We consider the motions and associated flow patterns of a swimming giant danio (Danio malabaricus). Experimental flow-visualization techniques have been employed to obtain the unsteady two-dimensional velocity fields around the straight-line swimming motions and a 60 ° turn of the fish in the centerline plane of the fish depth. A three-dimensional numerical method is also employed to predict the total velocity field through simulation. Comparison of the experimental and numerical velocity and vorticity fields shows good agreement. The fish morphology, with its narrow peduncle region, allows for smooth flow into the articulated tail, which is able to sustain a large load for thrust generation. Streamlines of the flow detail complex processes that enhance the efficiency of flow actuation by the tail. The fish benefits from smooth near-body flow patterns and the generation of controlled body-bound vorticity, which is propagated towards the tail, shed prior to the peduncle region and then manipulated by the caudal fin to form large-scale vortical structures with minimum wasted energy. This manipulation of body-generated vorticity and its interaction with the vorticity generated by the oscillating caudal fin are fundamental to the propulsion and maneuvering capabilities of fish.

Several theories have been developed to explain fish swimming performance. Wu (1961, 1971a,b,c) and Lighthill (1960, 1975) have investigated the propulsion characteristics of flexible two-dimensional and three-dimensional plates using linearized potential flow theory. While these seminal works provide insight into the basic swimming propulsive mechanisms, the details of the three-dimensional flow and the dynamics of the shed vorticity are still not well understood. Several comprehensive works on fish swimming thoroughly analyze the kinematics and the dynamics of straight-line and unsteady swimming motions for a variety of aquatic species (Videler, 1993; Blake, 1983) utilizing these linear theories to estimate the propulsive dynamics. However, such methods overpredict the thrust on more realistic fish-like forms (Videler, 1981) because basic assumptions about the linear slender body theory are violated.

The study of unsteady lifting surface motion for both swimming and flying has provided insight into the details of the dynamics of fish propulsion methods. The aerodynamics of hovering insect flight (Osborne, 1951; Ellington, 1984; Ellington et al., 1996) illustrates the importance of unsteady flow mechanisms in achieving a large loading on the wing, including a delay in stall. The unsteady motions of hydrofoils have been studied analytically and numerically by several investigators (von Kármán and Burgess, 1935; Wu, 1961, 1971a,b,c; Lighthill, 1960, 1975; Chopra, 1976; Chopra and Kambe, 1977; Lan, 1979), but these methods employ linearized body boundary conditions as well as assumed wake shapes.

Most fish move with amplitudes and frequencies that exceed the limit of linear lifting surface theories. Videler and Hess (1984) showed that two species of fish exhibited high propulsive efficiency, close to the actuator disk limit, with large lateral motion amplitude. Triantafyllou et al. (1993) showed that a variety of fish and cetaceans swim with a frequency and amplitude of tail motion that are within a narrow range of Strouhal numbers, minimizing energy lost in the wake for a given momentum and increasing efficiency. This Strouhal number range corresponds to the regime of maximum stability of the vortex wake thrust jet.

Several investigators have studied the mechanisms of vorticity control achieved by unsteady motion of a body in a fluid (Taneda and Tomonari, 1974; ffowcs-Williams and Zhao, 1989; Tokomaru and Dimotakis, 1991). Gopalkrishnan et al. (1994), Streitlien et al. (1996) and Anderson et al. (1998) utilized experimental and computational methods to demonstrate active control of a shear flow with an oscillating foil, revealing energy recovery flow control mechanisms that can increase or decrease the efficiency. Direct experimental dynamic measurements of the energetic make-up and efficiency of live swimming fish are difficult to obtain, however, and may not be characteristic of natural behavior. As a result, Barrett (1996) constructed a robotic underwater flexible hull vehicle, in the shape of a tuna with a lunate tail, on which direct dynamic measurements and energetic analyses could be performed. Over small ranges of various swimming variables, the robot could achieve extremely high propulsive efficiencies and lower drag than that of the rigid body hull through mechanisms of boundary layer relaminarization and vorticity control by the tail fin (Barrett et al., 1999).

Several researchers have studied, through visualization, the wake of a naturally swimming fish. Techniques that have been used include dye visualization (Rosen, 1959; Aleyev, 1977) and visualization using a stratified layer (Rosen, 1959; McCutchen, 1976). Neither method has elucidated all the details of the flow near the body and in the wake. Stamhuis and Videler (1995) utilized experimental particle image velocimetry to capture the flow dynamics around several live swimming organisms and to analyze the energetic make-up of the wake. Anderson (1996) used experimental digital particle image velocimetry (DPIV) to visualize the wake behind a swimming giant danio (Danio malabaricus) and identified the active manipulation of shed wake vorticity to create a reverse Kármán vortex street. Müller et al. (1997) analyzed the wake of a swimming mullet (Chelon labrosus Risso) using similar DPIV techniques and concluded that the manipulation of the wake structure resulted in high propulsive efficiencies.

In the present study, we present quantitative, multipoint measurements of the flow field around a small naturally swimming fish using DPIV. The results include flow measurements for straight-line swimming and rapid maneuvering of a giant danio (Danio malabaricus). We also develop a numerical method to simulate the hydrodynamics of an actively swimming fish-like body. The characteristics of real fish swimming at large Reynolds numbers are approximated by assuming that viscous effects are confined to a thin boundary layer and wake region. The formulation allows for the satisfaction of the exact body boundary conditions and for the nonlinear evolution of the shed wakes. This inviscid approach is computationally an order of magnitude more rapid than the fully viscous numerical methods that are essential at lower Reynolds numbers, such as those developed by Liu et al. (1997) to study tadpole locomotion. Our numerical method is then applied to analyze the wake dynamics and the near-body hydrodynamics of the motions of the giant danio, which were captured using DPIV, including both continuous straight-line swimming and a 60 ° turning maneuver. We present a sample of the experimental and numerical results, which demonstrate that the controlled actuation of the caudal fin and release of bound vorticity contributes to the efficient production of a reverse von Kármán street thrust jet.

Experimental methods and apparatus

We employ digital particle image velocimetry (DPIV), first introduced by Willert and Gharib (1991), for flow measurement and visualization. The DPIV algorithm is applied to two images of the flow separated by small time dt, and small subsets of image data (interrogation windows) are compared at the same location in each image by computing the spatial cross correlation. When the particle images match well, the cross-correlation function is sharply peaked, and the location of the peak marks the displacement of the particle image pattern. The algorithm is then applied throughout the images to obtain the entire flow field.

For the DPIV algorithm to work well, there must be adequate particle seeding, randomly distributed over the entire image plane. Willert and Gharib (1991) systematically studied the seeding requirement and found that 10–20 particles per window are sufficient to ensure good correlations. For a complete discussion of general DPIV capabilities and limitations, see Willert and Gharib (1991). Unique problems arise in using DPIV to visualize flows around moving, deformable bodies. As the fish swims and boundaries move, interrogation windows cross boundaries and contain spurious image data. This biases the cross-correlation peak towards the displacement of the boundary, resulting in spurious displacement data for the fluid near the boundary. A complete discussion of the near-body DPIV techniques we employ, including their capabilities, limitations and validation, can be found in Anderson (1996).

The windowing process in DPIV filters out wavenumbers (spatial frequencies) larger than the inverse interrogation window size. This is a problem when the flow field of interest has fine-scale motions and flow reversals, such as in a vortex core. To resolve such fine-scale features correctly, the image view should be enlarged in the region of interest by moving the camera or by reducing the window size. Willert and Gharib (1991) showed that there is approximately 70 % attenuation of 32 pixel features when the data are processed with 16 pixel ×?16 pixel interrogation windows, and one should always loosely interpret displacement data for scales less than twice the window size.

We calibrated our hardware and processing techniques in the same manner as Willert and Gharib (1991) using an artificially generated test image displaced with a rotary table instead of a linear table. The algorithm is very robust up to the Nyquist limit, despite curved particle trajectories. Our relative displacement error is generally less than 5 %, and usually less than 2 % for moderate (>1 pixel) displacement.

Fig. 1 summarizes the experimental apparatus for DPIV implementation in these experiments. The measurement volume, a 51 cm × 25 cm × 28 cm glass tank, was seeded with small (20–40 μm diameter), neutrally buoyant fluorescent polymer spheres. A planar slice of the flow approximately 3 mm thick was illuminated with a 6 W argon-ion laser connected to a fiberoptic cable and light sheet optics (Dantec).

Fig. 1.

Schematic diagram of experimental apparatus. The fish is confined within a thin layer of fluid between the water surface and a permeable screen. The horizontal midplane of the fish is illuminated from the side and imaged from above.

Fig. 1.

Schematic diagram of experimental apparatus. The fish is confined within a thin layer of fluid between the water surface and a permeable screen. The horizontal midplane of the fish is illuminated from the side and imaged from above.

The particle flow was imaged using a high-resolution (768 pixels × 480 pixels) black-and-white CCD video camera (Texas Instruments, Multicam MC-1134P) and recorded to video disc using a Sony CRVdisc videodisc recorder (LVR-5000A). Images were acquired in real time at the camera frame rate of 30 Hz in dual-field mode, allowing full vertical resolution with no interlace. Asynchronous shuttering of the laser light with a timing box (General Pixels) synchronized to the video signal allows for higher effective frame rates. Although image pairs are collected at 15 Hz, asynchronous shuttering allows the difference between images in a pair to be very small (approximately 5 ms).

As illustrated in Fig. 1, the fish was confined within a layer of fluid approximately 3 cm thick, between the free surface and a highly permeable screen consisting of a steel wire grid (2.5 cm openings) and a fine nylon mesh (approximately 3 mm openings). The clearance between the fish and the screen and free surface was estimated to be 3 mm. The screen was located approximately 25 cm from the tank bottom. The screen confined the fish yet allowed fluid to pass through it, and the plane in which the fish was allowed to swim was sufficiently wide to eliminate any wall effects. The flow in the horizontal symmetry plane of the fish was revealed by orienting the laser light sheet at the mid-depth of the confinement layer. In general, the light sheet illuminated the fish at the caudal fork; however, in some instances, slight bending of the caudal fin caused the tips of the tail to be illuminated. The free surface distortions caused by typical swimming and maneuvering motions were observed visually to be negligible. This suggests that there was minimal interaction between the wake vorticity and the free surface.

We chose to experiment with a fast-swimming tropical species, Danio malabaricus (Jerdon), the giant danio, whose morphology is very similar to that of the pearl danio studied by Rosen (1959). The dominant fins are the caudal fin, the flexible dorsal fin and the stiffer anal fin, which acts like a keel. Anterior to the anal fin, the ventral fins flare out during maneuvering but participate little in straight-line swimming. Similarly, the pectoral fins are used in maneuvering and are generally kept flush to the body during straight-line motion.

Several fish were used with body lengths between 5 and 10 cm. The maximum span of the caudal fin was approximately 23 % of the body length. Whenever possible, data will be presented normalized with respect to fish body length L. We found the giant danio well suited to our experiments since they are steady, fast swimmers and hardy with respect to handling. The laser light sheet was oriented horizontally in the midplane so that it hit the fish near the vertical plane of symmetry. The fish were allowed to swim freely in the confinement layer, during which time the video was recorded to laser disc. The fish showed no adverse reaction to either the particles or the laser light, and the kinematic behavior of the fish was identical whether the laser was on or off.

The DPIV images obtained of the fish swimming in a straight line or undergoing maneuvering motions are processed to produce velocity field images. Each arrow in a velocity field plot represents one cross-correlation calculation, and the length of each arrow indicates the relative velocity magnitude. The images obtained were processed using 32 pixel × 32 pixel interrogation windows, shifted in eight-pixel steps, which corresponds to oversampling the available image data four times. This interrogation window dimension is sufficiently small compared with the length of the fish (0.076L) such that flow scales of size 0.15L and above can be measured accurately. The velocity data are then used to compute the component of the vorticity perpendicular to the DPIV plane, so that large-scale vorticity components along the body and in the wake may be observed and quantified.

Numerical method

We develop a computational tool for investigating the swimming characteristics of a three-dimensional flexible body with arbitrary motions and geometry. We consider as a canonical problem the abrupt starting from rest to a constant velocity U of a streamlined flexible body geometry undergoing prescribed undulations about its mean line, with arbitrary distribution of sharp-trailing-edged fins. Thin shear layer wakes are shed continuously from the sharp trailing edges of the body as time proceeds. With the exception of the wake, the fluid is assumed to be inviscid and irrotational, as well as incompressible, allowing for the existence of a velocity potential We define two coordinate systems, an inertial global coordinate system O,X,Y,Z, fixed in space in the fluid, and a local coordinate system o,x,y,z, instantaneously fixed in the flexible body and orthonormal to the stretched-straight mean line and body section plane. The mean periodic undulations of the body are described with reference to o,x,y,z, and the global translational and rotational motions of the body are described with respect to O,X,Y,Z. Fig. 2 details the coordinate systems used in the problem. All time and length scales are chosen to be non-dimensional with respect to the body length L=1.

Fig. 2.

Coordinate system convention for the analysis of straight-line swimming (top view). Local body motions are described with respect to the body-fixed o,x,y,z coordinate system, and the body trajectory is described with respect to an inertial reference frame O,X,Y,Z. U, swimming speed; Sb, body surface in the computational domain; Sw, wake surface in the computational domain.

Fig. 2.

Coordinate system convention for the analysis of straight-line swimming (top view). Local body motions are described with respect to the body-fixed o,x,y,z coordinate system, and the body trajectory is described with respect to an inertial reference frame O,X,Y,Z. U, swimming speed; Sb, body surface in the computational domain; Sw, wake surface in the computational domain.

We consider the computational domain to be enclosed by three surfaces, the body surface Sb, an infinitesimally thin wake sheet Sw, and a far-field boundary at infinity S. Laplace’s equation for the velocity potential governs the conservation of mass of the fluid, and the unsteady Bernoulli equation describes the pressure everywhere in the field. The velocity potential can be further described by the linear superposition of a body perturbation velocity potential and a wake perturbation velocity potential each satisfying the Laplace field equation. Although the strength of the wake being shed at the body trailing edges at any time is unknown, the strengths of all previously shed wake surfaces are known, with the exception of the initial condition of the impulsive start when no previously shed wakes are present. Thus, the wake perturbation velocity potential is known, and we formulate the boundary value problem for the body perturbation velocity potential

The prescribed body motion gives the kinematic body boundary condition in terms of the body perturbation potential for a body normal vector :
where is a gradient vector. The prescribed motion of the body forces the length of the shed wake to increase, as an unsteady Kutta condition is applied to the sharp trailing edges of the body to enforce smooth flow separation. The strength of this portion of the shed wake sheet is unknown. At any time, however, we can set the strength of the shed wake velocity potential to the jump in the body perturbation potential between the upper and lower surfaces near the trailing edge (TE) (Kinnas and Hsin, 1992):
A radiation condition is imposed such that the influence of the velocity potential decays rapidly to zero in the far field. Green’s formulation was considered to solve the problem for the body perturbation velocity potential The second form of Green’s theorem is employed for the body perturbation potential and for the three-dimensional source potential Green’s function 1/|r|, both of which independently satisfy the Laplace equation. r is the distance from observation point to any point on the boundary, This choice of Green’s function is standard in the development of source-dipole boundary integral methods for an ideal fluid in the absence of a free surface. Thus, the body perturbation velocity potential at any point can be described in terms of the integral around the boundaries of the contributions to the perturbation velocity potential. If this point is on the boundary, Green’s theorem can be written:
where S is composed of all the bounding surfaces, S=Sb+Sw+S, ∂ n is a derivative in the direction of the outward surface normal vector and ds is an elemental unit of surface area. The influences of the source and dipole singularities in the integrands of equation decay as 1/r and 1/r2, respectively. Thus, the boundary at infinity S can be shown to have negligible influence.

The unsteady wake strength is a function of space and time, The pressure across the wake shear layer is continuous, and the wake surface is a material surface and can thus be represented by a smoothly varying strength dipole sheet. From Kelvin’s theorem, as the foil’s circulation changes, so does the strength of the shed wake dipole. At any time, the strength of the entire previously shed wake is known, and thus the strength of is known, except for the recently shed shear layer which smoothly separates from predetermined trailing edges. The unknown strength of the newest portion of the wake sheet is addressed through the Kutta condition as described by equation 2. Thus, the body perturbation velocity potential and the change in circulation of the various wake-shedding lifting surfaces of the body can be written in terms of surface integrals over the body surface Sb and the wake sheet surface Sw.

Representation of the time-dependent continuous potential distributions over the body and wake surfaces is developed in discrete form. The body and the wake are each divided into quadrilateral panels. The source-dipole distribution representing the body is approximated as piecewise constant over each body panel; similarly, the dipole distribution representing the thin shear layer wake is approximated as piecewise constant over each wake panel. The collocation point for each panel is the geometric center of the panel. The body grid is generated with higher panel densities in regions of presumed rapid potential variation and in regions with complex geometric considerations, such as areas of large curvature or lifting surface tips, to reduce panel skew and to increase accuracy with respect to the unsteady, continuous velocity potential distribution.

At any time t, the body perturbation velocity potential at each panel collocation point can then be found in terms of the perturbation potentials at all the other panels. A linear system of equations results for with the wake perturbation velocity potential of the new wake panels described in terms of at the trailing edges. A discrete form of the kinematic boundary condition on the body surface (equation 1) gives the normal velocity at each panel. A simple system of linear equations for at any given time t is expressed in the general form:
where is the matrix of influence coefficients for the source-dipole potential panel distribution, ϕb is a column vector of the body perturbation velocity potentials, and is the column vector of normal velocities for the source-dipole potential panel distribution. Details of the linear system development are found in Appendix A. At each time step, this linear system of equations is solved to find and through the use of an iterative scheme. is representative of the change in the total circulation of the body lifting surfaces. While the number of body panels k is constant, the total number of wake panels nw increases with time, as a wake of time-varying strength is continually shed from the trailing edges of the body. The numerical scheme is temporally integrated using a fourth-order Runge–Kutta scheme.

Thin fins and wakes

In general, we consider fish geometries which are composed of a main body, with smoothly varying elliptical cross sections, and a tail fin, with rounded leading and sharp trailing edges. On occasion, we model additional fins to investigate the effects of a more accurate geometrical representation. Modeling these fins as finite-thickness lifting surfaces with rounded leading and sharp trailing edges becomes impractical as the fins become smaller and thinner with respect to the main body. As the angle between adjacent panels at the leading edge approaches 2π, the resulting square-root singularity destabilizes the influence coefficient matrix (Katz and Plotkin, 1991).

The addition of thin fins to the main body and caudal fin arrangement is accomplished using vortex lattice surfaces, such as those investigated by Lan (1979), Cheng et al. (1991) and mostly recently H. Kagemoto, D. Yue and M. Triantafyllou (unpublished data). A surface of zero thickness is discretized into a grid of planar quadrilateral elements, and the doublet jump strength at each panel is assumed to be constant. The source distribution strength across each panel on the thin surface is set identically to zero, and a zero-normal velocity boundary condition is again applied at each thin panel’s collocation point with normal .
The application of this no-flux boundary condition results in a set of equations for a geometry with both a typical body and thin fins for the doublet jump strength ϕb across a thin fin panel at
The vortex lattice solution algorithm requires the differentiation of equation 3 an additional time to apply the Neumann boundary condition when the panel source strength is zero. The details of the development of this higher-order algorithm are analogous to those presented here for a source-dipole panel method and thus will not be discussed. However, it should be mentioned that some care must be taken in the geometric modeling of these thin fins not to use a fine panel mesh density, because a hypersingularity present in the self-influence coefficients can destabilize the numerical algorithm for very fine discretizations. The size of the thin fin panels should be of the order of the largest body panels to minimize the matrix condition number. For constant potential jump distributions over a quadrilateral geometry, the hypersingularity requires no special treatment for the typical grid discretizations applied in these simulations, and the influence coefficients calculations can performed by standard vortex-lattice algorithm (Katz and Plotkin, 1991). Removal of the hypersingularity for more complex geometries and potential jump distributions could be achieved through various numerical and analytical techniques (Martin and Rizzo, 1989; Guiggiani et al., 1992; Martin, 1996). The system of equations is expressed as a coupled mixed boundary-value problem, consisting of a set of matrix equations for ϕb on the body geometries ϕbb and for the ϕb jump across the thin fin geometries ϕbf, as well as the strength of the unsteadily shed wake. In general form, the system is expressed as:
where Q11, Q12, Q21 and Q22 are sub-matrices of the influence coefficients defined in the list of symbols, and P1 and P2 are partial column vectors of panel normal velocities defined in the list of symbols. A wake equivalent to that described in the previous section is shed from the thin fin trailing edge as time progresses, as prescribed by a Kutta condition. The Kutta condition for the thin wing states that the change in circulation around the lifting surface section must equal the change in vorticity shed into the wake, in accordance with Kelvin’s theorem, which states that the change in total circulation for the entire fluid domain must be zero:
where Γ is the circulation strength. Thus, the strength of a new wake doublet panel is equal to the strength of the thin fin trailing edge panel doublet at the previous time step, t−dt. The coupling of these two types of surface representation allows for the simulation of complex bodies undergoing unsteady arbitrary flexing motions.

When the solution of the boundary value problem has been obtained at each time step, the wake panel endpoints are convected by a velocity field described by a desingularized version of the Biot–Savart law. The desingularization technique employed (Krasny, 1986) eliminates the infinitesimal vortex sheet singularity and the associated ill-posedness. The inclusion of the desingularization is necessary to prevent nonlinear energy transfer to the highest wavenumber modes and simulation breakdown, in addition to non-physical solution growth caused by the numerical instabilities.

A wake vortex element of strength Γv, consisting of the side edge of one quadrilateral doublet panel, is assigned core radius δw, such that the velocity induced by this vortex element at can be expressed as:
where is the tangent vector to the vortex element, is the body Pb is the integral of the product of the local force and the vector with magnitude r from the vortex element to the field point the path of integration is along the length of the vortex element, and dl is an elemental unit of length along the vortex segment. As approaches the vortex, r approaches zero, and the velocity field expressed by equation 8 approaches a finite limit.

Similarly, the body and thin fin doublet panels are desingularized by the same technique, with arbitrary body desingularization radius δb. This desingularization assists in stabilizing the solution algorithm for coupled mixed boundary value problems with bodies of fine mesh discretizations. Additionally, when upstream-shed vorticity impinges on downstream body elements, non-physical free wake acceleration and deformation are avoided. These impinging wake panels pass around the outside of the body surface, often tangentially, convected by the body surface normal and tangential velocities. Without this body desingularization, free wake panels interacting with flexing body panels might convect inside the body.

Integrated quantities

At each time step, we find the velocity and force distribution over the body panels. Once the unknown body perturbation velocity potential ϕb is known for each panel, the tangential surface velocity on each panel can be found numerically by spatial differentiation of ϕb over the body using a finite-difference scheme. The pressure coefficient CP at each surface panel can then be found by employing the unsteady Bernoulli equation:
The partial derivative ∂ ϕb/∂ t is found by taking the material derivative of ϕb at each panel using a temporal second-order backward-difference technique and subtracting the convective term arising from the imposed motions of the unsteady flexing body. The force Fb is found by integrating the pressure on each panel over the body:
where ρ is the density of the fluid. For thick bodies, the pressure is calculated on only one side of each panel, whereas for thin fin surfaces, the total force realized is related directly to the pressure jump between the upper and lower surfaces of the fin. The lower surface velocities are calculated for the thin fins, and the pressure jump is integrated to find the total force on the thin fin Fb,thin. The method of Lan (1979) is employed to model the leading edge suction forces on any thin fins, provided that the sweep angle of the fin is not high. The caudal fin is modeled as a finite-thickness high-aspect-ratio foil, and the leading edge suction acting on the caudal fin is resolved through direct pressure integration over the body.

The power transmitted to the fluid by the motions of the local imposed body velocity. In general, for a geometry with a number of thick bodies and thin fin surfaces, the time records of the total force Fb and the total power delivered to the fluid Pb may be quite complex, even for sinusoidal transverse motions. The total force is decomposed into a side force Lb perpendicular to the direction of straight-line swimming, which should have a zero temporal mean, and a longitudinal force Tb, which should have a non-zero mean value of thrust or drag for a mean swimming speed U in the absence of skin frictional drag modeling. The mean value of thrust generated over one period for harmonic body undulations is denoted . Similarly, the mean power transmitted to the fluid over a period of harmonic body motion is denoted . This allows us to measure the classical hydrodynamic propulsive efficiency η for swimming speed U as

The unsteady code was validated by simulating the impulsively started motion to a constant speed U of a finite-aspect ratio foil at a small angle of attack, in a fluid otherwise at rest. Convergence of the numerical method was confirmed by systematically varying the time step size dt and the number of panels k around the foil, in addition to the aspect ratio and the desingularization parameter δw for the wake, and comparing the steady lift coefficient with the experimental values illustrated in Fig. 17-5 of Hoerner (1985).

Steady swimming of a giant danio

DPIV experimental results

In the DPIV experiments, the fish exhibited varied behavior with tail double amplitudes between 0.11L and 0.21L and frequencies up to 6.2 Hz. Consequently, the Strouhal number also showed variability. Typical relative errors in the analysis of the fish and wake kinematics were 25 % for the tail double amplitude measurements (0.02–0.05L), 15 % for the frequency measurements (up to 0.9 Hz), 5 % for the speed measurement (up to 0.055 L s−1) and 30 % for the Strouhal number calculation. Much of the error is simply because the fish did not always swim with a constant tail-beat amplitude or frequency and does not actually represent measurement error. The values reported here for experiments are means. The wake Strouhal numbers (St) clearly support the conclusions of Triantafyllou et al. (1993) that fish tend to swim within a narrow range of Strouhal numbers in order to exploit the natural instabilities of the thrust jet created. Although our kinematic measurements are approximate, most of the Strouhal numbers observed in our experiments are within the predicted optimal range of 0.25<St<0.4.

Several swimming cycles were captured and studied using DPIV; here, we present one particularly clear case where the fish swam directly away from the light source, thus revealing the flow on both sides of the body and in the wake without shadows. For this case, the fluid ahead of the fish was nearly stationary, and all motion can be attributed to the motion of the fish. Velocities very near the body boundary are not shown in the following results because the image of the fish itself contaminates the flow image data. Spurious data are directly removed, and adjacent data are treated specially to avoid the inclusion of contaminated data in the smoothing filter and calculations of vorticity and circulation.

Fig. 3 shows the velocity field and vorticity contours adjacent to and in the wake of the fish during steady straight swimming. The DPIV images reveal that the fish was swimming at an approximately constant mean velocity of 8.9 cm s−1 (1.1 L s−1), a normal observed swimming speed for the giant danio, for the 0.23 s duration of the cycle captured. The fish traveled steadily from one end of the tank to the other, passing through the imaging window in the center of the tank.

Fig. 3.

(A–D) Velocity field (left) and vorticity contours (right) during steady, straight-line swimming viewed from above. Each arrow represents one cross-correlation calculation with the length indicating the relative magnitude of the velocity. The maximum velocity shown is approximately 8 cm s−1. Vorticity contours range from −5.0 to −20.0 s−1 (dotted lines) and from +5.0 to +20.0 s−1 (solid lines) in ±1.67 s−1 increments. Time t is given in each view. Kinematic variables of the straight-line swimming motion: swimming speed U=1.1 L s−1; tail-beat frequency f=3.3 Hz; tail-tip double amplitude A=0.16 L; backbone wavelength λ=1.1 L; Strouhal number St=0.45. L fish body length.

Fig. 3.

(A–D) Velocity field (left) and vorticity contours (right) during steady, straight-line swimming viewed from above. Each arrow represents one cross-correlation calculation with the length indicating the relative magnitude of the velocity. The maximum velocity shown is approximately 8 cm s−1. Vorticity contours range from −5.0 to −20.0 s−1 (dotted lines) and from +5.0 to +20.0 s−1 (solid lines) in ±1.67 s−1 increments. Time t is given in each view. Kinematic variables of the straight-line swimming motion: swimming speed U=1.1 L s−1; tail-beat frequency f=3.3 Hz; tail-tip double amplitude A=0.16 L; backbone wavelength λ=1.1 L; Strouhal number St=0.45. L fish body length.

Any acceleration or deceleration of the body occurred at either end of the tank, well before or after the images shown in Fig. 3 in the center of the tank, so the flow features are likely to be typical of steady-state motion. Nearly an entire swimming cycle is shown with the successive data sets separated by approximately 0.22T, where T is the period of the tail oscillation. The figures show the fish body exactly as it appears in the first image of each image pair except that it is colored uniformly for clarity. Slight variations in the recorded body image exist due to the fact that in some images the caudal fork is illuminated by the laser sheet while in others the caudal fin tips are illuminated.

At the start of the swimming cycle (Fig. 3A), the fish tail is flexed maximally to the left. A clockwise vortex is forming with its center near the tip of the tail as a result of shed body-generated vorticity associated with the body undulation. The vortex is already well formed, although at this point the tail has not begun its rightward motion. Subsequent rightward tail motion will contribute same-signed vorticity to this vortex. Immediately to the left of the fish and slightly anterior to the tail, the flow follows the rightward velocity of the body. Slightly upstream of this position at approximately the mid-body of the fish, longitudinal flow towards the tail marks the formation of the next bound counterclockwise vortex.

In Fig. 3B, the tail has moved to the right, and the new clockwise vortex is now clearly visible in the wake. The region of longitudinal flow has moved rearward and to the right with the tail motion. In Fig. 3C, the tail is maximally displaced to the right ready to begin the return stroke. The flow to the right of the fish follows the body motion to the left which defines the circular flow that becomes the next counterclockwise vortex. Fig. 3D shows this new vortex in the wake as well as the next clockwise vortex forming along the body. The swimming cycle then concludes with the tail in the maximum downward position and the next clockwise vortex positioned at the tail tip, ready to be shed on the upstroke.

The closed circular contours shown in Fig. 3 indicate concentrated vortices arranged in a jet pattern. Body-bound vorticity cannot manifest itself, as it would require processing of the velocity field through the body of the fish. The wake organizes into a stationary reverse Kármán street: vortices align in a staggered pattern such that the net wake produces a jet flow. The maximum velocity along the centerline of the jet is approximately 90 % of the mean velocity of the fish. The flow adjacent to the fish is strongly influenced by the body undulation, as demonstrated by the clockwise flow at approximately three-quarters of the body length from the nose (Fig. 3D). This clockwise flow is the beginning of the next vortex destined to be shed by the tail into the wake. The new ‘vortex’ is formed well before encountering the tail. The increasing lateral motion of the posterior body region augments the circulation and strengthens the vortex before it is shed into the wake, just before and at the peduncle. The tail manipulates this vorticity through repositioning and shedding of additional vorticity.

Flow properties were quantified by averaging over one swimming cycle. The maximum wake velocity along the wake centerline was computed by averaging the streamwise velocity at several locations over one wavelength downstream, at one instant in time. The maximum wake velocity was 0.88U or 0.97 L s−1. In the regions near the body where the undulation wave has the most effect, the maximum longitudinal (streamwise) velocity component was approximately 0.14U, and the maximum transverse velocity (towards or away from the body in the normal direction) was approximately 0.39U. The experimental circulation re of the wake vortices was calculated by integrating along circular contours around several wake vortices using equation 11. The average circulation of the wake vortices was Γ/LU=0.3 with core radius R ≈0.1L:

Numerical modeling

In the numerical calculations, we approximate the body to be as close as possible to the giant danio body, including the caudal fin, and the dorsal and anal fins. The tail generates the largest wake, and a primary separation line is chosen to be the trailing edge of the tail. Although the giant danio possesses a large dorsal fin, a large anal fin and smaller ventral and pectoral fins, these edges are treated differently in the following cases. It is obvious from the experimental evidence that the tail realizes the greatest unsteady motion amplitudes, which would sustain the largest unsteady lifting forces and generate the strongest vorticity structures; however, the influence of vorticity shed upstream of the tail by secondary lifting surfaces on the flow around the tail, the total wake structure and the force on the body is unclear.

Lighthill (1960, 1975) hypothesized that if there was a large gap between the dorsal and caudal fins, any vorticity shed from dorsal or ventral lifting surfaces would behave very differently from that in the presence of a fin bridging the gap between the two fins. Specifically, for certain values of the phase between the tail and the dorsal fin motions, recovery of upstream shed energy may be achieved. Streitlien et al. (1996) showed computationally such energy recovery using an oscillating foil in a shear flow.

To clarify this point, a computational body geometry was chosen which includes the main portion of the body, the caudal fin and the smaller dorsal and anal fins (Fig. 4). The main body sections are assumed to be elliptical with a major-to-minor axis ratio of =2.2, where the major axis corresponds to the height of the body. A curve-fitting technique is used to determine the profile shape of the body with L=1.0, and is given simply by:

Fig. 4.

Computational form chosen for a giant danio. Wake separation lines are shown for the dorsal fin, anal fin and caudal fin separation scheme.

Fig. 4.

Computational form chosen for a giant danio. Wake separation lines are shown for the dorsal fin, anal fin and caudal fin separation scheme.

where z(x) is the vertical coordinate along the length of the body in the o,x,y,z coordinate system, and p(x) is the vertical coordinate of the mean line along the length of the body in that coordinate system. The caudal fin is assumed to have chordwise sections of NACA 0016 shape, to allow for efficient resolution of leading-edge suction forces. The caudal fin leading-edge and trailing-edge profiles for the semi-span are also determined through a curve-fitting technique, and are given simply by:
where x(z)LE and x(z)TE are the longitudinal coordinate of the leading and trailing edges, respectively, along the span of the fin, and where 0⩽z⩽0.15. The leading edge of the tail at midspan intersects the posterior end of the body’s contraction region, or the region of the body which narrows in cross-sectional area anterior to the caudal peduncle. The entire body length is then scaled to L=1. The smaller dorsal and anal fins are attached to the body along the midbody profile axes of the ventral and dorsal contraction regions anterior to the caudal penducle, and they are modeled as zero-thickness vortex-lattice lifting surfaces. These fins remain attached to the body along the same axes but are allowed to flex with the prescribed motion of the body. The trailing edge of the caudal fin is assumed to be the largest wake-producing edge on the fish in the wake-shedding scheme. Additionally, in order to model the effects of upstream shed vorticity on the flow into the caudal fin, the downstream wake structure and the total force on the fish, the thin dorsal fin and anal fin are assumed each to shed a vortex wake from the trailing edge. The separation lines of the wake sheets are shown in Fig. 4.

The numerical grid in the region of the caudal peduncle, where the body and tail intersect, is artificially smoothed. This simplification allows rapid repanelization of the structured grid and ease of relative panel motion. The differences in the peduncle contraction region between the numerical form and the actual giant danio are assumed to have a small effect on the hydrodynamic similitude, as this region sustains a small hydrodynamic load and serves to transmit force from the oscillating tail to the center of mass of the body (Lighthill, 1960, 1975). Additional geometric differences between the numerical body representation and the live giant danio include the zero-thickness representation of the anal and dorsal fins and the omission of the smaller ventral and pectoral fins. The smaller ventral and pectoral fins are assumed to have minimal effects on propulsion and are neglected in this analysis. The roles of the anal and dorsal fins in propulsion are investigated by implementation of the wake-shedding scheme.

Fig. 5 presents a case of typical straight-line swimming, revealing the nonlinear wake sheet deformations and interactions. Clean separation of the main wake from the trailing edge of the caudal fin is shown, and the wake rolls up behind the tail, forming a reverse Kármán vortex street structure and thrust jet, under the self-influence of the shed vorticity and also the body perturbation velocity influence. The thrust jet is composed of connected vortex-ring-like structures, which are evident from the concentric contours of dipole strength on the deforming wake surface. The wakes shed from the vortex-lattice lifting surfaces of the thin dorsal and anal fins are also evident, with strength which oscillates due to the unsteady fin undulations caused by the traveling backbone wave. While the thin fin trailing edges are not entirely vertical, they are close to the leading edge of the caudal fin and removed from the main body, which reduces the roll-up effects caused by the influence of oscillating body perturbation velocities and the self-influence of the shed vorticity. Thus, the interaction of the vortex sheets shed from the thin fins with the caudal fin wake occurs before the thin fin wakes have significantly tightened into vortex lines. These unrolled thin fin wakes may then affect more of the flow around the leading edge of the caudal fin than if they had an opportunity to deform significantly, increasing the likelihood of energy recapture by the caudal fin through increased leading-edge suction.

Fig. 5.

Effect of upstream-vorticity shedding on the wake structure behind the computational giant danio geometry during straight-line swimming simulation. Wake sheets are represented by deforming, ungridded surfaces, with superimposed contour lines of wake dipole strength (range −50 s−1 to 50 s−1) on the deforming wake surfaces.

Fig. 5.

Effect of upstream-vorticity shedding on the wake structure behind the computational giant danio geometry during straight-line swimming simulation. Wake sheets are represented by deforming, ungridded surfaces, with superimposed contour lines of wake dipole strength (range −50 s−1 to 50 s−1) on the deforming wake surfaces.

Imposed motion description

A kinematic history of giant danio straight-line swimming motions was obtained from experimental sequences of images taken over three swimming cycles using DPIV, such as that shown in Fig. 3. From this information, we derived an analytical representation of the swimming motions. Giant danio swim using carangiform motion. Carangiform swimmers possess tapering tails which terminate in a well-defined caudal fin. Carangiform motion is characterized by a smooth, amplitude-modulated traveling wave moving along the length of the fish backbone with a phase speed cp=ω/kw, which is usually different from the swimming speed U. In this description, kw=2π/λ is the wavenumber, for a wavelength of the backbone perturbation λ, and ω is the circular frequency of oscillation. The transverse amplitude of the backbone motion a(x) increases gradually along its length, such that the primary propulsive motions are confined to the tail region.

On the basis of experimental measurements of the giant danio using DPIV, discretized from Fig. 3, the analytical description of the motion was curve-fitted to be purely sinusoidal and to consist of a smooth amplitude-modulated traveling wave along the body length with constant phase speed cp=ω/kw for a constant swimming speed U. The imposed transverse motion y(x,t), with x measured from the nose, has the form:
where a(x) is the amplitude envelope, given the form of a quadratic function:
where c1 and c2 are adjustable coefficients, as described by Barrett et al. (1999), and are chosen to achieve a specific shape of the amplitude envelope a(x) along the length of the body and a specific value of the double amplitude of motion A at the tail. The proper frequency scaling of data is thus achieved by maintenance of a constant non-dimensional Strouhal number St, based on the wake Strouhal law (Triantafyllou et al., 1991, 1993) observed in live fish, where St is given by:
where f is the frequency of oscillation and A is the total mean lateral excursion of the tail fin. These variables, obtained from the discretizations of the DPIV data of Fig. 3, are used to develop the imposed motion description for the computational inviscid model. We chose to impose the measured motion because the basic purpose is to corroborate numerical predictions of the detailed flow against experimental results. The numerical results are then used to investigate details of the flow development that cannot be easily observed in the experiment.

Flow profile comparison

Our computational geometry is employed to investigate the straight-line swimming of the giant danio as obtained during DPIV experiments. The kinematic variables that prescribe the motion are those calculated from experimental DPIV images for the sequence shown in Fig. 3, specifically: swimming speed U=1.1 L s−1; tail-beat frequency f=3.3 Hz; tail-tip double amplitude A=0.16L; backbone wavelength λ=1.1L; Strouhal number St=0.45; phase angle between pitch and heave of the tail motion ϕ=95 °; and tail angle of attack α=6 °.

Focusing on the midbody plane flow dynamics, Figs 6 and 7 highlight many prominent in-plane flow features which develop during the course of a straight-line swimming period. Fig. 6 illustrates the midbody plane velocity vectors and streamlines at eight discrete intervals over the swimming cycle. Velocity vectors densely grid the plane, scaled in size by the magnitude of the local velocity, so that far from the fish they may appear as points. In-plane streamlines are superimposed on the velocity vector grid to clarify the direction and the structure of the flow perturbations. Although the streamlines in this representation are continuous, the magnitude of the fluid velocity decreases greatly in this reference frame as the distance from the fish body increases. While the velocity vectors in this plane, in general, may have three-dimensional vertical components, careful analysis reveals that the flow is strongly two-dimensional in most regions along the body at this depth for all separation schemes and body geometries. While the flow around the top and bottom of the fish body shows strong three-dimensional behavior, sectional plane profiles along the body length show that the flow in the midplane is entirely longitudinal and has no vertical components. The complexity of the flow at other depths is the subject of continuing investigation.

Fig. 6.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles for one period of the fish swimming motion viewed from above. In-plane velocity streamlines (black) are superimposed on a dense velocity vector grid. Velocity vectors are scaled in size by velocity magnitude and scaled in color by vertical vorticity (ωz) contours (range −10 s−1 to 10 s−1). Red vorticity indicates clockwise rotation, blue vorticity indicates counterclockwise rotation, green regions are irrotational. The sequence is shown in A to H at intervals of T/8, where T is the swimming cycle period.

Fig. 6.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles for one period of the fish swimming motion viewed from above. In-plane velocity streamlines (black) are superimposed on a dense velocity vector grid. Velocity vectors are scaled in size by velocity magnitude and scaled in color by vertical vorticity (ωz) contours (range −10 s−1 to 10 s−1). Red vorticity indicates clockwise rotation, blue vorticity indicates counterclockwise rotation, green regions are irrotational. The sequence is shown in A to H at intervals of T/8, where T is the swimming cycle period.

Fig. 7.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles for one period of the fish swimming motion viewed from above. Dynamic pressure coefficient contours are shown (range 0 to 2). Blue regions are high pressure, and red regions are low pressure, clearly revealing the wake thrust jet. Numerical desingularization of the rolled-up wake structures allows the vortex cores bounding the thrust jet to be easily identified. The sequence is shown in A to H at intervals of T/8, where T is the swimming cycle period.

Fig. 7.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles for one period of the fish swimming motion viewed from above. Dynamic pressure coefficient contours are shown (range 0 to 2). Blue regions are high pressure, and red regions are low pressure, clearly revealing the wake thrust jet. Numerical desingularization of the rolled-up wake structures allows the vortex cores bounding the thrust jet to be easily identified. The sequence is shown in A to H at intervals of T/8, where T is the swimming cycle period.

The color contours of the velocity vectors are correlated to the sign and magnitude of the vertical vorticity component ωz. The red vorticity indicates a clockwise or positive circulation, and the blue vorticity indicates a counterclockwise or negative circulation. The green areas are irrotational. These vorticity contours compare favorably with the DPIV experimental results shown in Fig. 3, with similar wake strength and developing wake dynamics. The formation of a steady thrust jet is evident through the unsteady dynamics of the oscillating caudal fin, continuously shedding and manipulating wake vorticity of oscillating strength. Streamline trajectories are determined using a fourth-order Runge–Kutta spatial integration with linear interpolation between grid points. The absence of small-scale turbulent fluctuations in the numerical representation of the fluid dynamics obviates the need for higher-order integration and interpolation schemes. The rectangular grid relevant length is dL=0.01 based on the body length L=1.

Fig. 7 illustrates the thrust jet region and the near-body flow dynamics by showing the dynamic pressure contours at midbody depth, at regular intervals over the course of the swimming cycle. The dynamic pressure contours reveal low-pressure regions formed upstream, along the midsection region and contraction region of the fish body length, which are then manipulated by the oscillation of the afterbody. Specifically, the controlled actuation of the caudal fin intercepts these low-pressure regions as they progress posteriorly in the local body frame of reference and pass into the region on the inside of the maximum lateral excursion of the caudal fin. As the caudal fin is swept to the other side through the low-pressure region, the separation from the sharp trailing edge of the caudal fin contributes to the formation of a reverse Kármán vortex street. This vorticity control and production interaction mechanism contributes to the efficient production of the thrust jet comprising the wake of the straight-line swimming motions, as is evident from the contours of dynamic pressure or momentum in the fluid (Fig. 7).

Turning of a giant danio

Fish are known to have outstanding capabilities for fast-starting and maneuvering. Fish can turn through 180 ° on a radius considerably less than their body length, whereas man-made underwater vehicles require several body lengths to execute a similar turn. Experimental measurements of live fish maneuvering and fast-starting can be found in Weihs (1972), Blake (1983), Harper and Blake (1990) and Domenici and Blake (1997). A thorough kinematic analysis of unsteady turning and maneuvering motions can be found in Videler (1993). In the present study, we employ our numerical method to simulate transient fish-like motions, such as turning and maneuvering, and then compare the simulation results with experimentally observed fish turning motions. By examining the near-body flow and the wake produced by the turning motions of the fish, we extend the concepts of vorticity shedding and manipulation by the tail to explain fish maneuvering performance.

Experimental results

The flow patterns around a swimming giant danio performing a 60 ° turn were captured on video. Some of the velocity-field data were missing because of fish shadows. In Fig. 8, horizontal slices of the flow are shown with the fish viewed from above. The laser is located at the top of each figure. The fish of length L=8.32 cm begins the maneuver coasting in a straight line at a speed U=1.50 L s−1, having recently completed another turn; then it rapidly contorts its body into a tight ‘C’-curve and subsequently recoils to resume straight-line swimming at U=1.58 L s−1 in a direction 60 ° towards the right as viewed by the fish from the starting direction. A total of 25 images were taken during the turn, at 0.0333 s intervals, for a total of 0.8 s. The backbone mean line positions of the fish were obtained from 13 of these image graphically, at 0.0667 s intervals, by manual discretization of the backbone at each time into 20 equal arc length segments. The backbone locations were later used to analyze the trajectory of the fish for numerical simulation. The fish intersected the laser plane at mid-body depth for the duration of the turn.

Fig. 8.

Experimental velocity fields captured using digital particle image velocimetry in the wake of a turning giant danio. Six images are shown during the course of the maneuver at different times t, detailing the formation of a ‘C’-shape through local body contortion and the generation of a strong pair of counter-rotating vortices which provide the large transient force needed for turning.

Fig. 8.

Experimental velocity fields captured using digital particle image velocimetry in the wake of a turning giant danio. Six images are shown during the course of the maneuver at different times t, detailing the formation of a ‘C’-shape through local body contortion and the generation of a strong pair of counter-rotating vortices which provide the large transient force needed for turning.

The DPIV results for this maneuver are shown in Fig. 8. The region immediately to the left of the fish was in its shadow, so these data were removed. Shortly preceding the position shown in Fig. 8A, the fish executed a turn originating in the upper left-hand corner of Fig. 8A and resumed straight-line swimming. Residual vorticity from these movements remains above and to the left of the fish. In Fig. 8B,C, the fish begins to bend, moving its head towards the new swimming direction while the tail moves in the opposite sense, forming a characteristic ‘C’-shape. The fluid motion follows that of the contorting body, moving directly towards the fish on the concave side of the body and directly away from the fish on the convex side of the body, at the points of maximum curvature.

In Fig. 8D, we see the flow organize into two circular-like flows, one centered at the tail and one near the head, as the tightening of the body into a ‘C’-shape nears completion. In Fig. 8E, the tail has begun to move towards the right of the viewed image, shedding in the wake the previously formed counterclockwise vortex. At this instant, the effect of bound vorticity near the head is obscured by reflection of light from the body. The next several frames are not presented because these body reflections obscured a large portion of the flow. However, analysis of the video record shows that during the stroke of the tail, sweeping downwards in the viewed image as the fish body heads to the left, the counterclockwise vortex sheds into the wake. During the next stroke of the tail, which sweeps back upwards in the viewed image as the fish continues moving to the left, the clockwise vortex initially associated with the front region of the body moves posteriorly and sheds in the wake.

The net result is shown in Fig. 8F: a strong vortex pair forms which provides a jet directed slightly downwards and to the right. The vortices in this pair are packets of counter-rotating large-scale wake vorticity, and Fig. 9 shows the vorticity contours at the same instant. The asymmetric shapes of the contours of the two vortices that make up the turning jet are due to the unsteady, non-periodic body motions and velocity. Additional vorticity shown in this figure, but not labeled with contour strength values, was created by the fish before the initiation of the turning maneuver. The entire turning sequence, from a straight-line coasting trajectory following a previous turn to a steady swimming trajectory 60 ° to the right as viewed by the fish from the starting direction, takes slightly more than 0.5 s. The vortices comprising the jet have average non-dimensional circulation Гe*e/LU=0.43, which is 42 % greater than the typical wake vortex strength for steady swimming at similar speeds. The core radius of the jet vortices is more than double that of those produced in straight swimming. The jet is 0.34L wide, with maximum jet velocity of ujU oriented approximately 60 ° to the left of the initial starting position as seen by the fish. Interestingly, the fish velocity after the turn is roughly the same as the jet velocity. Also, the direction of the jet is roughly in the direction required by horizontal momentum balance. Three-dimensional effects are significant, and the vigorous motion of the fish somewhat disrupts our two-dimensional planar results. After completing the turn, the fish began to swim steadily at U=1.5 L s−1.

Fig. 9.

Vorticity contours in the wake of a turning fish at the instant shown in Fig. 8F, time t=0.534 s. Solid lines denote positive vorticity, broken lines negative vorticity. Contour values are as follows: 6.7 s−1 to 16.9 s−1 and −6.7 s−1 to −20.0 s−1 in 1.03 s−1 increments. Unlabeled vorticity contours are remnants of prior turning motions and the cessation of straight-line swimming motions.

Fig. 9.

Vorticity contours in the wake of a turning fish at the instant shown in Fig. 8F, time t=0.534 s. Solid lines denote positive vorticity, broken lines negative vorticity. Contour values are as follows: 6.7 s−1 to 16.9 s−1 and −6.7 s−1 to −20.0 s−1 in 1.03 s−1 increments. Unlabeled vorticity contours are remnants of prior turning motions and the cessation of straight-line swimming motions.

Numerical modeling

We impose on the fish model the body motions observed in live fish, including the movements of the tail as well as the trajectory of the fish. The shape and position of the fish were only known at 13 time steps during the maneuver; an accurate wake picture could not be produced if these 13 positions were the sole simulation input with such a large time differential between body positions. An interpolative scheme was therefore necessary to calculate the location of the fish at intermediate time steps, from which the fish turn could be simulated.

Fig. 10 details the coordinate systems used. The O,X,Y,Z coordinate system is global and fixed. The o,x,y,z coordinate system is a coordinate system whose origin is fixed at the nose of the fish. The x-axis always passes through the tail of the fish. Thus, the flexing motion of the fish can be described in a local coordinate system, while its trajectory can be described in the global coordinate system. The z coordinates of the fish body depth were assumed to be invariant as the plane of the laser sheet intersected the fish at midbody depth for the duration of the turning motion.

Fig. 10.

Coordinate system conventions for the analysis of maneuvering (top view). Local unsteady body motions are described with respect to the body-fixed o,x,y,z coordinate system, and the body trajectory is described with respect to an inertial reference frame O,X,Y,Z. U, swimming speed or instantaneous translational velocity; Sb, body surface in the computational domain; Sw, wake surface in the computational domain; θ and θand rotational velocity, respectively, of the local coordinate system o,x,y,z about the z axis with respect to O,X,Y,Z.

Fig. 10.

Coordinate system conventions for the analysis of maneuvering (top view). Local unsteady body motions are described with respect to the body-fixed o,x,y,z coordinate system, and the body trajectory is described with respect to an inertial reference frame O,X,Y,Z. U, swimming speed or instantaneous translational velocity; Sb, body surface in the computational domain; Sw, wake surface in the computational domain; θ and θand rotational velocity, respectively, of the local coordinate system o,x,y,z about the z axis with respect to O,X,Y,Z.

The 13 discretized fish backbone images were then analyzed as follows. First, the position of the nose in the global coordinate system was recorded as a function of time, serving as the origin of the o,x,y,z coordinate system. The angle θ that the o,x,y,z coordinate system made with the global coordinate system was also recorded as a function of time, by constructing the x-axis through the fish tail at each time interval. The experimental data of fish head and tail locations during the turning maneuver can be seen in Fig. 11A.

Fig. 11.

Trajectory of a maneuvering fish. (A) Experimental data: circles and crosses indicate head and tail positions, respectively; the curves represent a few of the body backbone positions. The time interval between sequential plotted head/tail positions is dt=0.0333 s. The starting position is vertical with the head pointing downwards in the view shown. The fish contorts its body into a tight ‘C’-shape, then recoils to resume swimming to its right. (B) Simulation input: the fish mean backbone trajectory is shown at intervals of dt=0.02 s, corresponding to every other time step in the numerical simulation. Backbone lines are shown in bold every 0.1 s for clarity.

Fig. 11.

Trajectory of a maneuvering fish. (A) Experimental data: circles and crosses indicate head and tail positions, respectively; the curves represent a few of the body backbone positions. The time interval between sequential plotted head/tail positions is dt=0.0333 s. The starting position is vertical with the head pointing downwards in the view shown. The fish contorts its body into a tight ‘C’-shape, then recoils to resume swimming to its right. (B) Simulation input: the fish mean backbone trajectory is shown at intervals of dt=0.02 s, corresponding to every other time step in the numerical simulation. Backbone lines are shown in bold every 0.1 s for clarity.

A periodic motion was assumed for the local flapping motion of the fish. As the fish had a fairly uncambered mean line shape at both the first time step and the last time step of the turn, it was assumed that the local shape of the fish mean line at the extrapolated time of 0.8667 s would be equal to the original mean line shape. In this way, the local x and y coordinates of the mean line, as well as the thickness distribution h, could be expanded as Fourier sine series as functions of time. Additionally, as the local y coordinate and the thickness h are always zero at the nose and tail, these could be expanded as Fourier sine series as functions of backbone arc length, while the local x coordinate could be expanded as a linear function plus a Fourier sine series as a function of backbone arc length. Thus, the local behavior of the backbone can be found at any time during the simulation.

The orientation of the local coordinate system within the global frame of reference is similarly assessed. Expansion of the global X, Y and θ coordinates in terms of Fourier series in time yields the trajectory and inclination of the local coordinate system within the global reference frame for any time during the simulation. Details of the complete analytical description of the unsteady motion can be found in Appendix B. From this analytical description of the motion, we can produce a simulation with an arbitrary number of time steps, and the difficulties in accurately reproducing the body motion and the resulting wake are alleviated by having a sufficient number of time steps in the simulation. As an example, Fig. 11B shows the backbone trajectory and flexing behavior of the giant danio over the entire simulation employing 50 intermediate time steps of information.

The viscous effects of the body are again assumed to be confined to a thin boundary layer near the body and in a small wake described by shear layers. The Reynolds numbers Re of the giant danio before and after the turning motion are approximately Re=10 400 and Re=11 000, respectively. At these values, we would expect the boundary layer to be laminar or within a transition region, as illustrated by experimental data for the drag coefficients of axisymmetric streamlined bodies with varying L/d ratios (where d is the diameter of the streamlined body) found in Fig. 6.22 of Hoerner (1965). It is reasonable to assume that the near-body viscous effects on the body are small in comparison with the effects of unsteady vorticity generated by separation, such as that occurring at the sharp trailing edges of the body which generate wakes. Again, the tail is assumed to generate the largest wake, and a primary separation line is chosen to be the trailing edge of the tail. The large dorsal and anal fins are also included in the course of our investigation, to examine the influence of upstream-generated vorticity on the flow around the tail, the overall wake structure and the unsteady body forces generated.

The body shape can be described at each time step, and the imposed normal velocities at each panel midpoint are known; hence, the boundary integral equation can be solved to find the solution of the unsteady boundary value problem at each time step, as described above.

Numerical results and comparisons

The turning motion of the giant danio is simulated using the method described above. It was observed in the DPIV experiments that the near-tail flow dynamics are affected by the motions of the dorsal and anal fins during the maneuver. These interactions are less apparent during straight-line swimming, in which the dorsal and anal fin lateral motions are smaller. As a result, the accurate modeling of the fish geometry during the turning maneuver was deemed crucial to resolving the flow dynamics around the tail and around the dorsal and anal fins with their large excursions.

The simulation time step size dt=0.01 s was chosen to be small enough to describe the details of the motion completely and to resolve wake-body dynamics and wake self-interactions. Fourier expansions of the mean line shapes in the local coordinate system reference frame employ 15 temporal modes and 10 spatial modes. Similar expansions of the local coordinate system trajectories within the global coordinate system employ 10 temporal modes. Numerical desingularization radii for the wake and body were chosen to be δw=0.02 and δb=0.025, respectively, based on a body length of L=1.0, selected through a convergence process ensuring optimal stability of the numerical solution.

To eliminate the influence of any starting vorticity shed by the caudal, dorsal and anal fins during the impulsive start of the numerical simulation, a ramping-up time interval was added to the initial portion of the simulation. The initial conditions of the simulation are such that the fish mean line is locally perturbed to assume the shape of the giant danio at the start of the turning maneuver. The global position of the geometry is such that it translates at a constant velocity, equal to the velocity at the start of the turning maneuver. At the conclusion of the ramping-up time interval, the position, velocity and shape of the geometry are identical to those of the giant danio at the initial frame of the experimental DPIV data. The total simulation length of 160 time steps thus modeled the 0.800 s turning maneuver, with 0.800 s of initial coasting motion before the turn was initiated to eliminate unwanted starting vorticity dynamics.

The results of the simulation are shown in Figs 12 and 13. These are taken at 0.100 s intervals, shown from t=−0.100 s before the start of the turning maneuver until the turn is nearly completed at t=0.600 s. These frames are representative of time steps ts=70 to ts=140 at 10-time-step intervals.

Fig. 12.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles during the unsteady turning motion, viewed from above. In-plane velocity streamlines (black) are superimposed on a dense velocity vector grid. Velocity vectors are scaled in size by velocity magnitude and scaled in color by vertical vorticity (ωz) contours (range −20 s−1 to 20 s−1). Red vorticity indicates clockwise rotation, blue vorticity counterclockwise rotation, green regions are irrotational. The sequence is shown in A to H at intervals dt=0.100 s, corresponding to t=−0.100 s to t=0.600 s of the original experimental recording.

Fig. 12.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles during the unsteady turning motion, viewed from above. In-plane velocity streamlines (black) are superimposed on a dense velocity vector grid. Velocity vectors are scaled in size by velocity magnitude and scaled in color by vertical vorticity (ωz) contours (range −20 s−1 to 20 s−1). Red vorticity indicates clockwise rotation, blue vorticity counterclockwise rotation, green regions are irrotational. The sequence is shown in A to H at intervals dt=0.100 s, corresponding to t=−0.100 s to t=0.600 s of the original experimental recording.

Fig. 13.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles during the unsteady turning motion viewed from above. Dynamic pressure coefficient contours are shown (range 0 to 3). Blue regions are high pressure, and red regions are low pressure, revealing the formation of a turning-assisting jet. The sequence is shown in A to H at intervals dt=0.100 s, corresponding to t=−0.100 s to t=0.600 s of the original experimental recording.

Fig. 13.

Midplane depth (z=0.5H, where H is fish body maximum total depth) flow profiles during the unsteady turning motion viewed from above. Dynamic pressure coefficient contours are shown (range 0 to 3). Blue regions are high pressure, and red regions are low pressure, revealing the formation of a turning-assisting jet. The sequence is shown in A to H at intervals dt=0.100 s, corresponding to t=−0.100 s to t=0.600 s of the original experimental recording.

The sequence of images shown in Fig. 12 details the two-dimensional flow patterns around the turning fish in a plane at midbody depth on the fish. In the initial plots of the sequence, at the extreme top of the fluid plane there is residual vorticity resulting from the starting motions in the simulation. This is caused by the translation of the giant danio geometry from the top of the frame into its starting position. The turning maneuver produces additional vorticity which is largely unaffected by the decaying starting vorticity. The sequence of images shown in Fig. 13 illustrates the evolution of dynamic pressure contours during the turn.

From both Figs 12 and 13, it can be seen that as the fish contorts itself into a tight ‘C’-shape, the flow is organized into three circular patterns, around the head, the midbody and the tail fin, respectively. As the fish begins to sweep its tail towards its left side, the tail fin sheds positive vorticity, as negative bound vorticity moves from the contraction region towards the tail. A large low-pressure region coming from the right of the body crosses over to the left, passing over the tail and eventually pairing with the previous vortex to form a strong jet, which turns the fish. The negative vorticity in the contraction region moves posteriorly to the tail during the stroke of the tail to the left, and a new low-pressure region is formed below the fish. This negative vorticity is released into the wake during the subsequent return stroke to the right, as the second low-pressure region passes over the leading edge of the tail fin. As the tail concludes its stroke to the right and passes through the second region of high fluid momentum, it recovers energy from the jet through the enhanced separation from the trailing edge of the tail, reinforcing the negative vorticity that is being shed into the wake and resulting in additional loading on the tail, which propels the body forward. The dynamics of the maneuver are shown in Fig. 14. The forces on the fish body during the turn are depicted as a function of time, with the total force decomposed into its X and Y components of the global coordinate system (Fig. 14A). In addition, the fluid force vectors acting on the fish center of mass at each time step of the simulation are shown, with the mean line curves of the fish backbone superimposed on the plot every four time steps (Fig. 14B). The forces are moderately large at the outset of the maneuver, as the fish halts the oscillatory motions of straight-line swimming and releases residual body-bound vorticity into the wake, then increase sharply as the fish contorts into a tight curve and sweeps its caudal fin to the left. The oscillatory forces gradually subside as the turn is completed and straight-line swimming commences again. It is evident from the vector time history that the forces on the body allow it to turn in a short distance relative to its length.

Fig. 14.

Force versus time history for the turning giant danio simulation. (A) X (solid line) and Y (broken line) components of the force on the fish body center of mass in a global reference frame as a function of time. (B) Force vectors acting through the center of mass shown as a function of position at every time step dt=0.01 s during the maneuver. Body mean line locations are superimposed on the vector plot every four time-step intervals.

Fig. 14.

Force versus time history for the turning giant danio simulation. (A) X (solid line) and Y (broken line) components of the force on the fish body center of mass in a global reference frame as a function of time. (B) Force vectors acting through the center of mass shown as a function of position at every time step dt=0.01 s during the maneuver. Body mean line locations are superimposed on the vector plot every four time-step intervals.

Steady straight-line swimming

Comparison between the experimental data and the numerical simulation results shows good quantitative and qualitative agreement. Wake vortex strength and placement compare well, with a thrust jet width of approximately 1.3 times the maximum body width. Maximum computed numerical vorticity strengths are within 20 % of experimental values. Simulation results reveal that the side forces are distributed evenly between the tail and the main body, while the tail sustains over 90 % of the longitudinal force. The detailed patterns shown in Fig. 15 elucidate the complex processes of bound vorticity release and thrust wake formation in the experimental results and those of the simulation. The near-body and wake streamline patterns near the trailing-edge separation line of the caudal fin show qualitatively similar features.

Fig. 15.

Close-up of the separated and near-body flow at the midplane depth near the tail of the giant danio during straight-line swimming. Magnification of near-body streamlines illustrates the effect of body-bound and free vorticity on the adjacent flow. (A) Digital particle image velocimetry experimental data showing the in-plane velocity streamlines near the trailing edge, revealing the large wake vortex and the body-bound vorticity near the trailing edge. (B) Numerical simulation results for the same instant superimpose streamlines (black) on a dense in-plane velocity vector grid. Velocity vectors are scaled in size by velocity magnitude and in color by vertical vorticity contours (ωz) (range −10 s−1 to 10 s−1). Red vorticity indicates clockwise rotation, blue vorticity indicates counterclockwise rotation, green regions are irrotational.

Fig. 15.

Close-up of the separated and near-body flow at the midplane depth near the tail of the giant danio during straight-line swimming. Magnification of near-body streamlines illustrates the effect of body-bound and free vorticity on the adjacent flow. (A) Digital particle image velocimetry experimental data showing the in-plane velocity streamlines near the trailing edge, revealing the large wake vortex and the body-bound vorticity near the trailing edge. (B) Numerical simulation results for the same instant superimpose streamlines (black) on a dense in-plane velocity vector grid. Velocity vectors are scaled in size by velocity magnitude and in color by vertical vorticity contours (ωz) (range −10 s−1 to 10 s−1). Red vorticity indicates clockwise rotation, blue vorticity indicates counterclockwise rotation, green regions are irrotational.

Fig. 16 summarizes the development and control of vorticity by the flexible body undulation during the swimming cycle. This model simplifies the complex three-dimensional nature of the near-body flow, but accurately portrays the dominant features of the mid-depth plane flow and the principal mechanisms involved in manipulation of the large-scale two-dimensional vorticity features in the wake. In Fig. 16A, clockwise (negative) vorticity is shed into the wake. Along the contraction region of the body, counterclockwise (positive) vorticity is fully developed, due to the local undulation of the body and the coordinated upward stroke of the caudal peduncle as seen by the reader. Fig. 16B reveals positive bound vorticity moving towards the tail tip during the upward stroke of the tail. New negative bound vorticity develops in the midbody region of the fish length, due to the downward midbody motion. Fig. 16C illustrates the shedding of the positive bound vorticity into the wake, while negative vorticity is fully developed in the contraction region, owing to the local body contortion and coordinated downward caudal peduncle motion. In Fig. 16D, the cycle is complete and recommences. The alternating-sign free vortices in the wake form a reverse Kármán street, resulting in the development of a thrust jet. Optimal propulsion is achieved through proper selection of the frequency and spacing of the wake vortices. The process of the formation of wake vortices through the release of body-generated vorticity and subsequent tail manipulation as revealed in Figs 6, 7, 15 and 16 is an important result of the present paper, extending the analysis and conclusions of Triantafyllou et al. (1993).

Fig. 16.

A model of the development and release of body-bound and tail-generated vorticity during one swimming cycle of steady, straight-line swimming. A top view of the swimming fish is shown.

Fig. 16.

A model of the development and release of body-bound and tail-generated vorticity during one swimming cycle of steady, straight-line swimming. A top view of the swimming fish is shown.

Fig. 17.

A model summarizing vorticity control mechanisms in a giant danio executing a 60 ° turn to its right. A top view of the midplane of the swimming fish is shown.

Fig. 17.

A model summarizing vorticity control mechanisms in a giant danio executing a 60 ° turn to its right. A top view of the midplane of the swimming fish is shown.

Turning

Fig. 17 summarizes the development and control of vorticity by the flexible body deformation during the turning maneuver. This model simplifies the complex three-dimensional nature of the near-body flow, but accurately portrays the dominant features of the mid-depth plane flow and illustrates the principal mechanisms involved in manipulation of the large-scale two-dimensional vorticity. In Fig. 17A, the fish ceases its straight-line swimming undulations and releases body-bound vorticity associated with straight-line swimming motions into the wake (not shown). In Fig. 17B, the fish initiates its backbone curve. In the center of the body, a pair of oppositely signed bound vortices develop, owing to the midbody translation to the left and the opposite motion of the head and tail. In Fig. 17C, the local contortion of the backbone into a ‘C’-shape is complete, and the pair of body-generated, oppositely signed bound vortices separate as one of the vortices moves closer to the tail. Straightening of the body in a wave-like motion commences in Fig. 17D, which releases counterclockwise vorticity into the wake through manipulation by the caudal fin. As straightening continues, the clockwise bound vorticity travels down the length of the body towards the tail and is released into the wake. Initiation of straight-line swimming motions completes the release of clockwise vorticity into the wake, as shown in Fig. 17E, which then pairs with the counterclockwise vorticity to form a thrust jet in the wake. The rapid formation of this thrust jet accomplishes the turning of the fish. It is remarkable that no uncontrolled vorticity appears to be shed during the turning process.

Thus, control of bound vorticity and wake formation by the body and the tail fin for thrust vectoring to perform turning maneuvers is achieved through large localized backbone and tail actuation. This mechanism of vorticity control allows for the efficient manipulation of near-body flows and empowers the generation of large, short-duration maneuvering forces.

Conclusions

We have presented detailed experimental measurements of the flow around a live, naturally swimming fish and performed numerical simulations on a fish model performing the same motions. Using DPIV, we have clarified the principal propulsive mechanisms used by the fish. Three-dimensional unsteady flexible-body numerical simulations are corroborated against the experimental observations and then used to elucidate further details of the formation of the thrust jet. Specifically, fish control body-generated vorticity through body flexure and active manipulation by the caudal fin. Our results indicate that the structure of the jet wake is obtained through the regulated release of body-generated vorticity and its constructive interaction with vorticity formed and shed by the oscillating caudal fin.

We also presented detailed results for the near-body flow dynamics around a live fish executing an unsteady, rapid turning motion. Through precise body actuation, the fish regulates the formation and controlled release of body-generated vorticity, resulting in the production of a pair of counter-rotating vortices and, hence, a thrust jet. Thus, the fish is able to redirect its generated thrust rapidly and achieve its desired trajectory.

It has long been known that fish swimming in a straight line produce an unsteady wake in the form of a reverse Kármán street. However, it has generally been assumed that the wake vortices originate from the action of the tail as a lifting surface. Our results indicate that the tail fin has a secondary role in the generation of wake vortices. Vorticity is generated well upstream of the tail by the undulations of the body. Upon arriving at the tail, body-generated vorticity is favorably affected by the motion of the tail: it is reinforced by the tail, while vorticity shed at the trailing edge of the tail merges with and amplifies each body-generated vortex. The undulation of the fish body and the caudal fin motion are synchronized such that vorticity formation and evolution are actively controlled for both straight-line swimming and maneuvering.

Appendix

A. Unsteady panel method solution algorithm

From the discretized version of Green’s theorem at any time t, the body perturbation velocity potential at each panel collocation point can be found in terms of the perturbation potentials at all of the other panels:
where i is an index of observation panel number and j is an index of field panel number in a discrete representation of boundary integral formulation, k is the total number of body panels, and where:
where Pij is the source distribution influence of panel j on panel i, Qij is the dipole distribution influence of panel j on panel i and rij is the magnitude of the distance between the collocation point of panel i and the collocation point of panel j.
A linear system of equations results for the body perturbation velocity potential over each panel. However, the strengths of the most recently shed wake panels to the trailing edge are unknown at this point, and the unique solution of the newly shed wake strength cannot be determined using the above system of equations 20–23 alone. This unknown can be absorbed into the system of equations for by the introduction of a discrete form of the Kutta boundary condition (equation 2):
where |top panel and |bottom panel are the body perturbation velocity potentials on trailing-edge panels at on the top and the bottom of the lifting surface, respectively. Therefore, we can describe the perturbation velocity potential of the new wake panels in terms of the body perturbation velocity potential The normal velocity on each body panel is composed of contributions from the imposed velocity of the body and from the velocities induced by the discrete wake panels, each with piecewise constant perturbation velocity potentials We can write this boundary condition in its discrete form:
where nw is the total number of wake panels of known strength, and p is an index of these wake panels. We then rewrite the linear system (equation 20) for :
where the influence coefficients Qiw are integrals over the new wake panels Sw of unknown strength adjacent to the trailing edges with collocation point :
where riw is the magnitude of the distance between the collocation point of panel i and the collocation point of the wake panel adjacent to the trailing edge of unknown strength, and ∂ nw is the partial derivative in the direction of the wake surface normal vector.

The discrete forms of the Kutta condition (equation 24) and the kinematic body boundary condition (equation 25) can then be employed to arrive at a simple system of linear equations for at any given time t, expressed as equation 4.

Appendix

B. Kinematic modeling of unsteady maneuvering

Given the backbone shape and body thickness distribution at a discrete number of time steps for arbitrary fish swimming motions, we can determine the details of the local backbone undulations and the global body trajectory as functions of time using a series expansion technique. This allows the generation of the smooth trajectory and body shape needed for simulation.

At each time step, the local x and y coordinates of the backbone mean line and the body thickness distribution h can be expanded as Fourier sine series in backbone arc length s, where tp is the total number of arc length segments:
where Lo is the original arc length of the backbone, and An, Bn and Cn are the Fourier spatial expansion coefficients of lateral motion, longitudinal motion and unsteady body thickness distribution, respectively. The linear function added to the x(s) expansion is not necessary for the y(s) and h(s) expansions, which are zero-valued at each end of the backbone in the local coordinate system. In this way, the complex local shapes of the mean line are always accommodated, whereas a functional description such as y=f(x) for the expansion may suffer from multi-valued violations at discrete times.
Given these spatial expansions at each time step, we wish to link them through an expansion in time, where tt is the total number of original time steps. For each mode in the spatial expansion of x(s), y(s) and h(s), a periodic shape in time exists if we extrapolate to one more time step where all variables return to their original values. Thus, the coefficients can be expanded as functions of time:
where Anm is a Fourier coefficient of the lateral unsteady motion spatial and temporal expansion. The equations that describe the flapping motion of the fish in the local coordinate system as functions of backbone arc length and time then become:
where N and M are the number of Fourier sine modes in space and time, respectively, Bnm and Cnm are the Fourier coefficients of the spatial and temporal expansions of the longitudinal unsteady motion and the unsteady body thickness distribution, respectively, and Lm is the Fourier coefficient of the temporal expansion of the backbone arc length.
Given the motions of the fish in the local coordinate system, the orientation of the local coordinate system within the global frame of reference is found. Although not periodic, the location and inclination of the local coordinate system origin in the global coordinate frame are determined by a related method. The global coordinates X, Y and θ are expanded in terms of a Fourier sine series with an arbitrary number P of temporal modes. A linear term is subtracted from each expansion to force the coordinates to a value of zero at either end of the simulation period. Superposition of the linear term and the Fourier sine series expansion yields the trajectory and inclination of the local coordinate system within the global reference frame:
where the coefficients Cθp, CXp and CYp are found in the usual manner using the discretized mean line global location and inclination information. Simulations of the turning motion described in the present study of turning were found to result in optimal reproduction of the experimentally observed motions when at least 10 Fourier sine modes were employed for both temporal and spatial expansion of the local mean line behavior and for the temporal expansion of the local coordinate system global trajectory.
Given the position of the mean line (backbone) at a large number of intermediate time steps, the full three-dimensional body can be described at each time step. As the backbone deforms, the sectional planes are assumed to remain orthogonal to the backbone, and it is also assumed that the body volume is fixed. Therefore, an additional boundary condition must be applied to this numerical method, that the volume of the body must be constant. Thus, the body is scaled at each time so that the integral of the total normal velocity around the body must be zero:
     
  • A

    total lateral excursion of the tail at the caudal peduncle

  •  
  • An

    Fourier coefficients of lateral unsteady motion spatial expansion

  •  
  • Anm

    Fourier coefficients of lateral unsteady motion spatial and temporal expansion

  •  
  • aspect ratio of three-dimensional lifting surface

  •  
  • a(x)

    amplitude of the backbone motion along the length of the body x in the o,x,y,z coordinate system

  •  
  • Bn

    Fourier coefficients of longitudinal unsteady motion spatial expansion

  •  
  • Bnm

    Fourier coefficients of longitudinal unsteady motion spatial and temporal expansion

  •  
  • Cn

    Fourier coefficients of unsteady body thickness distribution spatial expansion

  •  
  • Cnm

    Fourier coefficients of unsteady body thickness distribution spatial and temporal expansion

  •  
  • CP

    pressure coefficient

  •  
  • CXp

    Fourier coefficients of the X-coordinate temporal expansion for unsteady body motions

  •  
  • CYp

    Fourier coefficients of the Y-coordinate temporal expansion for unsteady body motions

  •  
  • Cθp

    Fourier coefficients of the θ-coordinate temporal expansion for unsteady body motions

  •  
  • cp

    phase speed of propagating backbone wave

  •  
  • c1

    linear backbone wave amplitude coefficient

  •  
  • c2

    quadratic backbone wave amplitude coefficient

  •  
  • d

    diameter of streamlined body from Hoerner (1965) 

  •  
  • dl

    elemental unit of length along a vortex segment for Biot–Savart integration

  •  
  • dL

    velocity grid length spacing based on body length L

  •  
  • tangent vector for circulation contour integration

  •  
  • ds

    elemental unit of surface area

  •  
  • dt

    experimental DPIV image interval or computational time step

  •  
  • Fb

    force on the thick body surfaces

  •  
  • Fb,thin

    force on the vortex-lattice thin body surfaces

  •  
  • f

    frequency of tail oscillation (Hz)

  •  
  • H

    fish body maximum total depth

  •  
  • h

    fish body thickness distribution at midbody depth along the length of the body in o,x,y,z coordinate system

  •  
  • i

    index of observation panel number in discrete representation of boundary integral formulation

  •  
  • j

    index of field panel number in discrete representation of boundary integral formulation

  •  
  • k

    number of body panels

  •  
  • kw

    wavenumber of the backbone wave

  •  
  • L

    fish body length

  •  
  • Lm

    Fourier coefficients of the temporal expansion of the backbone arc length

  •  
  • Lo

    original arc length of the backbone in unsteady motion Fourier expansions

  •  
  • Lb

    component of total body force Fb in the lateral direction, perpendicular to the direction of swimming

  •  
  • M

    number of temporal modes in the Fourier expansions for unsteady body motion local trajectories

  •  
  • m

    index of unsteady motion temporal Fourier expansion

  •  
  • body outward normal vector

  •  
  • nb

    derivative in the direction of the body outward normal vector

  •  
  • vortex-lattice thin body surface normal vector

  •  
  • derivative in the direction of the thin fin outward normal vector

  •  
  • nw

    number of wake panels

  •  
  • nw

    derivative in the direction of the wake surface normal vector

  •  
  • O,X,Y,Z

    inertial reference frame

  •  
  • o,x,y,z P

    local body-fixed coordinate system

  •  
  • P

    number of temporal modes in the Fourier expansions for unsteady body motion global trajectories

  •  
  • column vector of panel normal velocities for source-dipole potential panel distribution

  •  
  • Pb

    power transmitted to the fluid by the body

  •  
  • temporal mean of Pb over an integral number of steady cycles

  •  
  • Pij

    source distribution influence of panel j on panel i

  •  
  • P1

    partial column vector of panel normal velocities for thick body panel distribution

  •  
  • P2

    partial column vector of panel normal velocities for vortex-lattice thin fin panel distribution

  •  
  • p(x)

    vertical coordinate of the mean line along length of body in o,x,y,z coordinate system

  •  
  • p

    index of wake panels for panel normal velocity calculation

  •  
  • matrix of influence coefficients for source-dipole potential panel distribution

  •  
  • Qij

    dipole distribution influence of panel j on panel i

  •  
  • Qiw

    dipole distribution influence of new wake panel

  •  
  • Q11

    sub-matrix of influence coefficients for source-dipole potential panel distribution; thick body panels’ influence on thick body panels

  •  
  • Q12

    sub-matrix of influence coefficients for source-dipole potential panel distribution; vortex-lattice surface panels’ influence on thick body panels

  •  
  • Q21

    sub-matrix of influence coefficients for vortex-lattice potential panel distribution; thick body panels’ influence on vortex-lattice surface panels

  •  
  • Q22

    sub-matrix of influence coefficients for vortex-lattice potential panel distribution; vortex-lattice surface panels’ influence on vortex-lattice surface panels

  •  
  • Re

    Reynolds number

  •  
  • Ro

    core radius of observed wake vortices

  •  
  • r

    magnitude of distance between two points

  •  
  • vector of magnitude r from a vortex element to the field point for Biot–Savart integration

  •  
  • rij

    magnitude of distance between collocation points of panel i and panel j

  •  
  • riw

    magnitude of distance between collocation points of panel i and new wake panel w

  •  
  • S

    all bounding surfaces of computational domain

  •  
  • Sb

    all body surfaces in computational domain, subset of S

  •  
  • Sw

    all wake surfaces in computational domain, subset of S

  •  
  • Sw

    new wake panels shed adjacent to trailing edge

  •  
  • S

    far-field surface at in computational domain, subset of S

  •  
  • St

    Strouhal number

  •  
  • s

    backbone arc length coordinate for unsteady motion parametization

  •  
  • tangent vector to vortex element for Biot–Savart integration

  •  
  • T

    period of one cycle of straight-line swimming motions

  •  
  • Tb

    thrust component of total body force Fb in the direction of swimming

  •  
  • temporal mean of Tb over an integral number of steady cycles

  •  
  • TE

    trailing edge of lifting surfaces

  •  
  • t

    time

  •  
  • tp

    total number of arc length segments for unsteady motion parametization

  •  
  • ts

    time step number

  •  
  • tt

    total number of time steps of experimental data for unsteady motion parametization

  •  
  • U

    swimming speed

  •  
  • uj

    velocity of fluid in thrust jet

  •  
  • velocity field in the X,Y plane of the O,X,Y,Z coordinate system for circulation contour integration.

  •  
  • velocity of the body at point on Sb or a panel at time t

  •  
  • velocity induced by vortex element at a point with respect to O,X,Y,Z coordinate system

  •  
  • X

    longitudinal coordinate of o,x,y,z origin with respect to O,X,Y,Z

  •  
  • x

    longitudinal coordinate along length of body in o,x,y,z coordinate system

  •  
  • x(z)LE

    longitudinal coordinate of the leading edge of the tail as a function of the height z in o,x,y,z coordinate system

  •  
  • x(z)TE

    longitudinal coordinate of the trailing edge of the tail as a function of the height z in o,x,y,z coordinate system

  •  
  • three-dimensional coordinate with respect to O,X,Y,Z

  •  
  • three-dimensional coordinate with respect to O,X,Y,Z on a vortex-lattice thin body surface three-dimensional coordinate of the collocation point of the observation panel with respect to O,X,Y,Z

  •  
  • three-dimensional coordinate of the collocation point of the field panel with respect to O,X,Y,Z

  •  
  • three-dimensional coordinate with respect to O,X,Y,Z at the trailing edge of lifting surfaces

  •  
  • three-dimensional coordinate with respect to O,X,Y,Z of the collocation point of the wake panel adjacent to the trailing edge of unknown strength.

  •  
  • Y

    lateral coordinate of o,x,y,z origin with respect to O,X,Y,Z

  •  
  • y(x,t)

    transverse motion of the fish backbone as a function of time t and length along the body x in the o,x,y,z coordinate system

  •  
  • z(x)

    vertical coordinate along length of body in the o,x,y,z coordinate system

  •  
  • α

    angle of attack of the tail

  •  
  • Γ

    circulation

  •  
  • Γe

    circulation of observed wake vortices

  •  
  • Γe*

    non-dimensional circulation of observed wake vortices

  •  
  • Γv

    wake vortex element circulation strength

  •  
  • total velocity potential field at point x at time t

  •  
  • body perturbation velocity potential field at point x at time t

  •  
  • wake perturbation velocity potential field at point x at time t

  •  
  • wake perturbation potential field at the trailing edge of lifting surfaces

  •  
  • ϕ

    phase angle between the pitch and heave motion of the tail

  •  
  • ϕb

    column vector of body perturbation velocity potentials

  •  
  • ϕbb

    partial column vector of thick body perturbation velocity potentials

  •  
  • ϕbf

    partial column vector of vortex-lattice thin body perturbation velocity potentials

  •  
  • body perturbation velocity potential on a panel at at time t

  •  
  • body perturbation velocity potential on a trailing edge panel at on the top of the lifting surface at time t

  •  
  • top panel

    body perturbation velocity potential on a trailing edge panel at on the bottom of the lifting surface at time t

  •  
  • bottom panel

    wake perturbation velocity potential field on a panel at at time t

  •  
  • wake perturbation velocity potential at a panel adjacent to the trailing edge of a lifting surface

  •  
  • gradient operator

  •  
  • divergence operator

  •  
  • TE

    jump in body perturbation potential field at the trailing edge of lifting surfaces

  •  
  • δb

    body panel desingularization radius

  •  
  • δw

    wake panel desingularization radius

  •  
  • λ

    wavelength of the backbone wave

  •  
  • η

    propulsive efficiency

  •  
  • ρ

    density of the fluid

  •  
  • θ

    angular rotation of the local coordinate system o,x,y,z about z axis with respect to O,X,Y,Z

  •  
  • angular rotational velocity of the local coordinate system o,x,y,z about z axis with respect to O,X,Y,Z

  •  
  • ω

    circular frequency of oscillation of the tail

  •  
  • ωz

    vertical component of vorticity

  •  
  • three-dimensional coordinate with respect to O,X,Y,Z

The financial support of the Office of Naval Research under contract N00014-96-1-1141 monitored by P. Purtell, T. McMullen and J. Fein, the Office of Naval Research under grants N00014-89-J-3186 and N00014-93-1-0774, the Advanced Research Project Agency under contracts N00014-92-1726 and N00014-94-1-0735 and the Sea Grant Program under Grant Number NA46RG0434 is gratefully acknowledged.

Aleyev
,
Y.
(
1977
).
Nekton. The Hague, Netherlands: Dr W. Junk b.v. Publishers
.
Anderson
,
J.
(
1996
).
Vorticity control for efficient propulsion. Doctoral thesis, Massachusetts Institute of Technology and the Woods Hole Oceanographic Institution
.
Anderson
,
J.
,
Streitlien
,
K.
,
Barrett
,
D.
and
Triantafyllou
,
M.
(
1998
).
Oscillating foils of high propulsive efficiency
.
J. Fluid Mech.
360
,
41
72
.
Barrett
,
D.
(
1996
).
Propulsive efficiency of a flexible hull underwater vehicle. Doctoral thesis, Massachusetts Institute of Technology
.
Barrett
,
D.
,
Triantafyllou
,
M.
,
Yue
,
D.
,
Grosenbaugh
,
M.
and
Wolfgang
,
M.
(
1999
).
Drag reduction in fish-like locomotion
.
J. Fluid Mech
.
392
,
183
212
.
Blake
,
R. W.
(
1983
).
Fish Locomotion
.
Cambridge
:
Cambridge University Press
.
Cheng
,
J.
,
Zhuang
,
L.
and
Tong
,
B.
(
1991
).
Analysis of swimming of three-dimensional waving plates
.
J. Fluid Mech.
232
,
341
355
.
Chopra
,
M.
(
1976
).
Large amplitude lunate-tail theory of fish locomotion
.
J. Fluid Mech.
74
,
161
182
.
Chopra
,
M.
and
Kambe
,
T.
(
1977
).
Hydromechanics of lunate-tail swimming propulsion. Part 2
.
J. Fluid Mech.
79
,
49
69
.
Domenici
,
P.
and
Blake
,
R. W.
(
1997
).
The kinematics and performance of fish fast-start swimming
.
J. Exp. Biol
.
200
,
1165
1178
.
Ellington
,
C. P.
(
1984
).
The aerodynamics of hovering insect flight. IV. Aerodynamic mechanisms
.
Phil. Trans. R. Soc. Lond. B
305
,
79
113
.
Ellington
,
C. P.
,
van den Berg
,
C.
and
Willmott
,
A. P.
(
1996
).
Leading-edge vortices in insect flight
.
Nature
384
,
626
630
.
ffowcs-Williams
,
J.
and
Zhao
,
B.
(
1989
).
The active control of vortex shedding
.
J. Fluids Struct.
3
,
115
122
.
Gopalkrishnan
,
R.
,
Triantafyllou
,
M.
,
Triantafyllou
,
G.
and
Barrett
,
D.
(
1994
).
Active vorticity control in a shear flow using a flapping foil
.
J. Fluid Mech.
274
,
1
21
.
Gray
,
J.
(
1936
).
Studies in animal locomotion. VI. The propulsive powers of the dolphin
.
J. Exp. Biol
.
13
,
192
199
.
Guiggiani
,
M.
,
Krishnasamy
,
G.
,
Rudolph
,
T.
and
Rizzo
,
F.
(
1992
).
A general algorithm for the numerical solution of hypersingular boundary integral equations
.
J. Appl. Mech
.
59
,
604
614
.
Harper
,
D.
and
Blake
,
R. W.
(
1990
).
Fast-start performance of rainbow trout Salmo gairdneri and northern pike Esox lucius during escapes
.
J. Exp. Biol
.
150
,
321
342
.
Hoerner
,
S.
(
1965
).
Fluid-Dynamic Drag
.
Vancouver, WA
:
Hoerner Fluid Dynamics
.
Hoerner
,
S.
(
1985
).
Fluid-Dynamic Lift
.
Vancouver, WA
:
Hoerner Fluid Dynamics
.
Katz
,
J.
and
Plotkin
,
A.
(
1991
).
Low-speed Aerodynamics: From Wing Theory to Panel Methods. Series in Aeronautical and Aerospace Engineering
.
New York
:
McGraw-Hill, Inc
.
Kinnas
,
S.
and
Hsin
,
C.
(
1992
).
Boundary element method for the analysis of the unsteady flow around extreme propeller geometries
.
AIAA J
.
30
,
688
696
.
Krasny
,
R.
(
1986
).
Desingularization of periodic vortex sheet roll-up
.
J. Comput. Phys
.
65
,
292
313
.
Lan
,
C.
(
1979
).
The unsteady quasi-vortex-lattice method with applications to animal propulsion
.
J. Fluid Mech.
93
,
747
765
.
Lighthill
,
M.
(
1960
).
Note on the swimming of slender fish
.
J. Fluid Mech.
9
,
305
317
.
Lighthill
,
M.
(
1975
).
Mathematical Biofluiddynamics. Philadelphia, PA: Society for Industrial and Applied Mathematics
.
Liu
,
H.
,
Wassenberg
,
R.
and
Kawachi
,
K.
(
1997
).
The three-dimensional hydrodynamics of tadpole swimming
.
J. Exp. Biol
.
200
,
2807
2819
.
Martin
,
P.
(
1996
).
Mapping flat cracks onto penny-shaped cracks, with application to somewhat circular tensile cracks
.
Q. Appl. Math
.
54
,
663
675
.
Martin
,
P.
and
Rizzo
,
F.
(
1989
).
On boundary integral equations for crack problems
.
Proc. R. Soc. Lond. A
421
,
341
355
.
McCutchen
,
C.
(
1976
).
Flow visualization with stereo shadowgraphs of stratified flow
.
J. Exp. Biol
.
65
,
11
20
.
Müller
,
U.
,
van den Heuvel
,
B.
,
Stamhuis
,
E.
and
Videler
,
J.
(
1997
).
Fish foot prints: morphology and energetics of the wake behind a continuously swimming mullet (Chelon labrosus Risso)
.
J. Exp. Biol
.
200
,
2893
2906
.
Osborne
,
M.
(
1951
).
Aerodynamics of flapping flight with application to insects
.
J. Exp. Biol
.
28
,
221
245
.
Rosen
,
M.
(
1959
).
Water flow about a swimming fish. NAVWEPS Report 2298. US Naval Ordinance Test Station, China Lake, CA
.
Stamhuis
,
E.
and
Videler
,
J.
(
1995
).
Quantitative flow analysis around aquatic animals using laser sheet particle image velocimetry
.
J. Exp. Biol
.
198
,
283
294
.
Streitlien
,
K.
,
Triantafyllou
,
G.
and
Triantafyllou
,
M.
(
1996
).
Efficient foil propulsion through vortex control
.
AIAA J.
34
,
2315
2319
.
Taneda
,
S.
and
Tomonari
,
Y.
(
1974
).
An experiment on the flow around a waving plate
.
J. Phys. Soc. Japan
36
,
1683
1689
.
Tokomaru
,
P.
and
Dimotakis
,
P.
(
1991
).
Rotary oscillation control of a cylinder wake
.
J. Fluid Mech.
224
,
77
90
.
Triantafyllou
,
G.
,
Triantafyllou
,
M.
and
Grosenbaugh
,
M.
(
1993
).
Optimal thrust development in oscillating foils with application to fish propulsion
.
J. Fluids Struct.
7
,
205
224
.
Triantafyllou
,
M.
,
Triantafyllou
,
G.
and
Gopalkrishnan
,
R.
(
1991
).
Wake mechanics for thrust generation in oscillating foils
.
Phys. Fluids A
3
,
2835
2837
.
Videler
,
J.
(
1981
).
Swimming movements, body structure and propulsion in cod Gadus morhua
.
Symp. Zool. Soc. Lond
.
48
,
1
27
.
Videler
,
J.
(
1993
).
Fish Swimming. London: Chapman & Hall
.
Videler
,
J.
and
Hess
,
F.
(
1984
).
Fast continuous swimming of two pelagic predators, saithe (Pollachius virens) and mackerel (Scomber scombrus): a kinematic analysis
.
J. Exp. Biol
.
109
,
209
228
.
von Kármán
,
T.
and
Burgess
,
J.
(
1935
).
General aerodynamic theory – Perfect fluids
. In
Aerodynamic Theory
, vol.
2
(ed.
W.
Durand
), pp.
346
349
.
Berlin
:
Springer-Verlag
.
Weihs
,
D.
(
1972
).
A hydrodynamical analysis of fish turning manoeuvres
.
Proc. R. Soc. Lond. B
182
,
59
72
.
Willert
,
C.
and
Gharib
,
M.
(
1991
).
Digital particle image velocimetry
.
Exp. Fluids
10
,
181
193
.
Wu
,
T.
(
1961
).
Swimming of waving plate
.
J. Fluid Mech.
10
,
321
344
.
Wu
,
T.
(
1971a
).
Hydromechanics of swimming propulsion. Part 1. Swimming of a two dimensional flexible plate at variable forward speeds in an inviscid fluid
.
J. Fluid Mech.
46
,
337
355
.
Wu
,
T.
(
1971b
).
Hydromechanics of swimming propulsion. Part 2. Some optimum shape problems
.
J. Fluid Mech.
46
,
521
544
.
Wu
,
T.
(
1971c
).
Hydromechanics of swimming propulsion. Part 3. Swimming and optimum movements of slender fish with side fins
.
J. Fluid Mech.
46
,
545
568
.