You are on page 1of 9

AIAA AVIATION Forum 10.2514/6.

2020-3015
June 15-19, 2020, VIRTUAL EVENT
AIAA AVIATION 2020 FORUM

Kinematics and Hydrodynamics of a Dolphin


in Forward Swimming
Pan Han1 and Junshi Wang2
University of Virginia, Charlottesville, VA, 22903, United States

Frank E. Fish3
West Chester University, West Chester, PA, 19383, United States
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

Haibo Dong4
University of Virginia, Charlottesville, VA, 22903, United States

Dorsoventral undulation is widely adopted by aquatic mammals for propulsion. As one of the
excellent swimmers, dolphins are well known for its capabilities of performing high-efficiency
cruising. In the present research, to explore the underlying mechanisms of the efficient
dorsoventral propulsion of swimming dolphins, a combined experimental and numerical process
is carried out to study the animal’s kinematics and hydrodynamics. Using the video graphic
technique and virtual skeleton-based surface reconstruction method, a three-dimensional high-
fidelity computational model is obtained for the swimming dolphin. Kinematic analysis is
performed on this computational model to achieve the time-varying kinematics of the dolphin, and
a sharp-interface immersed-boundary-method (IBM) based incompressible computational fluid
dynamic (CFD) solver is used to look into the corresponding hydrodynamics and wake structures.
The results from this work aim to bring new insight into understanding the force generation
mechanisms of dorsoventral swimming and offer potential suggestions to the future designs of
unmanned underwater vehicles.

I. Nomenclature
C = dolphin fluke chord length
BL = dolphin body length
A = fluke-tip peak-to-peak amplitude
 = body undulation wavelength
f = body undulation frequency
t = time
T = period of body undulation
U = incoming flow speed
Re = Reynolds number
St = Strouhal number
Sw = wetted surface area
FT = thrust force
CT = thrust coefficient
L = lift force

1
PhD, Department of Mechanical & Aerospace Engineering, University of Virginia, VA 22903, AIAA member.
2
PhD student, Department of Mechanical & Aerospace Engineering, University of Virginia, VA 22903, AIAA member.
3
Professor, Department of Biology, West Chester University, PA 19383.
4
Associate Professor, Department of Mechanical & Aerospace Engineering, University of Virginia, VA 22903, AIAA member.

Copyright © 2020 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
CL = lift coefficient
P = hydrodynamic power
CPW = power coefficient
p = surface pressure
Cp = pressure coefficient

II. Introduction
Among aquatic animals, dolphins are known for their outstanding swimming abilities, such as high-speed
swimming, high-efficiency cruising, porpoising, maneuverability, and rapid acceleration [1, 2]. The hydrodynamic
mechanisms behind their high-speed and high-efficiency propulsion are of particular interest to the design of high-
performance underwater vehicles [3, 4]. Since the famous “Gray’s paradox” [5] was raised in 1936, a number of
experimental studies [6-11] and numerical [12-16] studies were conducted in an attempt to explain the underlying
hydrodynamics of dolphin swimming. However, owing to technical difficulties in experimental measurements and
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

limited computational power, a clear understanding of the hydrodynamic performance and the associated flow physics
remains elusive.
Hydrodynamic force/pressure generations on the dolphin body and information about the flow field, such as
pressure, velocity, and vorticity, are key elements in understanding the flow mechanisms of dolphin swimming.
However, it is extremely difficult to obtain the force/pressure or flow field information on dolphin swimming in the
open ocean. Despite the difficulties, Fish et al. [11] first measured the flow field in the vicinity (and in the wake) of
the flukes of actively swimming dolphins in a large outdoor pool. Strong vortices were found to be generated by the
oscillations of the flukes, and thrust production was also estimated.
Compared with experimental measurements, numerical simulations can obtain much more detailed force/pressure
information relevant to the dolphin body and information of the flow field. However, the high Reynolds number (Re)
associated with dolphin swimming and the modeling of the complex body morphology and swimming kinematics are
the major challenges for numerical simulation. The Re of dolphin swimming can easily reach 107 which requires very
fine computational grids to resolve the boundary layer and the flow. Besides, dynamic morphing [17] and asymmetric
dorsoventral kinematics [1] of dolphin flukes during one tail beat cycle have made it very challenging to reconstruct
dolphin kinematics with great accuracy. Wang et al. [15] evaluated the swimming speed of a dolphin model using a
commercial Reynolds-Averaged Navier–Stokes (RANS) flow solver with posterior body motion prescribed from a
sine function. However, unlike the pitched flukes during mid-stokes in real dolphins, the sine function gave the model
a horizontal fluke during mid-strokes. On the flow solver side, the turbulence model used in RANS solver tends to
bring more numerical damping than direct numerical simulation (DNS) for transient simulations as such. Tanaka et
al. [16] measured the acceleration and deceleration of dolphin swimming and used RANS solver to calculate the
hydrodynamic drag of a geometrically realistic but static dolphin model. Thrust force and power were estimated based
on the static CFD calculation and body kinematics, rather than the direct simulation of a dynamic model with realistic
kinematics.
In this work, the underlying mechanisms of the efficient dorsoventral undulatory propulsion are explored using a
combined experimental and numerical process for a dolphin in forward swimming. With the video graphic technique
and virtual skeleton-based surface reconstruction method, a three-dimensional high-fidelity computational model is
obtained from the swimming dolphin with time-varying kinematics. The accurate morphology and kinematics
reconstruction captures important dolphin swimming features, including the dorsoventrally asymmetric kinematics
and the dynamic morphing of the fluke. A sharp-interface immersed-boundary-method (IBM) based incompressible
computational fluid dynamic (CFD) solver [18] is employed to look into the corresponding hydrodynamics and wake
structures. The results from this work aim to bring new insight into understanding the force generation mechanisms
of dorsoventral swimming and offer potential suggestions to the future designs of unmanned underwater vehicles.

III. Problem Definition and Numerical Approach

A. Video Recording and Computational Model Reconstruction


With two cameras recording at 60 frames/second, the steady forward swimming motion of an Atlantic bottlenose
dolphin (Tursiops truncatus) was recorded from two views (the perspective view and the lateral views) in an aquarium
(National Aquarium, Baltimore, MD, USA). The dolphin undulated at a frequency of 3.1 Hz and swam at a speed of
0.78 body length (BL) per stroke. Figure 1 (a) shows the time instant when the fluke tip is at its topmost position and
Figure 1 (b) shows the time instant when the fluke tip is at its bottommost position. With the obtained images, a virtual

2
skeleton-based reconstruction process is employed and the computational model is reconstructed in Maya (Autodesk,
Inc). The same technique has been successfully applied to swimming jackfish [19], and manta ray [20]. Details about
this technique can be found in Ref. [21]. In Figure 1, the reconstructed computational model is presented. There is
good agreement between the experimental images and the reconstructed computational model.
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

Figure 1. The images of the swimming dolphin at (a) t = 0.17s and (b) t = 0.33s and the reconstructed
computational model at the same time instants.
To define the swimming kinematics, two non-dimensional parameters are defined, the Strouhal number (St) and
the Reynolds number (Re):

f A U  BL
St = , Re = (1)
U 
where f is the body undulation frequency, A is the peak-to-peak amplitude of the fluke tip, U is the swimming speed,
and  is the kinematic viscosity. In this work, the dolphin was swimming at a Strouhal number of 0.35 and a Reynolds
number in the order of 1.0 107 .

B. Numerical Approach
To obtain the flow field, the incompressible Navier-Stokes equations are solved as the governing equations, written
in the indicial form as below
ui
=0 (2)
xi
ui  ( ui u j ) p 1   ui 
+ =− +   (3)
t x j xi Re x j  x j 
where the subscripts i and j define the x-, y- and z-directions, u is the velocity components and p is the pressure.
With the computational model imported as immersed moving boundaries, an inhouse finite difference-based
Cartesian-grid immersed-boundary-method based flow solver is used to compute the flow, which has been proven to
be successful in solving the flapping/undulating problems with moving boundaries [22-32]. Details of this solver can
be found in Ref. [18].
In Figure 2, the employed Cartesian grid is sketched, which has a computational domain size of
5.0BL  2.0BL  2.0BL with a non-uniform 513 165 129 (about 10.9 million points in total) grid points setup.
Around the computational model, the Cartesian grid is designed to be uniform with a higher-resolution to capture the
near-field wake structure. And further, to resolve the moving boundary and boundary layer flow, a block-wise octree-
like structured adaptive mesh refinement (SAMR) [33, 34] technique is applied in the higher-resolution uniform grid
zone. The SAMR block is denoted in Figure 2 as the blue block. With the SAMR technique, the mesh near the body
is refined to a resolution of x = 1.9 10−3 BL , which has been proven to be accurate enough for the current work
through our extensive grid independent studies. More details about this method as well as its applications in three-
dimensional bio-inspired flow simulations can be found in Ref. [33, 34].

3
Figure 2. Schematic of the computational domain and boundary condition setup.
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

To investigate the fundamental hydrodynamics of the swimming dolphin, the reconstructed computational model
is tethered in the flow domain with a uniform flow passing by. In Figure 2, the boundary condition setup is presented.
At the left-hand boundary, a constant velocity (U  ) incoming flow is provided. At the right-hand boundary, a
Neumann boundary condition is exerted on the flow velocity to let the flow convect out of the computational domain
without reflection. At all the lateral boundaries, the zero gradient boundary condition is applied for the velocity. For
the pressure, at all the boundaries, the Neumann boundary condition is used. In the current numerical simulations, a
lower swimming speed condition is considered. The two non-dimensional parameters are set to be: St = 0.93 ,
Re = 88, 000 .

IV. Results and Discussion


In this section, first, the kinematics results are shown. Then, the numerical simulation results, including the
hydrodynamic performance, the wake structures and surface pressure distributions of the computational model are
reported. For all the simulations, to achieve periodic swimming state, five periodic swimming cycles are simulated
and the results from the fifth cycle are adopted for analysis.

A. Undulatory Kinematics
In Figure 3 (a) and (b), the computational model of the swimming dolphin is shown from the lateral and top views.
The fluke part is highlighted by blue color and the midline is denoted by the dashed red line. In Figure 3 (c), the
midlines of the forward swimming dolphin are presented. In x- and y-directions, the length is normalized by the body
length. The fluke upstroke midlines and downstroke midlines are denoted by the red and blue lines respectively. From
Figure 3 (c), it can be observed that, during the forward swimming, the beak of the dolphin oscillates slightly, with a
peak-to-peak amplitude of about 0.05BL. The minimum body oscillation happens at 0.27BL. And the largest
oscillation happens at the fluke tip. At the fluke tip, for this dolphin, an amplitude of 0.27BL is observed. Based on
the midline shapes, the wavelength of the forward swimming dolphin is estimated to be 1.0BL. From the midline
results, asymmetries can be observed between the upstroke and downstroke, especially at the peduncle region.
Compared to the downstroke, the peduncle bends more during the upstroke. This is probably due to the
skeleton/muscle structures at the peduncle region, which will be looked into in the near future. During the fluke
upstroke, the middle part of the body ( 0.3BL  x  0.7 BL ) arches toward the dorsal and forms a concave shape at the
abdomen region.

4
Figure 3. The reconstructed computational model viewed from (a) the lateral view and (b) the top view. The
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

fluke part is colored by blue. In (a), the midline of the swimming dolphin is denoted by the dashed red line. (c)
Midlines of the swimming dolphin during one fluke beating cycle. The red lines denote the upstroke and the
blue lines denote the downstroke of the fluke.

B. Hydrodynamic Performance
In this section, the hydrodynamic performance of the swimming dolphin is investigated with two models, one of
which is the full-body model, and the other one is the fluke-only model.
To describe the hydrodynamic performance and surface pressure distributions, four non-dimensional coefficients
are employed, the thrust coefficient ( CT ) , the lift coefficient ( C L ) , the power coefficient ( CPW ) , and the pressure
( )
coefficient C p . They are defined as:
FT L P p
CT = , CL = , CPW = , Cp = (4)
0.5U  S
2
0.5U  S
2
0.5U  S
3
0.5U  2
where FT is the thrust force acting on the computational model towards the negative x-direction, L is the lift force
towards the positive y-direction, P is the hydrodynamic power, p is the pressure, U  is the incoming flow velocity
and S is the wetted surface area.

Figure 4. (a) The instantaneous thrust and lift force coefficients, and (b) the instantaneous power coefficient of
the full-body and fluke-only models during one fluke beating cycle. The shadow areas denote the downstroke
of the fluke.
Based on the solutions of the Navier-Stokes equations, the surface pressure and shear stress are integrated over the
body surface of the computational model to obtain the instantaneous hydrodynamic forces. Figure 4 shows the thrust
force, lift force, and power coefficients plotted with respect to time in one full fluke beating cycle. The shadow area
denotes the time period of the fluke downstroke. In Figure 4 (a), the thrust force is observed to share a similar trend
between the full-body and fluke-only models, except that during most of the fluke beating cycle, the full-body model

5
shows a smaller thrust value, which implies a drag force generation on the body trunk. In Figure 4 (a), it is also noticed
that, for the lift force, during each fluke beating cycle, the fluke-only model shows one peak value, while the full-body
model shows three peak values. Out of these three peak values, the first peak shows a minor value and is found when
the fluke just passes its bottommost position and starts to move upwards, the second peak value is observed when the
fluke is about to arrive at its topmost position, and the third peak is observed during the halfway of the fluke
downstroke. The lift force coefficient results here indicate that, during the forward swimming, the body trunk part of
the swimming dolphin generates significant lift force when the fluke is moving upwards. In Figure 4 (a), the thrust
force and lift force of the full-body model show a general out-of-phase trend, except at the end of the fluke downstroke.
Corresponding to the force generation, in Figure 4 (b), two major peak values can be observed for the hydrodynamic
power.
( )
Table 1 lists the fluke upstroke-, downstroke- and full stroke-averaged thrust coefficient C T , lift coefficint

(C ) and power coefficient (C ) of these two computational models. It is first observed that, during the fluke
L PW

upstroke, the full-body model generated thrust is about 22.3% lower than the thrust generated by the fluke-only case,
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

which implies a drag force generation on the body trunk. Meanwhile, while the lift force of the fluke-only case shows
a negative value during the fluke upstroke, the full-body model shows a positive lift, which implies a positive lift
generation on the body trunk. During the fluke upstroke, the full-body model consumes a power 35.1% higher than
the fluke-only case. During the fluke downstroke, similar to the upstroke, the fluke is generating thrust while the body
trunk is generating drag force. The lift force generated by the fluke is countered by the body trunk, and the downstroke-
averaged lift force of the full-body model shows a value close to zero. From the full stroke-averaged force and power
results in Table 1, it is estimated that, during one full fluke beating cycle, the body trunk generates about 42.3% of the
total lift while the fluke generates about 57.7%. It is the fluke that consumes the majority of the power, about 85.3%,
while the trunk cost about only 14.7%. However, it needs to be noted that, from the previous studies of swimming fish
[19, 22, 35], significant interactions are observed between the body trunk and fish fins. These interactions increase the
generated thrust and reduce the power consumption of the fish. In the current work, the interactions between the body
trunk and fluke of dolphin swimming are not considered. These interaction effects will be included in the future study.

Table 1. Fluke upstroke-, downstroke- and full stroke-averaged thrust coefficient ( C T ), lift coefficient ( C L )
and power coefficient ( C PW ) of the full-body model and the fluke-only model.
fluke upstroke fluke downstroke full stroke
CT CL C PW CT CL C PW CT CL C PW
full body 0.223 0.105 1.290 0.145 -0.001 1.216 0.184 0.052 1.253
fluke 0.287 -0.364 0.955 0.178 0.423 1.183 0.233 0.030 1.069

C. Wake Structures and Pressure Force Distribution

Figure 5.Wake structure of a forward swimming dolphin at (a) t T = 0.24 and (b) t T = 0.80 .

In this section, the wake structures and surface pressure distributions are looked into to investigate the
corresponding force generation mechanisms. Figure 5 shows the wake structures of the forward swimming dolphin

6
with the Q-criterion iso-surface ( Q=8.0 ). In each fluke beating stroke, four obvious disconnected vortex rings can be
found in the downstream, as denoted in Figure 5 (a). R1 and R3 are two vortex rings generated during the fluke
upstroke and downstroke, respectively. R2 and R4 are two vortex rings generated at the start of the downstroke and
upstroke, when the fluke changes its moving directions. In Figure 5 (a), the fluke is moving upwards, it can be observed
that, the vortex is mainly generated at the ventral side of the fluke, and in Figure 5 (b), when the fluke is moving
downwards, the vortex is mainly generated at the dorsal side of the fluke.
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

Figure 6. The surface pressure distributions at time instants (a) t T = 0.07 , (b) t T = 0.24 , (c) t T = 0.43 and
(d) t T = 0.80 . The corresponding time instants are denoted as (i), (ii), (iii) and (iv) in Figure 4 (a).

Corresponding to the time instants denoted in Figure 4, in Figure 6, the surface pressure is plotted on the whole
computational model to show the pressure distribution changes during one full fluke beating cycle. Figure 6 (a)
corresponds to the first minor lift peak in Figure 4. At this time instant, a higher-pressure region is observed on the
dorsal side of the fluke while on the ventral side of the fluke, a lower-pressure region is observed. Considering the
bending of the body, these two regions generate a forward force and a downward force. The minor lift force peak at
t T = 0.07 is mainly due to the higher-pressure zone at the belly region. In Figure 6 (b), with the fluke moving
upwards, a high-pressure pressure zone is found at the dorsal side of the fluke and the low-pressure suction zone is
found at the ventral side of the fluke. This is in accordance with the near field vortex structures in Figure 5 (a). Due
to the larger bending angle of the peduncle and the paired pressure and suction zones, when t T = 0.24 , the thrust
force in Figure 4 reaches to its peak value. Meanwhile, a high-pressure region is observed at the dorsal side of the
computational model anterior body part. This results in the lift force trough value in Figure 4. When t T = 0.43 , the
fluke is almost at its topmost position. At this time instant, high pressure is observed at the ventral side and low
pressure is observed at the dorsal side of the computational model trunk part. This echoes the large lift force in Figure
4. Figure 6 (d) shows the pressure contours of the computational model when the fluke is at its downstroke halfway
position. At this time instant, the lift force is also at its peak value. However, this is an integrated result of the fluke
ventral side high-pressure pressure zone and fluke dorsal side low-pressure suction zone. The pressure distribution
results of this part show that, for the dolphin in forward swimming, the thrust is mainly generated by the fluke in both
upstroke and downstroke, while the lift force is mainly generated by the trunk part in the fluke upstroke and by the
fluke part in the fluke downstroke.

7
V. Conclusions and Future Work
In this work, a combined experimental and computational methodology is used to study the kinematics and
hydrodynamics of the forward swimming dolphin. The three-dimensional model reconstructed from high-speed video
shows that, the midlines of the swimming dolphin display an asymmetric pattern between the fluke upstroke and
downstroke. During the fluke upstroke, the swimming dolphin bends its peduncle region with a larger angle and arches
its body middle part towards its back. In each fluke beating stroke, four distinct vortex rings can be generated. Two
of these vortex rings are generated during the fluke upstroke/downstroke. The other two vortex rings are generated
during the transition stages between the upstroke and the downstroke. Another thing we notice from this study is that,
during each fluke beating cycle, three lift force peak values can be observed on the dolphin model. Two of them are
found during the fluke upstroke, when the lift force is mainly generated by the body trunk, and the other one is found
during the fluke downstroke, when the lift force is mainly generated by the fluke. Findings of this work help to bring
insight into the understaning of dorsoventral undulatory propulsion mechanisms and may inspire the design of next-
generaton Bio-inspired Autonomous Underwater Vehicles (BAUV).
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

Acknowledgments
This work is supported by ONR grant N00014-14-1-0533, NSF grant CBET-1605232 and University of Virginia
research innovation award (RIA).

References

1. Fish, F. E., and Rohr, J. "Review of dolphin hydrodynamics and swimming performance," Space and Naval
Warfare Systems Command San Diego CA, 1999, pp. No. SPAWAR/CA-TR-1801.
2. Fish, F. E. "The myth and reality of Gray's paradox: implication of dolphin drag reduction for technology,"
Bioinspiration & Biomimetics Vol. 1, No. 2, 2006, pp. R17-R25.
3. Fish, F. E. "Advantages of aquatic animals as models for bio-inspired drones over present AUV technology,"
Bioinspiration & Biomimetics Vol. 15, No. 2, 2020, p. 025001.
4. Roper, D. T., Sharma, S., Sutton, R., and Culverhouse, P. "A review of developments towards biologically
inspired propulsion systems for autonomous underwater vehicles," Proceedings of the Institution of
Mechanical Engineers, Part M: Journal of Engineering for the Maritime Environment Vol. 225, No. 2, 2011,
pp. 77-96.
5. Gray, J. "Studies in animal locomotion," Journal of Experimental Biology Vol. 13, No. 2, 1936, pp. 192-199.
6. Fish, F. E. "Comparative kinematics and hydrodynamics of odontocete cetaceans: morphological and
ecological correlates with swimming performance," Journal of Experimental Biology Vol. 201, No. 20, 1998,
pp. 2867-2877.
7. Lang, T. G., and Norris, K. S. "Swimming speed of a pacific bottlenose porpoise," Science Vol. 151, No.
3710, 1966, pp. 588-590.
8. Parry, D. A. "The swimming of whales and a discussion of Gray's paradox," Journal of Experimental Biology
Vol. 26, No. 1, 1949, pp. 24-34.
9. Purves, P., Dudok van Heel, W., and Jonk, A. "Locomotion in dolphins. Part I: Hydrodynamic experiments
on a model of the bottle-nosed dolphin, Tursiops truncatus,(Mont.)," Aquatic. Mammals Vol. 3, 1975, pp. 5-
31.
10. Webb, P. W. "Hydrodynamics and energetics of fish propulsion," Bulletin of the Fisheries Research Board
of Canada Vol. 190, 1975, pp. 1-159.
11. Fish, F. E., Legac, P., Williams, T. M., and Wei, T. "Measurement of hydrodynamic force generation by
swimming dolphins using bubble DPIV," Journal of Experimental Biology Vol. 217, No. 2, 2014, pp. 252-
260.
12. Chopra, M. G., and Kambe, T. "Hydromechanics of lunate-tail swimming propulsion. Part 2," Journal of
Fluid Mechanics Vol. 79, No. 1, 1977, pp. 49-69.
13. Ayancik, F., Fish, F. E., and Moored, K. W. "Three-dimensional scaling laws of cetacean propulsion
characterize the hydrodynamic interplay of flukes' shape and kinematics," Journal of The Royal Society
Interface Vol. 17, No. 163, 2020, p. 20190655.
14. Isogai, K. "Effect of flexibility of the caudal fin on the propulsive performance of dolphins," Transactions of
the Japan Society for Aeronautical and Space Sciences Vol. 57, No. 1, 2014, pp. 21-30.

8
15. Wang, X., Wei, P., Yuan, Y., Zhang, Z., and Feng, D. "Evaluation of dolphin swimming speed and thrust
based on CFD," International Journal of Offshore and Polar Engineering Vol. 28, No. 02, 2018, pp. 120-
127.
16. Tanaka, H., Li, G., Uchida, Y., Nakamura, M., Ikeda, T., and Liu, H. "Measurement of time-varying
kinematics of a dolphin in burst accelerating swimming," PLOS ONE Vol. 14, No. 1, 2019, p. e0210860.
17. Fish, F. E., Nusbaum, M. K., Beneski, J. T., and Ketten, D. R. "Passive cambering and flexible propulsors:
cetacean flukes," Bioinspiration & Biomimetics Vol. 1, No. 4, 2006, pp. S42-S48.
18. Mittal, R., Dong, H., Bozkurttas, M., Najjar, F. M., Vargas, A., and von Loebbecke, A. "A versatile sharp
interface immersed boundary method for incompressible flows with complex boundaries," Journal of
Computational Physics Vol. 227, No. 10, 2008, pp. 4825-4852.
19. Liu, G., Ren, Y., Dong, H., Akanyeti, O., Liao, J. C., and Lauder, G. V. "Computational analysis of vortex
dynamics and performance enhancement due to body-fin and fin-fin interactions in fish-like locomotion,"
Journal of Fluid Mechanics Vol. 829, 2017, pp. 65-88.
20. Liu, G., Ren, Y., Zhu, J., Bart-Smith, H., and Dong, H. "Thrust producing mechanisms in ray-inspired
underwater vehicle propulsion," Theoretical and Applied Mechanics Letters Vol. 5, No. 1, 2015, pp. 54-57.
Downloaded by UNIVERSITY OF VIRGINIA on June 8, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3015

21. Ren, Y. "Low Dimensional Morphology Analysis and Computational Optimization of Flapping Propulsors
in Nature," Department of Mechanical and Aerospace Engineering. Vol. Ph.D., University of Virginia, 2016.
22. Han, P., Lauder, G. V., and Dong, H. "Hydrodynamics of median-fin interactions in fish-like locomotion:
Effects of fin shape and movement," Physics of Fluids Vol. 32, No. 1, 2020, p. 011902.
23. Han, P., Wang, J., and Dong, H. "Effects of Intermittent Swimming Gait in Fish-Like Locomotion," AIAA
Scitech 2020 Forum. American Institute of Aeronautics and Astronautics, 2020, pp. 2020-1779.
24. Li, C., Dong, H., and Zhao, K. "A balance between aerodynamic and olfactory performance during flight in
Drosophila," Nature Communications Vol. 9, No. 1, 2018, p. 3215.
25. Li, C., Dong, H., and Cheng, B. "Tip vortices formation and evolution of rotating wings at low Reynolds
numbers," Physics of Fluids Vol. 32, No. 2, 2020, p. 021905.
26. Lei, M., and Li, C. "The aerodynamic performance of passive wing pitch in hovering flight," Physics of
Fluids Vol. 32, No. 5, 2020, p. 051902.
27. Wang, J., Ren, Y., Li, C., and Dong, H. "Computational investigation of wing-body interaction and its lift
enhancement effect in hummingbird forward flight," Bioinspiration & Biomimetics Vol. 14, No. 4, 2019, p.
046010.
28. Liang, Z., Dong, H., and Wei, M. "Computational Analysis of Hovering Hummingbird Flight," 48th AIAA
Aerospace Sciences Meeting. AIAA Paper, 2010, pp. 2010-555.
29. Wang, J., Wainwright, D. K., Lindengren, R. E., Lauder, G. V., and Dong, H. "Tuna locomotion: a
computational hydrodynamic analysis of finlet function," Journal of The Royal Society Interface Vol. 17,
No. 165, 2020, p. 20190590.
30. Wang, J., Xi, J., Han, P., Wongwiset, N., Pontius, J., and Dong, H. "Computational analysis of a flapping
uvula on aerodynamics and pharyngeal wall collapsibility in sleep apnea," Journal of Biomechanics Vol. 94,
2019, pp. 88-98.
31. Wang, J., Han, P., Deng, X., Dong, H., Akoz, E., and Moored, K. "Effects of flapping waveforms on the
performance of intermittent swimming in viscous flows," 2018 AIAA Aerospace Sciences Meeting. AIAA
paper, 2018, pp. 2018-0812.
32. Wang, J., Han, P., Zhu, R., Liu, G., Deng, X., and Dong, H. "Wake capture and aerodynamics of passively
pitching tandem flapping plates," 2018 Fluid Dynamics Conference. AIAA paper, 2018, pp. 2018-3236.
33. Deng, X., Han, P., Wang, J., and Dong, H. "A level set based boundary reconstruction method for 3-D bio-
inspired flow simulations with sharp-interface immersed boundary method," 2018 Fluid Dynamics
Conference. AIAA paper, 2018, pp. 2018-4163.
34. Wang, J., Han, P., Deng, X., and Dong, H. "A Versatile Level Set Based Immersed Boundary Reconstruction
for Bio-Inspired Flow Applications," AIAA Scitech 2020 Forum. American Institute of Aeronautics and
Astronautics, 2020, pp. 2020-2235.
35. Han, P., Liu, G., Ren, Y., and Dong, H. "Computational Analysis of 3D Fin-Fin Interaction in Fish’s Steady
Swimming," ASME 2016 Fluids Engineering Division Summer Meeting collocated with the ASME 2016 Heat
Transfer Summer Conference and the ASME 2016 14th International Conference on Nanochannels,
Microchannels, and Minichannels. ASME paper, 2016, pp. V01AT04A006-V01AT04A006.

You might also like