swing arm mode. Renew. 712–727 (2016)
Energy 88, 130–142 (2016)
123
M. Zhang, et al.
26. Ducoin, A., Astolfi, J.A., Deniset, F., et al.: Computational and 41. Haller, G., Yuan, G.: Lagrangian coherent structures and mixing in
experimental investigation of flow over a transient pitching hydro- two-dimensional turbulence. Phys. D 147, 352–370 (2000)
foil. Eur. J. Mech. B Fluids 28, 728–743 (2009) 42. Wu, Q., Huang, B., Wang, G.: Lagrangian-based investigation of
27. Bhat, S.S., Govardhan, R.N.: Stall flutter of NACA 0012 airfoil at the transient flow structures around a pitching hydrofoil. Acta
low Reynolds numbers. J. Fluids Struct. 41, 166–174 (2013) Mech. Sin. 32, 64–74 (2016)
28. Wang, S., Ingham, D.B., Ma, L., et al.: Numerical investigations 43. Tseng, C.C., Liu, P.B.: Dynamic behaviors of the turbulent cavi-
on dynamic stall of low Reynolds number flow around oscillating tating flows based on the Eulerian and Lagrangian viewpoints. Int.
airfoils. Comput. Fluids 39, 1529–1541 (2010) J. Heat Mass Transf. 102, 479–500 (2016)
29. Huang, B., Ducoin, A., Young, L.Y.: Physical and numerical inves- 44. Wang, Z.Y., Huang, B., Zhang, M.D., et al.: Experimental and
tigation of cavitating flows around a pitching hydrofoil. Phys. numerical investigation of ventilated cavitating flow structures with
Fluids 25, 102109 (2013) special emphasis on vortex shedding dynamics. Int. J. Multiph.
30. Chen, Y.L., Zhan, J.P., Wu, J., et al.: A fully-activated flapping foil Flow 98, 79–95 (2018)
in wind gust: energy harvesting performance investigation. Ocean 45. Roache, P.J.: Quantification of uncertainty in computational fluid
Eng. 138, 112–122 (2017) dynamics. Annu. Rev. Fluid Mech. 29, 123–160 (2003)
31. Teng, L., Deng, J., Pan, D., et al.: Effects of non-sinusoidal pitching 46. Gohil, P.P., Saini, R.P.: Effect of temperature, suction head and flow
motion on energy extraction performance of a semi-active flapping velocity on cavitation in a Francis turbine of small hydro power
foil. Renew. Energy 85, 810–818 (2016) plant. Energy 93, 613–624 (2015)
32. Lai, J.C.S., Platzer, M.F.: Jet characteristics of a plunging airfoil. 47. Roache, P.J.: Verification of codes and calculations. AIAA J. 36,
AIAA J. 37, 1529–1537 (2015) 696–702 (2012)
33. Gharali, K., Johnson, D.A.: Dynamic stall simulation of a pitching 48. Kwasniewski, L.: Application of grid convergence index in FE
airfoil under unsteady freestream velocity. J. Fluids Struct. 42, 228– computation. Bull. Pol. Acad. Sci. Tech. Sci. 61, 123–128 (2013)
244 (2013) 49. Kleinhans, M.G., Jagers, H.R.A., Mosselman, E., et al.: Procedure
34. Guo, Q., Zhou, L., Wang, Z.: Comparison of BEM-CFD and full of estimation and reporting of uncertainty due to discretization in
rotor geometry simulations for the performance and flow field of a CFD applications. J. Fluids Eng. 130, 078001 (2008)
marine current turbine. Renew. Energy 75, 640–648 (2015) 50. Hunt, J.C.R., Wray, A.A., Moin, P.: Eddies, stream, and conver-
35. Wu, Q., Wang, Y.N., Wang, G.Y.: Experimental investigation of gence zones in turbulent flows. Center for Turbulence Research
cavitating flow-induced vibration of hydrofoils. Ocean Eng. 144, Report CTR-S88. pp. 193–208 (1988)
50–60 (2017) 51. Choudhry, A., Leknys, R., Arjomandi, M., et al.: An insight into
36. Wu, Q., Huang, B., Wang, G.Y., et al.: Experimental and numerical the dynamic stall lift characteristics. Exp. Therm. Fluid Sci. 58,
investigation of hydroelastic response of a flexible hydrofoil in 188–208 (2014)
cavitating flow. Int. J. Multiph. Flow 74, 19–33 (2015) 52. Martinat, G., Braza, M., Hoarau, Y., et al.: Turbulence modeling of
37. Menter, F,R.: Improved two-equation k − ω turbulence models for the flow past a pitching NACA0012 airfoil at 105 and 106 Reynolds
aerodynamic flows. NASA Tech. Memo. 34, 103975 (1992) numbers. J. Fluids Struct. 24, 1294–1303 (2008)
38. Langtry, R.B., Menter, F.R., Likki, S.R., et al.: A correlation-based 53. Velkova, C., Todorov, M., Dobrev, I., et al.: Approach for numerical
transition model using local variables-part I: model formulation. J. modeling of airfoil dynamic stall. in: Proceedings of BulTrans-
Turbomach. 128, 413–422 (2006) 2012, 26–28 September, Sozopol, 1–6 (2012)
39. Langtry, R.B., Menter, F.R., Likki, S.R., et al.: A correlation-based 54. Tseng, C.C., Cheng, Y.E.: Numerical investigations of the vortex
transition model using local variables-part II: test cases and indus- interactions for a flow over a pitching foil at different stages. J.
trial applications. J. Turbomach. 128, 423–434 (2006) Fluids Struct. 58, 291–318 (2015)
40. Menter, F.R., Langtry, R.B., Völker, S.: Transition modeling for
general purpose CFD codes. Flow Turbul. Combust. 77, 277–303
(2006)
123