Sei sulla pagina 1di 6

INTEGR. COMP. BIOL.

, 42:1026–1031 (2002)

Vorticity Control in Fish-like Propulsion and Maneuvering1

M. S. TRIANTAFYLLOU, A. H. TECHET, Q. ZHU, D. N. BEAL, F. S. HOVER, AND D. K. P. YUE


Department of Ocean Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139

SYNOPSIS. Vorticity control is employed by marine animals to enhance performance in maneuvering and
propulsion. Studies on fish-like robots and experimental apparatus modelling rigid and flexible fins provide
some of the basic mechanisms employed for controlling vorticity.

INTRODUCTION mechanism of unsteady vorticity injection appears to


Systematic observations of the kinematics of swim- control separation and turbulence production.
ming fish and marine mammals have revealed a new The possibility of extracting energy from oncoming
paradigm of locomotion, distinctly different from con- vortices through an oscillating foil was investigated
ventional propulsion used in man-made vehicles experimentally by Gopalkrishnan et al., (1994), and
(Gray, 1969; Aleyev, 1973; Videler, 1993). Relatively Anderson (1996); and theoretically by Streitlien and
Triantafyllou (1995, 1996). It was shown that energy

Downloaded from http://icb.oxfordjournals.org/ by guest on December 21, 2015


large amplitude, rhythmic, unsteady body motions are
used by fast-swimming animals to achieve remarkable contained in large scale eddies can be retrieved hence
agility underwater. Such large amplitude motions must enhancing propulsive efficiency, or amplifying thrust
be accompanied by optimized flow control mecha- production.
nisms, otherwise these same motions may lead to un- We review evidence of flow control achieved in fish
controlled flow separation and the generation of large like swimming, including separation elimination and
adverse pressure forces and hence considerable deg- turbulence reduction, up to Reynolds numbers of 1.5
radation of performance. Recent work with live ani- 3 106; as well as energy extraction from oncoming
mals has provided important information on the re- vortical flows.
sulting flow, structures (Stamhuis and Videler, 1995; WAKE CONTROL
Anderson, 1996; Mueller et al., 1997; Wolfgang et al.,
1999; Drucker and Lauder, 1999; Lauder, 2000). As in man-made systems, steady forward motion by
Vorticity control, therefore, is of considerable inter- fish is achieved through the generation of a jet which
est, because it is the only known way to control flow provides the thrust needed to overcome the body drag.
Unsteady motion of the body of the fish means that
in an effective way, requiring only a finite amount of
the flow left in the wake is also unsteady; time aver-
sensors and actuators. Indeed, the presence of a few
aging of the flow, however, must have the form of a
large scale vortices renders flow sensing and control
jet. The free shear layers of a jet flow are known to
manageable. The basic rules of vorticity control are
be unstable, i.e., small perturbations are amplified, ini-
understood at least for some basic flows, such as con-
tially exponentially—until nonlinear terms saturate the
trolling the vortex shedding process behind a bluff
growth of these instabilities. The continuous flapping
body through rotational oscillation of the cylinder (Ta-
of the tail and the body provides for a harmonic ex-
neda, 1977; Tokomaru and Dimotakis, 1991), and con-
citation of the flow and hence unsteady patterns are
trolling oncoming vortices using a flapping foil (Go-
expected to form, in addition to any organized vortical
palkrishnan et al., 1994). Taneda (1977) and Tokomaru
patterns shed by the tail, the body or the fins of the
and Dimotakis (1991) imposed a rotational oscillation
animal.
on a two-dimensional cylinder in cross-flow, reducing Minimization of the energy expended in propulsion
the width of the wake and hence the drag coefficient is contingent on satisfying a number of conditions,
significantly, for properly selected parametric values. principally that: (a) the flow does not separate from
Injection of unsteady vorticity is the mechanism the body to generate adverse pressure losses; (b) the
through which the flow control was implemented. energy put in the fluid by the fish is the minimal need-
A flat plate undergoing transverse oscillations in the ed for the given thrust generation. The first condition
form of a traveling wave, placed within an oncoming requires that the body motion is well synchronized to
steady flow, was found to exhibit reduced turbulence avoid the uncontrolled separation and vortex shedding
intensity and separation as the phase speed of the trav- encountered, for example, in a flapping flag. The sec-
eling wave is increased to reach values comparable to ond condition requires that the flow patterns contain
the free stream velocity (Taneda, 1977). The same minimal energy for a given generated thrust.
Hence the concept of a minimum energy wake has
1 From the Symposium Dynamics and Energetics of Animal Swim-
been developed on the basis that the vortical patterns
ming and Flying presented at the Annual Meeting of the Society for
developing through body motion must be compatible
Integrative and Comparative Biology, 2–6 January 2002, at Ana- with the instability properties of the average wake
heim, California. flow. This concept is not so simple to apply, because

1026
VORTICITY CONTROL 1027

the average flow is affected by the unsteady vortical


patterns through the Reynolds stresses. It is coinciden-
tal that the average jet flow in an actual wake behind
a flapping foil, has qualitatively the same bell-shape
expected for a jet: Still, the quantitative values vary
as function of the amplitude and frequency of oscil-
lation.
As a result of these considerations, it was proposed
that a wake-based non-dimensional frequency, or
Strouhal number, St, should be the controlling param-
eter (Triantafyllou et al., 1991, 1993):
St 5 fA/U,
where f is the frequency of oscillation, A the peak-to-
peak motion of the tail of the fish or trailing edge of FIG. 1. The average velocity profiles for swimming and non-swim-
the foil, and U the speed of forward motion. The ming cases. The solid lines show the ‘‘law of the wall.’’
Strouhal number, as defined above, is an indicative
quantity, since there are major simplifications intro-

Downloaded from http://icb.oxfordjournals.org/ by guest on December 21, 2015


duced: A should be the width of the wake, which is or body producing them. The idea of a chain of alter-
unknown beforehand, although it is reasonable to as- nately inclined—with respect to the direction of mo-
sume that it approximately equal to the amplitude of tion—interconnected vorticity rings, as sketched by
oscillation; U should be an average jet velocity, which Lighthill (1975), provides a consistent qualitative pic-
is also unknown. For moderately loaded foils, the av- ture for the flow structure behind oscillating foils. Re-
erage speed of the jet should be close to the speed of cent work by Drucker and Lauder (1999) and Lauder
locomotion; for heavily loaded foils this speed may be (2000) shows the formation of ring-like vortices by
significantly different. fish fins.
For the streamlined body of the fish, one may expect Detailed flow visualization in flapping foils provides
the load to be light to moderate; hence the approxi- a more complex picture: The vortical patterns close
mations used in the definition of the Strouhal number partially on themselves to form apparent ring loops,
are reasonable. Indeed, observations from fish show but the vorticity of each loop connects all the way back
that it is very likely that the Strouhal number of fish to the foil (Freymuth, 1989, Hart et al., 1992). This is
will fall within the theoretically predicted range, i.e., similar to the way Karman vortices interconnect with
between 0.25 and 0.35 (Fish, 1997; Rohr et al., 1998). themselves and to the body, in the wake of a finite
The theoretical prediction is based on the experi- length bluff cylinder, or—in a more simplistic descrip-
mentally observed fact that thrust develops optimally tion—to the way helical vortices in the propeller go
when two counter-rotating vortices develop within a all the way back to the propeller blade. The only dif-
cycle, resulting in a reverse Karman street, and, when ference here is that the connecting vortices to the foil
time-averaged, in a bell-shaped average velocity pro- are entangled in a ‘‘spaghetti’’ like structure, resem-
file. Also, on the fact that such a jet flow has a con- bling the hub vortex of the propeller but with the dif-
vective instability with a preferred frequency and ference that the constituent vortices do not have the
wavelength of maximum spatial amplification: Exci- same rotational sign. This picture has a significant ef-
tation of the wake at this preferred frequency results fect on induced drag.
in the development of a reverse Karman street with
minimum excitation energy; the vortical patterns form SEPARATION AND TURBULENCE REDUCTION THROUGH
partially through rearrangement of the shed vorticity. VORTICITY CONTROL
Similar arguments have been used by Gharib et al. Flow visualization with a 1.20 m long robotic mech-
(1994, 1998) in studying the development of a single anism, the RoboTuna, whose external shape has the
vortex ring through the ramped up motion of a piston form of a bluefin tuna (Triantafyllou and Triantafyllou,
with diameter D, moving at constant velocity U for a 1995), capable of emulating the swimming motion of
time period t. In this case, a non-dimensional forma- a live tuna, exhibited lack of separation effects along
tion time is formed, tU/D, which is qualitatively the its body, even when the motion amplitude reached val-
inverse of the Strouhal number, expressed for an un- ues 10% of the body length. Also, measurement of the
steady, single-event flow. It is found that the circula- flow characteristics of the boundary layer showed ap-
tion of the ring grows optimally, i.e., with minimum parent re-laminarization of the flow when the traveling
energy until a value of tU/D roughly equal to 4, when wave phase speed was close in value to the stream
a maximum is found; after this parasitic vorticity is velocity (Techet et al., 1999; Anderson et al., 1999;
generated, wasting energy. Techet and Triantafyllou, 1999, 2000; Techet, 2001).
The arguments above are based on two-dimensional Under conditions of towing the mechanism without
flow considerations. In three-dimensions, the vortical transverse motion, the boundary layer was character-
patterns must connect with each other and with the foil ized by a turbulent velocity profile. Figure 1 demon-
1028 M. S. TRIANTAFYLLOU ET AL.

FIG. 2. The turbulence intensity along the length of a flapping plate for c/U 5 0.40 (left) and c/U 5 1.2 (right).

strates the difference between the two measured av- shows reduction or even elimination of separation, and
erage velocity profiles: The form of the velocity curve significant reduction of turbulence intensity as c/U ap-

Downloaded from http://icb.oxfordjournals.org/ by guest on December 21, 2015


has the characteristic law of the wall shape when there proaches 1.2 (Techet, 2001). Figure 3 shows a sketch
is no transverse motion, while a laminarized shape ap- of the experimental apparatus, which allows for a trav-
pears when there is active swimming motion with c/U eling wave motion using a series of scotch yokes driv-
5 1.14, where c is the speed of traveling wave and U en by a single motor, activating 8 pistons. Figure 4
the speed of the external stream. shows a comparison between DPIV data at Reynolds
Computational studies on a flat plate undergoing numbers 106 (left) and direct numerical simulations at
traveling wave oscillations within an oncoming Reynolds number 6,000 (right), demonstrating the
stream, show reduction of separation as the phase qualitative similarity between the two cases, despite
speed c increases, with complete elimination of sepa- the large difference in Reynolds number. Figure 5
ration at c/U close to 1 (Zhang, 2001). Similarly, tur- shows the intensity of the horizontal (streamwise) and
bulence intensity is found to decrease non-uniformly vertical components of the velocity, normalized with
across the length of the plate, with increasing c/U, up respect to U2, calculated from ensemble averaged LDV
to a value of about 1.5. The study was conducted at records obtained 4 mm under the plate surface at pis-
Reynolds number, based on plate length, of 6,000 and ton #5. The inflow velocity is U 5 1.0 m/sec and the
then 18,000. Although the flow features around the phase speed varies from c/U 5 0.3 to 2.0. As shown,
plate change with Reynolds number for a non-vibrat- the local turbulence intensity is reduced for c/U up to
ing plate, the flow remains qualitatively similar at the 1.5, but increases for phase speeds beyond this value.
Reynolds numbers studied when c/U is near 1, with a Red denotes points taken at the crest of motion,
preferred value of c/U 5 1.2. The energy to propel the while blue the corresponding data in a trough, showing
flapping plate, defined as the energy to tow the plate that turbulence reduction varies during the cycle.
(which can be negative if the plate produces thrust) Recent studies with a three-dimensional body, in the
plus the energy to oscillate the plate, is minimal at a form of a water snake, undergoing traveling wave os-
value of c/U 5 1.2 (Zhang, 2001; Zhang et al., 2001). cillations, shows non-uniform turbulence reduction
Observations with live fish show that the preferred val- along the body, but the clear trends observed with a
ue of c/U is about 1.2 as well. Figure 2 shows the two-dimensional plate could not be established (Mich-
turbulence intensity along the length of a flapping el et al., 2001).
plate, for two values of c/U, 0.4 and 1.2. There is non-
uniform, but substantial reduction of turbulence inten- ENERGY EXTRACTION FROM ONCOMING
sity in the case of c/U 5 1.2. VORTICAL FLOWS
Experimental work with an 1-meter flapping plate Anecdotal evidence of energy recovery by live fish
in a water tunnel at Reynolds numbers up to 2,000,000 in turbulent flow containing large scale eddies is sub-
stantiated by experimental work with live fish (Beal et
al., 2002); and with simpler experiments using the
controlled motion of a flapping foil within the wake
of a bluff body, such as a D-section cylinder. The foil
is used to extract energy from the flow by properly
timing its motion relative to the arrival of large eddies
(Gopalkrishnan et al., 1994; Beal et al., 2001; Beal,
2002).
Gopalkrishnan et al. (1994) placed a two-dimen-
sional foil in the wake of a cylinder within an oncom-
FIG. 3. The experimental apparatus to generate the wave motion. ing stream to measure the forces and torque needed to
VORTICITY CONTROL 1029

Downloaded from http://icb.oxfordjournals.org/ by guest on December 21, 2015


FIG. 4. Comparison between DPIV data at Re 5 106 (left) and direct numerical simulations at Re 5 6,000 (right).

oscillate the foil so as to produce thrust; and to observe ciency of thrust production was found to be in-
the flow patterns around the oscillating foil. The foil creased compared to other conditions.
would undergo a flapping motion, i.e., harmonic, con- ● A constructive interference pattern, where vortices
trolled heave (transverse) and pitch (angular) motions shed by the foil would coalesce with same-signed
with adjustable amplitude and relative phase. They cylinder vortices forming strong vortices arranged
found that depending on the timing of the motion of in a Karman street or reverse Karman street, de-
the foil relative to the arrival of oncoming vortices, pending on the parametric conditions. Efficiency
three distinct patterns could be observed: was found to be minimal under such conditions.
● An intermediate condition, where foil vortices
● A destructive interference pattern, where vortices would pair with opposite signed cylinder vortices
shed by the foil would first pair and then coalesce forming ‘‘mushroom’’-like structures, expanding the
with opposite-sign vortices shed by the upstream wake width.
cylinder to form a wake consisting of weak vortices
aligned near the centerline of the wake. The effi- Theoretical studies by Streitlien and Triantafyllou
(1995, 1996) showed that energy extraction is possible
when a foil operates in the wake of a bluff body, in
an ‘‘intercepting’’ mode, where the foil would inter-
cept with its leading edge oncoming vortices. Efficien-
cy, defined as the ratio of the useful energy (thrust
times free stream velocity) over expended energy,
could exceed 100% under conditions of energy ex-
traction. The flow patterns associated with the inter-
cepting mode consist of opposite-signed vortices, one
from the cylinder and the other from the foil, which
are pushed close together, resulting in an effective mu-
tual elimination (the inviscid code could not predict
destructive vortex interference, since two inviscid vor-
tices of the opposite sign do not coalesce). A different
mode, named ‘‘slaloming mode,’’ is associated with a
path that keeps away from the core of the oncoming
vortices (slaloming around them): Energy was not re-
covered and the resulting wake consisted of vigorous
vortices resulting from coalescence of same signed
vortices, one form the cylinder and the other from the
foil.
FIG. 5. The intensity of the horizontal (upper figure) and the ver-
tical (lower figure) component of the velocity calculated from en- Figure 6 shows the thrust coefficient Ct and effi-
semble averaged LDV records obtained 4 mm under the mat bound- ciency h of a flapping foil within the wake of a vortex-
ary at piston #5. shedding cylinder. The cylinder has diameter D 5 7.5
1030 M. S. TRIANTAFYLLOU ET AL.

reducing turbulence when the phase speed is close to


1.2 times the free stream velocity.
Energy extraction is seen to be feasible both by flap-
ping two-dimensional foils and by actively swimming
live fish, or fish-like, three-dimensional bodies.
ACKNOWLEDGMENTS
Support by ONR under contract N00014-00-1-0198,
monitored by Dr. P. Bandyopadhyay and by the MIT
Sea Grant Program under Grant NA46RG0434 is
gratefully acknowledged.
REFERENCES
Anderson, E. J., W. R. McGillis, M. A. Grosenbaugh, A. H. Techet,
and M. S. Triantafyllou. 1999. Visualization and analysis of
boundary layer flow in swimming fish. Proc. First Intern. Symp.
Turbulence and Shear Flow Phenomena, Santa Barbara, CA.
Anderson, J. M. 1996. Vortex control for efficient propulsion. Ph.D.
FIG. 6. The thrust coefficient Ct and efficiency h of a flapping foil Diss., MIT/WHOI Joint Program.

Downloaded from http://icb.oxfordjournals.org/ by guest on December 21, 2015


within the wake of a vortex-shedding cylinder, plotted here as func- Bandyopadhyay, P., J. M. Castano, J. Q. Rice, R. B. Phillips, W. H.
tions of the phase angle between the foil motion and the cylinder Nederman, and W. K. Macy. 1997. Low speed maneuvering
motion. hydrodynamics of fish and small underwater vehicles. J. Fluids
Engng. 119:136–119.
Beal, D. N., F. S. Hover, and M. S. Triantafyllou. 2001. The effect
of a vortex wake on the thrust and efficiency of an oscillating
cm and is placed upstream of an oscillating foil; both foil. Proc. 12th Inter. Symp. Unmanned Untethered Submersible
the cylinder and the foil translate at constant velocity Technology (UUST01), Durham, NH.
U in a towing tank. The cylinder oscillates transversely Beal, D. N. 2002. Flow control and energy extraction in fish-like
propulsion. Ph.D. Diss., MIT.
with amplitude 0.4D, while the foil moves transversely Drucker, E. G. and G. V. Lauder. 1999. Locomotor forces on a swim-
both in a linear and angular motion. The linear motion ming fish: Three dimensional vortex wake dynamics quantified
is equal to 0.5D, while the nominal angle of attack is using DPIV. J. Exp. Biol. 202:2393–2412.
ao 5 0 degrees. The Reynolds number, based on the Fish, F. E. 1997. Biological designs for enhanced maneuverability:
foil chord c 5 10 cm, is Re 5 40,000 (Beal 2002). Analysis of marine mammal performance. 10th Int. Symp. Un-
manned Untethered Submersible Technology, Special Session
The horizontal axis is the phase angle between the foil on Bio-Engineering Research Related to AUV, Durham, NH,
motion and the cylinder motion, which affects the rel- pp. 109–117.
ative phase between the arrival of the cylinder-gener- Freymuth, P. 1989. Visualizing the connectivity of vortex systems
ated vortices and the position of the foil. As shown, for pitching wings. J. Fluids Engng. 111:217–220.
Gharib, M., E. Rambod, D. Dabiri, and M. Hammache. 1994. Pul-
thrust and efficiency are affected significantly by the satile heart flow: A universal time scale. Proc. 2nd Intern. Con-
timing of vortex arrival. The relative timing between fer. Experimental Fluid Mech., ed. M Onorato, Torino, Italy.
the arrival of cylinder-generated eddies and eddies Gharib, M., E. Rambod, and K. Shariff. 1998. A universal time scale
shed by the trailing edge of the foil is crucial in de- for vortex ring formation. J. Fluid Mech. 360:121–140.
Gopalkrishnan, R., M. S. Triantafyllou, G. S. Triantafyllou, and D.
termining the type of vortex interaction that will pre- S. Barett. 1994. Active vorticity control in a shear flow using a
vail. Visualization of the flow shows that the three flapping foil. J. Fluid Mech. 274:1–21.
principal patterns identified in Gopalkrishnan et al. Hart, D. P., A. Acosta, and A. Leonard. 1992. Observations of cav-
(1994) can be associated with the efficiency peaks (de- itation and wake structure of unsteady tip vortex flows. Proc.
structive interference mode), efficiency troughs (con- Intern. STG Symp. on Propulsors and Cavitation, Hamburg,
Germany, 121–127.
structive interference mode), and efficiency nodes Koochesfahani, M. 1989. Vortical patterns in the wake of an oscil-
(vortex pairing). Efficiency can even exceed 100% due lating foil. AIAA J. 27:1200–1205.
to energy recovery by the foil (Beal et al., 2001). Lauder, G. V. 2000. Function of the caudal fin during locomotion in
fishes: Kinematics, flow visualization and evolutionary patterns.
CONCLUSIONS Amer. Zool. 40:101–122.
Kumph, J. M., A. H. Techet, D. K. P. Yue, and M. S. Triantafyllou.
Experimental and numerical studies show that fish- 1999. Flow Control of flexible Hull Vehicles. Proc. 11th Inter.
like locomotion employs mechanisms of separation Symp. Unmanned Untethered Submersible Technology
elimination, turbulence reduction and energy extrac- (UUST99). August 22–25, 1999, Durham, New Hampshire.
tion from oncoming flow, to minimize the energy Michel, A. P. M., A. H. Techet, F. S. Hover, and M. S. Triantafyllou.
2001. Experiments with an undulating snake robot. Proc.
needed for locomotion. Oceans, Honolulu.
A two-dimensional plate undergoing traveling wave Mueller, U., B. van den Heuvel, E. Stamhuis, and J. Videler. 1997.
motion captures the essence of the phenomena in- Fish foot prints: Morphology and energetics of the wake behind
volved in turbulence reduction and separation elimi- a continuously swimming mullet (Chelon Labrosus Risso). J.
Exp. Biol. 200:2893–2806.
nation, showing that over a wide range of Reynolds Rohr, J. J., E. W. Hendricks, L. Quigley, F. E. Fish, J. W. Gipatrick,
numbers qualitatively similar mechanisms are at work, and J. Scardina-Ludwig. 1998. SPAWAR Technical Report
providing flow without separation and non-uniformly 1769, San Diego, CA.
VORTICITY CONTROL 1031

Stamhuis, E. and J. Videler. 1995. Quantitative flow analysis around in swimming fish. Proc. Symp. Intern. Soc. Offshore and Polar
aquatic animals using laser sheet particle image velocimetry. J. Engineers (ISOPE’99). Brest France, June 1999.
Exp. Biol. 198:283–294. Techet, A. H. 2001. Experimental visualization of the near-boundary
Streitlien, K. and M. S. Triantafyllou. 1995. Force and moment on flow about fish-like swimming bodies. Ph.D. Diss., MIT/WHOI,
a Joukowski profile in the presence of point vortices. AIAA J. Cambridge/Woods Hole, MA.
33:603–610. Tokomaru, and P. Dimotakis. 1991. Rotary oscillation control of a
Streitlien, K., G. S. Triantafyllou, and M. S. Triantafyllou. 1996. cylinder wake. J. Fluid Mech. 224:77–90.
Efficient foil propulsion through vortex control. AIAA J. 34: Triantafyllou, M. S. and G. S. Triantafyllou. 1995. An efficient
2315–19. swimming machine. Sci. Am. 272(3):64–70.
Taneda, S. 1977. Visual study of unsteady separated flows around Triantafyllou, M. S., D. K. P. Yue, and G. S. Triantafyllou. 2000.
bodies. Progr. Aerosp. Sci. 17:287–348. Hydrodynamics of fish swimming. Ann. Rev. Fluid Mech. 32:
Techet, A. H. and M. S. Triantafyllou. 1999. Experimental study of 33–53.
a waving plate. 52nd Annual Meeting of the Division of Fluid Videler, J. 1993. Fish swimming. Chapman and Hall, London.
Dynamics. American Physical Society, New Orleans, LA, 21– Wolfgang, M. J., J. M. Anderson, M. A. Grosenbaugh, D. K. P. Yue,
24 November 1999. and M. S. Triantafyllou. 1999. Near-body flow dynamics in
Techet, A. H. and M. S. Triantafyllou. Near boundary visualization swimming fish. J. Exp. Biol. 202:2303–2327.
of the flow about fish-like swimming bodies. 2000. 53rd Annual Zhang, X. 2001. Direct numerical simulation of the flow over a
Meeting of the Division of Fluid Dynamics. American Physical flexible plate. Ph.D. Diss., MIT, Cambridge, Mass.
Society, November 19–21, 2000, Washington, DC. Zhang, X., D. K. P. Yue, and M. S. Triantafyllou. 2001. Flow around
Techet, A. H., M. S. Triantafyllou, E. J. Anderson, W. R. McGillis, a flexible plate undergoing traveling wave motion. (under re-
and M. A. Grosenbaugh. 1999. Boundary layer relaminarization view).

Downloaded from http://icb.oxfordjournals.org/ by guest on December 21, 2015

Potrebbero piacerti anche