Next Article in Journal
Distribution of Bisphenol A in Sediment and Suspended Matter and Its Possible Impact on Marine Life in Kaštela Bay, Adriatic Sea, Croatia
Next Article in Special Issue
Performance Assessment of a Planing Hull Using the Smoothed Particle Hydrodynamics Method
Previous Article in Journal
Comprehensive Natural Environment and Landscape Signs in Coastal Settlement Hazard Assessment: Case of East Taiwan between the Coastal Mountain and the Pacific Ocean
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydrodynamics of Biomimetic Marine Propulsion and Trends in Computational Simulations

1
Escola Politécnica Superior, Universidade da Coruña, 15403 Ferrol, Spain
2
Technical Department, Norplan Engineering S.L., 15570 Naron, Spain
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2020, 8(7), 479; https://doi.org/10.3390/jmse8070479
Submission received: 1 June 2020 / Revised: 26 June 2020 / Accepted: 27 June 2020 / Published: 29 June 2020
(This article belongs to the Special Issue Unconventional Marine Vehicles)

Abstract

:
The aim of the present paper is to provide the state of the works in the field of hydrodynamics and computational simulations to analyze biomimetic marine propulsors. Over the last years, many researchers postulated that some fish movements are more efficient and maneuverable than traditional rotary propellers, and the most relevant marine propulsors which mimic fishes are shown in the present work. Taking into account the complexity and cost of some experimental setups, numerical models offer an efficient, cheap, and fast alternative tool to analyze biomimetic marine propulsors. Besides, numerical models provide information that cannot be obtained using experimental techniques. Since the literature about trends in computational simulations is still scarce, this paper also recalls the hydrodynamics of the swimming modes occurring in fish and summarizes the more relevant lines of investigation of computational models.

1. Introduction

Biomimetics, also called biomimicry or bionics, involves the imitation of nature in man-made systems. An area of biomimetics that is gaining an increasing interest is marine propulsion. Fish swimming is the result of millions of years of evolutionary optimization to develop a very efficient movement. For this reason, several fish movements inspire technology for marine propulsion. Marine propulsion is a multi-disciplinary area and involves numerous uses. Biomimetic marine propulsors are currently employed in oceanographic observation, military, underwater mine detection, surveillance, pipeline inspection, gas drilling, environmental awareness, rescue, etc. Fish have evolved to achieve different capabilities. Some fish are maneuverable, some are efficient, some are fast swimmers, etc. For engineering purposes, it is necessary to design the most appropriate swimming form depending on the application. For instance, oceanographic observation and pipeline inspection require high maneuvering. Other applications, such as excavation of sunken wreckage or gas drilling, require robustness and maneuvering in narrow spaces. Biological monitoring throughout large regions requires fast vehicles with the capability to travel long distances. In the military area, where spying and avoidance of detect are the two most important fields, noiseless and inconspicuous wakes are very important. In merchant ships, the most important parameter is efficiency in order to overcome the traditional propellers.
In 1936, a decisive paper was published by Gray [1], who initiated the interest in the hydrodynamics of fish swimming. Gray studied the movement of dolphins and postulated that these mammals can obtain high speeds and accelerations with a small muscle mass. He estimated the power required by a rigid model of a dolphin as about a sevenfold shortage in muscle mass to attain such speeds. He concluded that dolphin skin has special drag reduction properties. His research, known as Gray’s paradox, inspired several works about the drag reduction capabilities of dolphins [2,3,4,5,6,7,8,9]. Gray’s paradox remained controversial for the following sixty years, mainly because it is extremely difficult to obtain reliable force measurements from live fish. In the 1990s, several investigations from Massachusetts Institute of Technology (MIT) reported that Gray´s calculations were wrong [10,11,12,13]. Gray had employed a rigid model of a dolphin, and these works showed that drag is lower for an undulating body than that corresponding to the rigid equivalent. Nevertheless, these authors agreed that fish employ techniques to reduce drag forces while swimming. Basically, these techniques are based on the utilization of the energy that exists in the eddies of an oncoming flow by repositioning the vorticities [14].
Great progress was made since the 1990s. Due to the increasing number of investigations about biomimetic marine propulsion, several reviews can be found in the literature. The first and most relevant one was developed by Sfakiotakis et al. [15] in 1999. This paper presented an overview of fish swimming modes, providing a classification and discussion about the relevance to underwater vehicle design. They analyzed the advantages, disadvantages, kinematics, and recommendable applications of several swimming modes. Although many mechanisms were developed after the publication of this paper, the work of Sfakiotakis et al. is still highly referenced due to its excellent explanation and classification of fish swimming modes for aquatic locomotion. Another review is the one of Cheng and Chahine [16], who presented the state of the art of swimming hydrodynamics and their biological relevance. Triantafyllou et al. [17] focused their review on the hydrodynamics of flapping foils and analyzed the implementation of such propulsion systems in marine propulsors. Colgate et al. [18] reviewed the control of robot fish, analyzing the main problems and solutions. Mittal et al. [19] reviewed the works in bio-hydrodynamics, from flapping foils to fin flexion mechanisms. Bandyopadhyay [20] reviewed aspects related to hydrodynamics, artificial muscle technology, and neuroscience-based control. Kato [21] reviewed the studies about kinematics, hydrodynamics, and performance of median and paired fin (MPF) locomotion in fish and biomimetic mechanical systems from the viewpoint of enhancing the propulsive and maneuvering performance of marine vehicles at low speeds. Chu et al. [22] centered on smart actuators used on biomimetic robots, particularly shape memory alloys, ionic polymer metal composites, and lead zirconate titanates. Raj and Thakur [23] discussed several design features like sensing, actuation, autonomy, waterproofing, and morphological structure. Other researchers [24,25,26,27,28] focused on flapping foils.
These reviews constitute a comprehensive state of the art of biomimetic marine propulsion. Nevertheless, the literature about trends in computational techniques is still scarce. According to this, this paper aims to supplement the literature providing the state of the works in the field of hydrodynamics and computational simulations to analyze biomimetic marine propulsors. After a description of fish swimming modes and biomimetic marine propulsors which mimic fishes, aspects related to computational simulations, such as hydrodynamics, non-dimensional parameters, turbulence, grid strategies, and experimental validations, are analyzed.

2. Fish Swimming Modes and Biomimetic Marine Propulsors

A comprehensive classification and description of the fish swimming modes was developed by Breder [29] in 1926 and expanded by Lindsey [30] in 1978. Basically, they differenced two propulsion modes: swimming and non-swimming locomotion. These are explained below.

2.1. Swimming Locomotion

The most common fish swimming modes are summarized in Figure 1. Basically, there are two factors to classify them: the extent to which the propulsion is undulatory or oscillatory and the body structures or fins that generate thrust. Most fish generate thrust using their body and/or caudal fin (BCF), for instance anguilliforms, caranguiforms, thunniforms, and ostraciiforms. Other fish, however, employ their median and/or paired fins (MPF) while the body remains practically rigid. Examples of MPF movement are amiiforms, gymnotiforms, rajiforms, labriforms, and tetraodontiforms. Several reviews [15,16,17,18,19,20,21,22] indicated that BCF modes, especially thunniform and, to a lesser extent, caranguiform, are more efficient than MPF modes. Nevertheless, MPF modes are characterized by a high maneuverability in terms of fast start, short-radius turning, and other factors.
The force actuating on a fish is the result of a viscous (friction) force due to the viscous shear stresses acting on the body and a pressure force due to pressure gradients. Hydrodynamically, fish locomotion can be generated by three procedures: drag-based, added mass, and momentum injection (Figure 2). Most swimmers use a combination of more than one of these procedures.
In the drag-based procedure (Figure 2a) a force due to the friction between the fin and water acts opposite to the direction of motion. In the added mass procedure (Figure 2b), a reaction force is produced by the inertia of the water pushed backward by the body [31]. This propulsive force is approximately the product of the water mass accelerated and its acceleration. The momentum injection procedure (Figure 2c) is based on impulsively injecting momentum into the surrounding fluid. A force is produced by jetting fluid from the body. The momentum, which is approximately the product of the mass flow rate and the jet velocity, creates a reaction force.
Besides these locomotion procedures, most fish evolved vortex control capabilities [10,11,12,13]. This is typical of thunniform swimming mode and is presented in most fishes. As mentioned previously, dolphins have drag reduction capabilities because they employ the energy that exists in the eddies of an oncoming flow by repositioning the vortices. Despite some superficial similarities between thunniform and ostraciiform swimmers, the hydrodynamics of the latter are less efficient and based on a rigid body which achieves less vorticity control. Oscillatory movements also obtain a poorer vorticity control.
A detailed analysis of each movement and repercussion on marine propulsion is shown below.

2.1.1. BCF Swimming Modes

As indicated in Figure 1, the most common BCF swimming modes are anguilliform, caranguiform, thunniform, and ostraciiform. In anguilliform swimming mode, an undulation with constant amplitude takes place along the whole body, from head to tail. It is typical of eels, congers, muraenas, lampreys, lancelets, etc. One advantage of this movement is the turning and accelerating abilities due to the flexibility of the body. Another advantage is the possibility to swim backward, simply passing the undulation from tail to head. Due to its lower efficiency compared with other BCF modes, anguilliform swimming movement has only inspired a few robots that require high maneuverability, mimicking lampreys [32,33,34,35,36,37,38,39,40,41], snakes [42], salamanders [43,44,45], eels [46,47,48,49,50,51], etc.
The caranguiform mode is similar to the anguilliform one, but with the amplitude undulations growing toward the tail. Since the caranguiform swimming movement is more efficient than the anguilliform one, it inspired many more mechanisms mimicking pikes [52], salmons [53,54,55], sea breams [56], mackerels [57,58], etc.
As in the case of caranguiform, in thunniform swimming mode, the amplitude undulations grow toward the tail. Nevertheless, in the latter, the undulations practically take place only in the posterior half of the fish. It is typical of tunas, sharks, whales, dolphins, etc. Many researchers agree that thunniform is the most efficient swimming mode [15,16,17,18,19,20,21,22]. Among thunniform swimmers, tuna is considered the fastest fish in nature. For these reasons, most studies deal with tuna-like mechanisms [59,60,61,62,63,64,65,66,67]. The pioneer tuna-like mechanism was RoboTuna [11,59], made at MIT. This mechanism is a replica of a real tuna. Results from the RoboTuna project did indeed suggest a reduction in the drag force, agreeing with Gray’s paradox. Following the success of the RoboTuna project, Anderson and Chapra at MIT improved this robot providing more vorticity control propulsion and maneuvering [60]. Another important animal that inspired thunniform mechanisms is dolphin [68,69,70,71,72,73,74,75,76,77,78,79,80]. The popularity of Gray’s paradox made dolphins an important source of bioinspiration, and the kinematics of dolphins are among the best understood in nature. At Istanbul Technical University, Dogangil et al. also constructed a robotic dolphin [78] and analyzed the hydrodynamics and kinematics [79]. To a lesser extent, shark also inspired biomimetic propulsion mechanisms [81,82].
Ostraciiform locomotion is characterized by the pendulum-like oscillation of a rather stiff caudal fin, while the body remains essentially rigid. It is typical of cowfishes, boxfishes, coffer fishes, trunkfishes, etc. Due to its low efficiency [15,18,21], this movement has received limited attention in engineering, although an oscillating movement is easier to implement than an undulating one. Most of the investigations found in the literature are inspired by boxfish [83,84,85,86]. Other authors employed flapping foils for oscillating marine propulsors [87,88,89,90,91,92,93,94,95,96,97,98,99,100,101].
Table 1 summarizes the BCF swimming mechanisms mentioned, including the authors, institution, propulsion mode, and source of inspiration.

2.1.2. MPF Swimming Modes

As indicated in Figure 1, the most common MPF swimming modes are rajiform, amiiform, gymnotiform, labriform, and tetraodontiform. Rajiform (rays, skates, mantas, etc.), amiiform (bowfins, etc.), and gymnotiform (electrical eels, American knifefishes, etc.) modes are based on the undulation of the pectoral, dorsal, and anal fins, respectively. Nevertheless, labriform (wrasses, angelfish, surfperches, etc.) is based on the oscillation of the pectoral fin and tetraodontiform (blowfishes, etc.) on the oscillation of both anal and dorsal fins. Speed, performance, and maneuverability are key aspects of pectoral fin locomotion which provide specific application ideas for underwater vehicle technology. MPF locomotion mode is less efficient than BCF, but more maneuverable, and it is easy to reverse thrust [15,16,18]. By controlling the movement of the fins, it is possible to obtain maneuvering forces in the full six degrees of freedom and both rowing and flapping.
Due to its low efficiency, much fewer MPF mechanisms were developed by comparison with BCF mechanisms. Both oscillatory and undulatory MPF mechanisms were developed in the literature, with a preference towards pectoral fin labriform-type swimming. Concerning oscillatory-type MPF movement, labriform swimming has received special attention inspiring mechanisms that mimics black basses [102], sea turtles [103,104,105,106,107], sunfishes [108,109], flapping foils [110,111,112], rays [113,114,115,116,117,118,119,120], knifefishes [121,122,123,124,125,126,127,128], etc. Concerning undulatory-type MPF locomotion, little interest has been showed by comparison with oscillatory MPF movement. Table 2 summarizes the MPF swimming mechanisms mentioned, including the authors, institution, propulsion mode, and source of inspiration.
In order to analyze propulsive effectiveness, most researchers use the so-called Froude efficiency. William Froude (1810–1879) was the first who used the propulsive efficiency, defined as the ratio of power output to the rate of energy input. Lamas et al. [49] developed an undulating propulsor actuated by several ribs. After that, Lamas et al. [50,51] employed CFD (Computational Fluid Dynamics) to study several amplitudes, wavelengths, and oscillation frequencies in anguilliform, carangiform, thunniform, ostraciiform, and gymnotiform movements. They concluded that thunniform is the most efficient swimming mode in terms of the Froude efficiency, followed by caranguiform, ostraciiform, anguilliform, and finally, gymnotiform. Caranguiform, ostraciiform, and anguilliform swimming modes are less efficient than thunniform because the latter produce thrust only at the tail. Nevertheless, the large undulation amplitudes along the entire body of caranguiform, anguilliform, ostraciiform, and gymnotiform swimming modes produce wasted power at the head.
It is worth mentioning that a single most appropriate swimming form does not exist. Efficiency is only one among other important assets of bioinspired aquatic propulsion such as low-speed maneuverability, fast start, and short-radius turning [129,130]. These are the most important advantages of employing a bioinspired design. Efficiency is a controversial aspect of aquatic bioinspired propulsion since bioinspired vehicles are not merely built for their energetic efficiency, but for their suitability at dealing with tasks for which common vehicles are unfit. Fish have grown to become specialists; they excel at a range of movements their habitat and respective ecosystems dictate, but at the expense of not being well adapted to other ranges of motion. For instance, while thunniform swimming does provide superior efficiency for high cruise speed in calm water, it performs poorly at lower speeds, and provides only low levels of agility in turning maneuvers [15].

2.2. Non-Swimming Locomotion

Some aquatic animals use other mechanisms than swimming. The most important is jet propulsion (cephalopods, shellfishes, jellyfishes, etc.), based on the ejection of water behind the fish. The phenomenon is shown in Figure 3, which illustrates a vortex ring generated during contraction and another one generated during expansion. Pulsed jetting seems to have gained a growing attention in the scientific community in recent years.
California Institute of Technology, through the works [131,132,133,134,135,136], was the first institution which studied jet propulsion. Particularly, the hydrodynamics of jet propulsion were explained in the review of Dabiri [121]. Briefly, it is based on the momentum injection procedure, i.e., water behind the fish produces a jet which generates thrust. Besides, these researchers found an important vortex optimization capability by analyzing a piston-cylinder mechanism. In their configuration, a boundary layer of vorticity forms on the inner cylinder wall as the piston moves downstream inside the cylinder, and through oscillatory jets they found that the boundary layer is less susceptible to separation; this fact mitigates turbulence. Their measurements of the impulse associated with vortex ring formation showed to be much larger than that expected from the jet velocity alone.
Table 3 summarizes the main non-swimming mechanisms, including the authors, institution, propulsion mode, and source of inspiration.

3. Numerical Models

3.1. Hydrodynamics

A proper design of a mechanism requires a comprehensive characterization of the fluid pattern. In this regard, numerical models can be used as an important tool to analyze the fluid flow in detail. Between them, CFD is highly employed. This is a branch of fluid mechanics based on the split of the domain into small elements called grid or mesh. For each grid element, the governing equations are solved using numerical techniques. In a biologically inspired mechanism, the complex hydrodynamics may be computed by solution of the Navier–Stokes equations using numerical analysis. CFD provides important information that cannot be measured experimentally, such as details of the pressure field, velocity field, hydrodynamic forces, power, and efficiency. It is very useful to complement experimental works using less time and lower cost.
During the movement of the mechanism, water is pushed backward and produces a wake of alternating sign vortices known as a reverse Karman street [50,64,78,110]. This phenomenon is shown in Figure 4a,b, which represents the velocity field overlaid with the pressure field for BCF and MPF swimming modes, respectively [50]. As can be seen in these figures, several vortices are formed behind the fin. Van Buren et al. [147] manipulated the vortex structure in the wake on a pitching panel and verified important effects on thrust and efficiency. As indicated previously, it is possible to control the thrust by manipulating the flow vortex, i.e., by a mechanism of vorticity control [10,12,64]. A fish consumes much less energy to displace than a rigid body because the motion of the fish is associated with a reduction of the drag force and, thus, an increment of the propulsive efficiency by reducing separation and suppressing turbulence [148]. Recent works have shown that undulating fish movement is very efficient compared to rigid bodies [14,59,149].
The pressure differences promote a pressure force, given by:
F p = S p n ^ d S
where p is the pressure, S the surface, and n ^ a unit vector normal to the surface. On the other hand, a viscous (friction) force is produced due to the viscosity of water, and is given by the following expression:
F v = S τ i j n ^ d S
where τij is the viscous stress tensor. The total force is the sum of the pressure and viscous forces. A typical distribution of the components of these forces in the direction of the movement as well as the average total force are shown in Figure 5 for anguilliform swimming mode [50]. As can be seen in this figure, the pressure and viscosity forces exhibit two peaks per cycle, corresponding to the forward and backward tail strokes. Under the conditions indicated in the figure, the propulsor accelerates since the total force is positive. If the body is accelerated, the inertia of the system contributes as an additional force, and the additional inertia of the system also promotes an added mass effect [131,150,151].
The average forces per cycle against the cruising velocity for anguilliform swimming are shown in Figure 6 [50]. As can be seen, when the velocity is zero, the pressure force is the maximum, and the viscous force is the minimum. As the velocity is increased, the pressure force is reduced, and the viscous force is increased until the velocity is such that the pressure force equals the viscous force, and thus, the total force is zero. Under this condition, the propulsor displaces at constant velocity, i.e., there is neither acceleration nor deceleration. In Figure 6, this velocity corresponds to 5.5 m/s. If the velocity exceeds this value, the propulsor decelerates.

3.2. Non-Dimensional Parameters

In biomimetic systems, it is useful to employ non-dimensional numbers to generalize the involving parameters [49]. The governing equations corresponding to a flow around a surface are the Navier–Stokes equations of conservation of mass and momentum, given by:
u = 0
u t + ( u u ) = p ρ + ν 2 u + g
where u represents the velocity, p the pressure, ρ the density, ν the kinematic viscosity, and g the gravitational acceleration. The parameters in the equations above can be converted to non-dimensional quantities by using reference values Lref, uref, pref, tref, and gref for length, velocity, pressure, time, and gravitational acceleration, respectively. The non-dimensional length, velocity, pressure, time, and gravitational acceleration are given by L* = L/Lref, u* = u/uref, p* = p/pref, t* = t/tref, and g* = g/gref, respectively, and the resulting equations yield:
1 L r e f * ( u * u r e f ) = 0
1 t r e f ( u * u r e f ) t * + 1 L r e f ( u * u r e f u * u r e f ) = 1 ρ 1 L r e f * ( p * p p r e f ) + ν 1 L r e f 2 * 2 ( u * u r e f ) + g * g r e f
Employing the reference parameters given in Table 4, where L is the fin length and U the cruising velocity, the resulting equations of conservation of mass and momentum are given by Equations (7) and (8).
1 L * ( u * U ) = 0
U L ( u * U ) t * + 1 L ( u * U u * U ) = 1 ρ 1 L * ( p * ρ U 2 ) + ν 1 L 2 * 2 ( u * U ) + g * g
which yields:
* u * = 0
u * t * + ( u * u * ) = * p * + ν U L * 2 u * + L g U 2 g *
Expressing the Reynolds number as Re = UL/ν and the Froude number as F r = U / L g , the resulting governing equations in non-dimensional form are:
* u * = 0
u * t * + * ( u * u * ) = * p * + 1 Re * 2 u * + 1 F r 2 g *
As can be seen, the Reynolds number, which represents the relation between inertial and viscous effects, and the Froude number, which represents the relation between inertial and gravity effects, are decisive parameters in the hydrodynamics of marine propulsors.
The frequency is usually expressed as the non-dimensional Strouhal number, St, given by St = fA/U, where f is the frequency and A the amplitude. It was highly reported that the Strouhal number is related to the vortex formation. In fact, typical values of the Strouhal number in nature are in a narrow range, approximately 0.2–0.4. Several researchers verified that the Strouhal number is related to the optimal vortex formation, which is produced in this narrow range observed in nature [59,87,149,152,153,154,155].

3.3. Turbulence

A few studies [93,156,157,158] employed inviscid models to analyze the hydrodynamics of biomimetic propulsion. The most important advantage of inviscid simulations is the cheap computational cost, but their accuracy is questionable. An important dimensionless parameter that involves viscosity is the Reynolds number. This parameter constitutes an indication of turbulence. The simulations that can be found in the literature span a wide range of Reynolds numbers due to the variety of values found in nature. Figure 7 shows the Reynolds number for different fish (from zebrafish larvae to stingrays and sharks), amphibians (tadpoles), reptiles (alligators), marine birds (penguins), and large mammals (from manatees and dolphins to belugas and blue whales) according to Gazzola et al. [151].
Another important parameter is the swimming number, Sw = fAL/ν. Although most numerical studies employ the Strouhal number, the swimming number provides the advantage of accounting for two length scales: amplitude and length. The swimming number is a transverse Reynolds number which characterizes the undulating movement. Gazzola et al. [152] analyzed more than 1000 measurements of fish varying in size from millimeters to meters and found that Re ̴ Swα, with α = 4/3 for laminar flows and α = 1 for turbulent flows. They verified that around Re 103–104, the transition from the laminar to the turbulent regime takes place, as indicated in Figure 8.
Some organisms swim at low Reynolds numbers and thus under laminar flow [149,159]. Nevertheless, most engineering biomimetic mechanisms operate at high Reynolds numbers and thus under turbulent flow. According to this, the turbulent regime was the most employed one in the numerical simulations of biomimetic propulsors.
When a body moves along a fluid, a boundary layer exists due to the “no-slip” condition and viscosity. The fluid in contact with the surface has zero velocity and, close to the surface, the fluid is slowed by frictional forces associated to the viscosity of the fluid. The result is a thin layer where the tangential velocity of the fluid increases from zero at the surface to the cruising velocity far away from the surface. The so-called boundary layer is the normal distance between the surface, where the velocity is zero, and the position where the tangential velocity is 0.95–0.99U, depending on the author. Flow over a wavy surface experiences adverse (flow against an increasing pressure) and favorable pressure gradients induced by the wave motion. When a boundary layer undergoes an adverse pressure gradient, the flow near the wall decelerates. If this boundary layer has travelled far enough in the adverse pressure gradient that the velocity becomes negative, i.e., reverses the direction, separation occurs. Figure 9 shows the velocity field in the boundary layer, the last profile corresponds to reverse flow which leads to flow separation.
Flow separation is the result of momentum losses, or decelerations, that eventually prevent the continued streamwise progress of the boundary layer fluid along the body surface. Drag is increased, and thrust is reduced. For this reason, much research has gone in the study of surfaces which delay flow separation and keep the flow attached for as long as possible. Some wave motions reduce the drag force and increase the efficiency by restraining separation and suppressing turbulence [148,160,161]. The travelling wave laminarizes the flow, and the fluid motion in the wave direction is accelerated. The turbulence intensity is increased when separation takes place since separation is the major mechanism for turbulence production in undulating surfaces. Several authors [57,162] found that approximately when the wave phase velocity is larger than the external flow velocity, the wavy surface pushes the fluid so strongly that separation occurs.
RANS (Reynolds-averaged Navier–Stokes), LES (Large-Eddy Simulation), and DNS (Direct Numerical Simulation) approaches were employed in the literature in order to simulate turbulence. RANS procedure computes the mean flow quantities. It includes only the largest flow structures, while the smaller scales are not included. It is robust, computationally cheap, and reasonably accurate for a wide range of engineering problems. Nevertheless, the main disadvantage is that the information about the flow is limited, including the simulation of flow separation [19,163]. In biomimetic marine propulsion, the effect of separation may be important, and thus, the RANS approach does not provide an accurate prediction of the flow in the separated region. Kim et al. [164] compared six two-equation models for detachment flow and determined that the most accurate in their simulations was k-ω and k-ωSST models. Another disadvantage of RANS models is the simulation of transitional laminar-turbulent flows.
Regarding DNS, this computes all the turbulent scales in the grid and temporal resolution. In DNS, the Reynolds stresses are thus not modeled. The velocity field is obtained by integrating the three-dimensional time-dependent Navier–Stokes equations. The main advantage is that this model provides accurate results, but the disadvantage is the computational cost. In the recent years, computational resources have improved noticeably and made possible several works about direct numerical simulations [165,166,167,168,169]. The spatial resolution required increases as the Reynolds number increases. For this reason, DNS is mostly used to study turbulent flows at low or moderate Reynolds numbers [148,168,169].
Finally, the LES approach lies between RANS and DNS. LES resolves directly large eddies, while small eddies are modeled. Resolving only the large eddies allows to use much coarser meshes and larger time-step sizes than those employed in DNS. However, LES requires substantially finer meshes than RANS. The mesh is related to the turbulence scale. The small turbulence scale becomes smaller at high Reynolds flows, and thus, it needs finer meshes, especially in the near-wall region. The high computational cost has thus become a limitation of LES for high Reynolds simulations and the main advantage is that LES provides flow separation using less computational cost than DNS [170,171,172].

3.4. Grid Strategies

Depending on the mathematical representation of the moving boundary, two approaches to simulate bioinspired marine propulsors can be found in the literature. The first one, known as the interface-tracking or Lagrangian method, consists in employing a moving mesh which follows the movement of the boundary. Therefore, the edges of the mesh cells are aligned to the moving boundary, as indicated in Figure 10a. The second approach, known as interface-capturing or Eulerian method, consists in employing a fixed mesh and implementing the movement of the boundary in the governing equations (Figure 10b). In this approach, the governing equations for both solid and fluid are solved simultaneously and the fluid–solid interaction in computed by source terms in the governing equations.
The habitual approach in the literature was to employ a grid that moves with the surface. Moving meshes provide accuracy but the main disadvantages are the difficulty to handle complex moving boundaries and the high computational cost. Moving meshes must be reconstructed at each time step, and the remeshing techniques increment the computational cost noticeably; in some cases, the convergence may result difficult. Fixed meshes eliminate the computational cost of remeshing procedures during the simulations. Nevertheless, the main disadvantage of fixed meshes is the accuracy. In biomimetic marine propulsion fixed meshes have been employed since the 1980s [173] and have been continued in the recent years [170,171,172,173,174,175,176,177,178,179,180,181,182,183,184]. These are appropriate to represent sharp movements such as free-surface boundary on the incident waves [185,186,187,188].

4. Experimental Validations

The validation process constitutes a crucial part of computational simulations. In order to validate the models and visualize the wake structures behind the propulsor, it is very common to employ PIV (Particle Image Velocimetry) [87,108,109,166,175,189,190]. This is an optical fluid measurement technique that provides instantaneous fluid velocity fields using image processing and statistics. By seeding particles in the fluid and shooting a laser sheet behind the moving surface, the flow pattern is analyzed.
During PIV, the particle concentration is such that it is possible to identify individual particles in an image, but not with certainty to track it between images. When the particle concentration is so low that it is possible to follow an individual particle, it is called PTV (Particle Tracking Velocimetry). This is a velocimetry method, i.e., a technique to measure the velocity of particles that are resident in a fluid. There are several works about PIV analysis of fishes [191,192,193,194,195,196,197] and robotic fishes [198,199]. Anderson et al. [200] compared PIV with PTV and found that both procedures solved the outermost regions of the boundary layer but failed to resolve the flow very close to a moving surface. In the outer regions of the boundary layer, PIV and PTV provided the same velocities; nevertheless, they recommended PTV to analyze the velocity near the surface although the errors in the results also depend on camera pixel resolution, field of view dimensions, particle shape, size, centroid analysis, and image quality. The development of high-speed digital video technology in the recent years and the availability of lower cost continuous wave lasers has facilitated these studies, and PIV and PTV are becoming a common technique among engineers and biologists. Nevertheless, these techniques must be carefully employed in order to analyze the region near the surface. An accurate determination of the velocities in this region is too important to analyze friction.
Other procedures to measure flows are LDV (Laser Doppler Velocimetry) [170], also known as LDA (Laser Doppler Anemometry), and hot-wire anemometry. The main difference between PIV and these techniques is that the former produces two-dimensional or even three-dimensional vector fields, while the other techniques measure the velocity at a point. LDV is the technique of using the doppler shift in a laser beam to measure velocities. Buckles et al. [201] and Kuzan et al. [202] employed LDV to measure velocities working with separating flows and identified the turbulent layer and separated region. Hudson et al. [203,204] used LDV to analyze a wave surface and also analyzed the separated flow and Cherukat et al. [168] developed a computational model to analyze Hudson’s results.

5. Conclusions

This work provides a state about the works in the fields of hydrodynamics and computational simulations to analyze biomimetic marine propulsion. This is an emerging field due to the advantages of biomimetic propulsion over traditional rotary propellers. The advancement of numerical methods in the recent years provides a cheap and efficient tool to analyze the hydrodynamics. The tendency of many current works is to mimic the exact shape of a real fish. One of the reasons to develop a system as a copy is the development of mechanisms capable of capturing energy from a flow. There is a great body of literature describing this phenomenon, and consequently, the hydrodynamics are now better understood. Nevertheless, numerical models provide information that cannot be obtained using experimental techniques such as details of the pressure field, velocity field, hydrodynamic forces, power, and efficiency; analyze in detail all the aspects of the problem; and isolate the effect of each parameter. The characterization of vortex formation is a complex three-dimensional problem. Given the limitations to measure these flows experimentally, numerical models constitute an important tool to characterize the vortex formation.
In the upcoming years, it is expected that the number of numerical analysis grow as computation evolve and improve some deficiencies of current works. For instance, it is very important to improve the computational resources in order to make LES and DNS more affordable. Besides, it is necessary to develop robust moving meshes algorithms which allow abrupt movements with an appropriate convergence.
In order to validate these numerical models and visualize the wake structures behind the propulsor, it is important to develop experimental works such as PIV, PTV, LDV, and other procedures. Besides, it is necessary to improve the current technology in order to obtain more precise data. Future works will focus on developing different numerical models and validate their results using experimental techniques. Once validated, these numerical models will be useful to develop new marine propulsors.

Author Contributions

Conceptualization, M.I.L. and C.G.R.; methodology, M.I.L. and C.G.R.; software, M.I.L. and C.G.R.; validation, M.I.L. and C.G.R.; Formal analysis, M.I.L.; Investigation, M.I.L. and C.G.R.; Resources, M.I.L. and C.G.R.; data curation, M.I.L. and C.G.R.; Writing—original draft preparation, M.I.L.; writing—review and editing, M.I.L.; Visualization, M.I.L.; Supervision, M.I.L. and C.G.R.; Project administration, M.I.L. and C.G.R.; Funding acquisition, M.I.L. and C.G.R. Both authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to express their gratitude to Norplan Engineering S.L. and recommend the courses “CFD with OpenFOAM” and “C++ applied to OpenFOAM” available at www.technicalcourses.net.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gray, J. Studies in animal locomotion: VI. The propulsive powers of dolphin. J. Exp. Biol. 1936, 13, 192–199. [Google Scholar]
  2. Gawn, R.W.L. Aspects of the locomotion of whales. Nature 1948, 161, 44–46. [Google Scholar] [CrossRef]
  3. Parry, D.A. The swimming of whales and a discussion of Gray’s paradox. J. Exp. Biol. 1949, 26, 24–34. [Google Scholar] [PubMed]
  4. Gero, D.R. The hydrodynamic aspects of fish propulsion. Am. Mus. Novit. 1952, 1601, 1–32. [Google Scholar]
  5. Kramer, M.O. Boundary layer stabilization by distributed damping. J. Am. Soc. Nav. Eng. 1960, 72, 25–33. [Google Scholar] [CrossRef]
  6. Webb, P.W. Hydrodynamics and Energetics of Fish Propulsion; Department of the Environment Fisheries and Marine Service: Ottawa, ON, Canada, 1975; pp. 1–158.
  7. Aleyev, Y.G. Hydrodynamic resistance and speed of movement of nekters. Zool. Zhurnal 1974, 53, 493–507. [Google Scholar]
  8. Van Oossanen, P.; Oosterveld, M.W.C. Hydrodynamic resistance characteristics of humans, dolphins, and ship forms. Schiffstechnik 1989, 36, 31–48. [Google Scholar]
  9. Fish, F.E.; Hui, C.A. Dolphin swimming—A review. Mammal Rev. 1991, 21, 181–195. [Google Scholar] [CrossRef]
  10. Gopalkrishnan, R.; Triantafyllou, M.S.; Triantafyllou, G.S.; Barrett, D. Active vorticity control in a shear flow using a flapping foil. J. Fluid Mech. 1994, 274, 1–21. [Google Scholar] [CrossRef] [Green Version]
  11. Barrett, D.S. Propulsive Efficiency of a Flexible Hull Underwater Vehicle. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 1996. [Google Scholar]
  12. Anderson, J.M. Vorticity Control for Efficient Propulsion. Ph.D. Thesis, Massachusetts Institute of Technology, Cambridge, MA, USA, 1996. [Google Scholar]
  13. Streitlien, K.; Triantafyllou, G.S.; Triantafyllou, M.S. Efficient foil propulsion through vortex control. Am. Inst. Aeronaut. Astronaut. J. 1996, 34, 2315–2319. [Google Scholar] [CrossRef]
  14. Triantafyllou, M.S.; Triantafyllou, G.S. An efficient swimming machine. Sci. Am. 1995, 272, 64–70. [Google Scholar] [CrossRef]
  15. Sfakiotakis, M.; Lane, D.M.; Davies, J.B.C. Review of fish swimming modes for aquatic locomotion. IEEE J. Ocean. Eng. 1999, 24, 237–252. [Google Scholar] [CrossRef] [Green Version]
  16. Cheng, J.Y.; Chahine, G.L. Computational hydrodynamics of animal swimming: Boundary element method and three dimensional vortex wake structure. Comp. Biochem. Physiol. Part A 2001, 131, 51–60. [Google Scholar] [CrossRef]
  17. Triantafyllou, M.S.; Techet, A.; Hover, F.S. Review of experimental work in biomimetic foils. IEEE J. Ocean. Eng. 2004, 29, 585–594. [Google Scholar] [CrossRef] [Green Version]
  18. Colgate, J.E. Mechanics and control of swimming: A review. IEEE J. Ocean. Eng. 2004, 29, 660–673. [Google Scholar] [CrossRef]
  19. Mittal, R. Computational modelling in bio-hydrodynamics: Trends, challenges and recent advances. IEEE J. Ocean. Eng. 2004, 29, 595–604. [Google Scholar] [CrossRef]
  20. Bandyopadhyay, P.R. Trends in biorobotic autonomous undersea vehicles. IEEE J. Ocean. Eng. 2005, 30, 109–139. [Google Scholar] [CrossRef]
  21. Kato, N. Median and paired fin controllers for biomimetic marine vehicles. Appl. Mech. Rev. 2005, 58, 238–252. [Google Scholar] [CrossRef]
  22. Chu, W.S.; Lee, K.T.; Song, S.H.; Han, M.W.; Lee, J.Y.; Kim, H.S.; Kim, M.S.; Park, Y.J.; Cho, K.J.; Ahn, S.H. Review of biomimetic underwater robots using smart actuators. Int. J. Precis. Eng. Manuf. 2012, 13, 1281–1292. [Google Scholar] [CrossRef]
  23. Raj, A.; Thakur, A. Fish-inspired robots: Design, sensing, actuation, and autonomy—A review of research. Bioinspir. Biomim. 2016, 11. [Google Scholar] [CrossRef]
  24. Rozhdestvensky, K.; Ryzhov, V. Aerodynamics of flapping-wing propulsors. Prog. Aerosp. Sci. 2003, 39, 585–633. [Google Scholar] [CrossRef]
  25. Naito, S.; Isshiki, H. Effect of bow wings on ship propulsion and motions. Appl. Mech. Rev. 2005, 58, 253–268. [Google Scholar] [CrossRef]
  26. Shyy, W.; Aono, H.; Chimakurthi, S.K.; Trizilia, P.; Kang, C.K.; Cesnik, H.L.C.E.S. Recent progress in flapping wing aerodynamics and aeroelasticity. Prog. Aerosp. Sci. 2010, 46, 284–327. [Google Scholar] [CrossRef]
  27. Xiao, Q.; Zhu, Q. A review on flow energy harvesters based on flapping foils. J. Fluids Struct. 2014, 46, 174–191. [Google Scholar] [CrossRef]
  28. Wu, X.; Zhang, X.; Tian, X.; Li, X.; Lu, W. A review on fluid dynamics of flapping foils. Ocean. Eng. 2019, 195, 106712. [Google Scholar] [CrossRef]
  29. Breder, C.M. The locomotion of fishes. Zoologica 1926, 4, 159–297. [Google Scholar]
  30. Lindsey, C.C. Form, function and locomotory habits in fish. Fish Physiol. 1978, 7, 1–100. [Google Scholar] [CrossRef]
  31. Metin, S.; Menciassi, A.; Low, K.H.; Kim, S. Survey and introduction to the focused section on bio-inspired mechatronics. IEEE/ASME Trans. Mechatron. 2013, 18, 409–418. [Google Scholar] [CrossRef] [Green Version]
  32. Cohen, A.H.; Colmes, P.J.; Rand, R.H. The nature of the coupling between segmental oscillators of the lamprey spinal generator for locomotion: A mathematical model. J. Math. Biol. 1982, 13, 345–369. [Google Scholar] [CrossRef]
  33. Grillner, S.; Deliagina, T.; Ekeberg, A.; Manira, R.H.; Lansner, H.A.; Orlovsky, G.N.; Wallen, P. Neural networks that coordinate locomotion and body orientation in the lamprey. Trends Neurosci. 1995, 18, 270–279. [Google Scholar] [CrossRef]
  34. Grillner, S.; Matsushima, T. The neural network underlying locomotion in lamprey—Synaptic and cellular mechanisms. Neuron 1991, 7, 1–15. [Google Scholar] [CrossRef]
  35. Ayers, J.; Wilbur, C.; Olcott, C. Lamprey robots. In Proceedings of the 1st International Symposium of Aqua Bio-Mechanisms, ISABMEC 2000, Tokai University Pacific Center, Honolulu, HI, USA, 27–30 August 2000. [Google Scholar]
  36. Ayers, J.; Rulkov, N.; Knudsen, D.; Kim, Y.B.; Volkovskii, A.; Selverston, A. Controlling underwater robots with electronic nervous systems. Appl. Bionics Biomech. 2010, 7, 57–67. [Google Scholar] [CrossRef] [Green Version]
  37. Ayers, J.; Westphal, A.; Blustein, D. A conserved neural circuit-based architecture for ambulatory and undulatory biomimetic robots. Mar. Technol. Soc. J. 2011, 45, 147–152. [Google Scholar] [CrossRef]
  38. Wilbur, C.; Vorus, W.; Cao, Y.; Currie, S.N. A lamprey-based undulatory vehicle. In Neurotechnology for Biomimetic Robots; MIT Press: Cambridge, MA, USA, 2002; pp. 285–296. [Google Scholar]
  39. Westphal, A.; Rulkov, N.F.; Ayers, J.; Brady, D.; Hunt, M. Controlling a lamprey-based robot with an electronic nervous system. Smart Struct. Syst. 2011, 8, 39–52. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, W.; Li, F.; Stefanini, C.; Chena, D.; Dario, P. Biomimetic flexible/compliant sensors for a soft-body lamprey-like robot. Robot. Auton. Syst. 2010, 58, 1138–1148. [Google Scholar] [CrossRef]
  41. Xu, J.X.; Niu, X.L.; Ren, Q.Y. Modeling and control design of an anguilliform robotic fish. Int. J. Model. Simul. Sci. Comput. 2012, 3, 1250018. [Google Scholar] [CrossRef]
  42. Crespi, A.; Badertscher, A.; Guignard, A.; Ijspeert, A.J. AmphiBot I: An amphibious snake-like robot. Robot. Auton. Syst. 2005, 50, 163–175. [Google Scholar] [CrossRef] [Green Version]
  43. Ijspeert, J.A. A connectionist central pattern generator for the aquatic and terrestrial gaits of a simulated salamander. Biol. Cybern. 2001, 84, 331–348. [Google Scholar] [CrossRef]
  44. Ijspeert, A.J.; Crespi, A.; Cabelguen, J.M. Simulation and robotics studies of salamander locomotion: Applying neurobiological principles to the control of locomotion in robots. Neuroinformatics 2005, 3, 171–195. [Google Scholar] [CrossRef] [Green Version]
  45. Crespi, A.; Karakasiliotis, K.; Guignard, A.; Ijspeert, A.J. Salamandra Robotica II: An amphibious robot to study salamander-like swimming and walking gaits. IEEE Trans. Robot. 2013, 29, 308–320. [Google Scholar] [CrossRef] [Green Version]
  46. McIsaac, K.P.; Ostrowski, J.P. Experiments in closed-loop control for an underwater eel-like robot. In Proceedings of the IEEE International Conference on Robotics and Automation, Washington, DC, USA, 11–15 May 2002; pp. 750–755. [Google Scholar] [CrossRef]
  47. McIsaac, K.P.; Ostrowski, J.P. Experimental verification of open-loop control for an underwater eel-like robot. Int. J. Robot. Res. 2002, 21, 849–859. [Google Scholar] [CrossRef]
  48. McIsaac, K.P.; Ostrowski, J.P. Motion planning for anguilliform locomotion. IEEE Trans. Robot. Autom. 2003, 19, 637–652. [Google Scholar] [CrossRef]
  49. Lamas, M.I.; Rodríguez, J.D.; Rodríguez, C.G.; González, P.B. Design aspects and two-dimensional CFD simulation of a marine propulsor based on a biologically-inspired undulating movement. J. Marit. Res. 2020, 7, 73–88. [Google Scholar]
  50. Lamas, M.I.; Rodríguez, J.D.; Rodríguez, C.G. CFD analysis of biologically-inspired marine propulsors. Brodogradnja 2012, 63, 125–133. [Google Scholar]
  51. Lamas, M.I.; Rodríguez, J.D.; Rodríguez, C.G.; González, P.B. Three-dimensional CFD analysis to study the thrust and efficiency of a biologically-inspired marine propulsor. Pol. Marit. Res. 2011, 18, 10–16. [Google Scholar] [CrossRef] [Green Version]
  52. Kumph, J.M. Maneuvering of a Robotic Pike. Mater’s Thesis, Department of Ocean Engineering, Massachussets Institute of Technology, Cambridge, MA, USA, 2000. [Google Scholar]
  53. Watts, C.M. A Comparison Study of Biologically Inspired Propulsion Systems for an Autonomous Underwater Vehicle. Ph.D. Thesis, University of Glasgow, Glasgow, UK, 2009. [Google Scholar]
  54. Watts, C.; McGookin, E.; Macauley, M. Biomimetic propulsion systems for mini-autonomous underwater vehicles. In Proceedings of the IEEE Conference OCEANS 2007, Vancouver, BC, Canada, 29 September–4 October 2007. [Google Scholar] [CrossRef]
  55. Watts, C.; McGookin, E. Modeling and simulation of a biomimetic underwater vehicle. In Proceedings of the Grand Challenges in Modeling and Simulation symposium, Edinburgh, UK, 16–19 June 2008; pp. 402–408. [Google Scholar]
  56. Hirata, K.; Takimoto, T.; Tamura, K. Study on turning performance of a fish robot. In Proceedings of the 1st International Symposium of Aqua Bio-Mechanisms, ISABMEC 2000, Tokai University Pacific Center, Honolulu, HI, USA, 27–30 August 2000. [Google Scholar]
  57. Heo, S.; Wiguna, T.; Park, H.C.; Goo, N.S. Effect of an artificial caudal fin on the performance of a biomimetic fish robot propelled by piezoelectric actuators. J. Bionic Eng. 2007, 4, 151–158. [Google Scholar] [CrossRef]
  58. Li, W.; Tianmiao, W.; Guanhao, W.; Jianhong, L. Hybrid undulatory kinematics of a robotic Mackerel (Scomber scombrus): Theoretical modeling and experimental investigation. Sci. China Technol. Sci. 2012, 55, 2941–2952. [Google Scholar] [CrossRef]
  59. Barrett, D.S.; Triantafyllou, M.S.; Yue, D.K.P. Drag reduction in fish-like locomotion. J. Fluid Mech. 1999, 392, 183–212. [Google Scholar] [CrossRef] [Green Version]
  60. Anderson, J.M.; Chabra, N.K. Maneuvering and stability performance of a robotic tuna. Integr. Comp. Biol. 2002, 42, 118–126. [Google Scholar] [CrossRef] [Green Version]
  61. Liang, J.; Wen, L.; Guo, Y. Experimental design and performance of underwater vehicle based on capacity of voyage. In Proceedings of the IEEE International Conference on Robotics, Automation and Mechatronics, Chengdu, China, 21–24 September 2008; pp. 587–591. [Google Scholar] [CrossRef]
  62. Liang, J.; Zheng, W.; Wen, L.; Wang, T.; Xie, C. Propulsive and maneuvering performance of two joints biorobotic autonomous undersea vehicle. In Proceedings of the IEEE International Conference on Robotics and Biomimetics, Guilin, China, 19–23 December 2009; pp. 314–320. [Google Scholar] [CrossRef]
  63. Liang, J.; Wang, T.; Wen, L. Development of a two-joint robotic fish for real-world exploration. J. Field Robot. 2011, 28, 70–79. [Google Scholar] [CrossRef]
  64. Wang, T.; Wen, L.; Liang, J.; Wu, G. Fuzzy vorticity control of a biomimetic robotic fish using a flapping lunate tail. J. Bionic Eng. 2010, 7, 56–65. [Google Scholar] [CrossRef]
  65. Liu, Y.X.; Chen, W.S.; Liu, J.K. Research on the swing of the body of two-joint robot fish. J. Bionic Eng. 2008, 5, 159–165. [Google Scholar] [CrossRef]
  66. Suleman, A.; Crawford, C. Design and testing of a biomimetic tuna using shape memory alloy induced propulsion. Comput. Struct. 2008, 86, 491–499. [Google Scholar] [CrossRef]
  67. Kruusmaa, M.; Fiorini, P.; Megill, W.; de Vittorio, M.; Akanyeti, O.; Visentin, F.; Chambers, L.; El Daou, H.; Fiazza, M.C.; Ježov, J.; et al. FILOSE for svenning: A flow sensing bioinspired robot. IEEE Robot. Autom. Mag. 2014, 21, 51–62. [Google Scholar] [CrossRef]
  68. Nakashima, M.; Tokuo, K.; Kaminaga, K.; Ono, K. Experimental study of a self-propelled two-joint dolphin robot. In Proceedings of the Ninth International Offshore and Polar Engineering Conference, Brest, France, 30 May–4 June 1999. [Google Scholar]
  69. Nakashima, M.; Ono, K. Development and experiment of two-joint dolphin robot. In Neurotechnology for Biomimetic Robots; MIT Press: Cambrige, MA, USA, 2002; pp. 309–324. [Google Scholar]
  70. Nakashima, M.; Takashi, Y.; Ono, K. Three dimensional manoeuverability of the dolphin robot. In Bio-Mechanisms of Swimming and Flying; Springer: Tokio, Japan, 2004; pp. 79–92. [Google Scholar] [CrossRef]
  71. Nakashima, M.; Tsubaki, T.; Ono, K. Three-dimensional movement in water of the dolphin robot—Control between two positions by roll and pitch combination. J. Robot. Mechatron. 2006, 18, 347–355. [Google Scholar] [CrossRef]
  72. Yu, J.; Hu, Y.; Fan, R.; Wang, L.; Kuzucu, A. Mechanical design and motion control of biomimetic robotic dolphin. Adv. Robot. 2007, 21, 499–513. [Google Scholar] [CrossRef]
  73. Yu, J.; Hu, Y.; Huo, J.; Wang, L. An adjustable Scotch Yoke mechanism for robotic dolphin. In Proceedings of the IEEE International Conference on Robotics and Biomimetics, Sanya, China, 15–28 December 2007; pp. 513–518. [Google Scholar] [CrossRef]
  74. Yu, J.; Hu, Y.; Huo, J.; Wang, L. Dolphin-like propulsive mechanism based on an adjustable Scotch yoke. Mech. Mach. Theory 2009, 44, 603–614. [Google Scholar] [CrossRef]
  75. Shen, F.; Wei, C.; Cao, Z.; Xu, D.; Yu, J.; Zhou, C. Implementation of a multi-link robotic dolphin with two 3-DOF flippers. J. Comput. Inf. Syst. 2011, 7, 2601–2607. [Google Scholar]
  76. Shen, F.; Cao, Z.Q.; Xu, D.; Zhou, C. A dynamic model of robotic dolphin based on Kane method and its speed optimization method. Acta Autom. Sin. 2012, 38, 1247–1256. [Google Scholar] [CrossRef]
  77. Liu, P.; He, K.; Ou, X.; Du, R. Mechanical design, kinematic modeling and simulation of a robotic dolphin. In Proceedings of the IEEE International Conference on Information and Automation, Shenzhen, China, 6–8 June 2011. [Google Scholar] [CrossRef]
  78. Dogangil, G.; Ozcicek, E.; Kuzucu, A. Design, construction, and control of a robotic dolphin. In Proceedings of the IEEE International Conference on Robotics and Biomimetics, Shatin, China, 5–9 July 2005; pp. 51–56. [Google Scholar] [CrossRef]
  79. Dogangil, G.; Ozcicek, E.; Kuzucu, A. Modeling, simulation, and development of a robotic dolphin prototype. In Proceedings of the IEEE International Conference on Mechatronics and Automation, Niagara Falls, ON, Canada, 29 July–1 August 2005; pp. 952–957. [Google Scholar] [CrossRef]
  80. Ho, T.; Lee, S. Design of a multi-locomotion underwater robot. Adv. Mater. Res. 2012, 488, 1732–1736. [Google Scholar] [CrossRef]
  81. Aghbali, B.; Yousefi-Koma, A. Design and fuzzy control of the shark robot-fish dorsal fin using SMA. In Proceedings of the 10th Biennial Conference on Engineering Systems Design and Analysis, Istanbul, Turkey, 12–14 July 2010. [Google Scholar] [CrossRef]
  82. Long, J.H., Jr.; Koob, T.; Schaefer, J.; Summers, A.; Bantilan, K.; Grotmol, S.; Porter, M. Inspired by sharks: A biomimetic skeleton for the flapping, propulsive tail of an aquatic robot. Mar. Technol. Soc. J. 2011, 45, 119–129. [Google Scholar] [CrossRef]
  83. Kodati, P.; Hinkle, J.; Deng, X. Micro autonomous robotic ostraciiform (MARCO): Design and fabrication. In Proceedings of the IEEE International Conference on Robotics and Automation, Rome, Italy, 14–15 April 2007. [Google Scholar] [CrossRef]
  84. Kodati, P.; Hinkle, J.; Winn, A.; Deng, X. Microautonomous robotic ostraciiform (MARCO): Hydrodynamics, design, and fabrication. IEEE Trans. Robot. 2008, 24, 105–111. [Google Scholar] [CrossRef]
  85. Gordon, M.; Hove, J.; Webb, P.; Weihs, D. Boxfishes as unusually well-controlled autonomous underwater vehicles. Physiol. Biochem. Zool. 2000, 74, 663–671. [Google Scholar] [CrossRef] [PubMed]
  86. Hu, Y.; Zhao, W.; Xie, G.; Wang, L. Development and target following of vision-based autonomous robotic fish. Robotica 2009, 27, 1075–1089. [Google Scholar] [CrossRef]
  87. Anderson, J.M.; Streitlien, K.; Barrett, D.S.; Triantafyllou, M.S. Oscillating foils of high propulsive efficiency. J. Fluid Mech. 1998, 360, 41–72. [Google Scholar] [CrossRef] [Green Version]
  88. Read, D.A.; Hover, F.S.; Triantafyllou, M.S. Forces on oscillating foils for propulsion and maneuvering. J. Fluids Struct. 2003, 17, 163–183. [Google Scholar] [CrossRef]
  89. Yamamoto, I.; Terada, Y.; Nagamatu, T.; Imaizumi, Y. Propulsion system with flexible/rigid oscillating fin. IEEE J. Ocean. Eng. 1995, 20. [Google Scholar] [CrossRef]
  90. Streitlien, K.; Triantafyllou, G.S. On thrust estimates for flapping foils. J. Fluids Struct. 1998, 12, 47–55. [Google Scholar] [CrossRef]
  91. Paterson, E.G.; Stern, F. Computation of unsteady viscous marine-propulsor blade flows—Part 1: Validation and analysis. J. Fluids Eng. Trans. ASME 1997, 119, 145–154. [Google Scholar] [CrossRef]
  92. Paterson, E.G.; Stern, F. Computation of unsteady viscous marine-propulsor blade flows—Part 2: Parametric study. J. Fluids Eng. Trans. ASME 1999, 121, 139–147. [Google Scholar] [CrossRef]
  93. Karpouzian, G.; Spedding, G.; Cheng, H.K. Lunate-tail swimming propulsion. Part 2. Performance analysis. J. Fluid Mech. 1990, 210, 329–351. [Google Scholar] [CrossRef]
  94. Yamaguchi, H.; Bose, N. Oscillating foils for marine propulsion. In Proceedings of the 4th International Offshore and Polar Engineering Conference, Osaka, Japan, 10–15 April 1994; Volume 3, pp. 539–544. [Google Scholar]
  95. Saimek, S.; Li, P.Y. Motion planning and control of a swimming machine. In Proceedings of the American Control Conference, Arlington, VA, USA, 25–27 June 2001; pp. 125–130. [Google Scholar] [CrossRef] [Green Version]
  96. Herr, H.; Dennis, B. A swimming robot actuated by living muscle tissue. J. Neuroeng. Rehabil. 2004, 1, 6–9. [Google Scholar] [CrossRef] [Green Version]
  97. Guo, S. A new type of fish-like underwater microrobot. IEEE/ASME Trans. Mechatron. 2003, 8, 136–141. [Google Scholar] [CrossRef]
  98. Belibassakis, K.A.; Politis, G.K. Hydrodynamic performance of flapping wings for augmenting ship propulsion in waves. Ocean. Eng. 2013, 72, 227–240. [Google Scholar] [CrossRef]
  99. Belibassakis, K.A.; Filippas, E.S. Ship propulsion in waves by actively controlled flapping foils. Appl. Ocean. Res. 2015, 52, 1–11. [Google Scholar] [CrossRef]
  100. Filippas, E.; Gerostathis, G.; Belibassakis, K.A. Semi-activated oscillating hydrofoil as a nearshore biomimetic energy device system in waves and currents. Ocean. Eng. 2018, 154, 396–415. [Google Scholar] [CrossRef]
  101. Koutsogiannakis, P.E.; Filippas, E.S.; Belibassakis, K.A. A study of multi-component oscillating-foil hydrokinetic turbines with a GPU-accelerated boundary element method. J. Mar. Sci. Eng. 2019, 7, 424. [Google Scholar] [CrossRef] [Green Version]
  102. Kato, N. Control performance in the horizontal plane of a fish robot with mechanical pectoral fins. IEEE J. Ocean. Eng. 2000, 25, 121–129. [Google Scholar] [CrossRef]
  103. Kato, N. Swimming and walking of an amphibious robot with fin actuators. Mar. Technol. Soc. J. 2011, 45, 181–197. [Google Scholar] [CrossRef]
  104. Licht, S.; Polidoro, V.; Flores, M.; Hover, F.S.; Triantafyllou, M.S. Design and projected performance of a flapping foil AUV. IEEE J. Ocean. Eng. 2004, 29, 786–794. [Google Scholar] [CrossRef]
  105. Licht, S.C.; Wibawa, M.S.; Hover, F.S.; Triantafyllou, M.S. In-line motion causes high thrust and efficiency in flapping foils that use power downstroke. J. Exp. Biol. 2010, 213, 63–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Kim, H.J.; Song, S.H.; Ahn, S.H. A turtle-like swimming robot using a smart soft composite (SSC) structure. Smart Mater. Struct. 2013, 22, 014007. [Google Scholar] [CrossRef]
  107. Salumae, T.; Raag, R.; Rebane, J.; Ernits, A.; Toming, G.; Ratas, M.; Kruusmaa, M. Design principle of a biomimetic underwater robot U-CAT. In Proceedings of the 2014 Oceans, St. John’s, NL, Canada, 14–19 September 2014; pp. 1–5. [Google Scholar] [CrossRef]
  108. Bozkurttas, M.; Mittal, R.; Dong, H.; Lauder, G.V.; Madden, P. Low-dimensional models and performance scaling of a highly deformable fish pectoral fin. J. Fluid Mech. 2009, 631, 311–342. [Google Scholar] [CrossRef] [Green Version]
  109. Bozkurttas, M.; Tangorra, J.; Lauder, G.; Mittal, R. Understanding the hydrodynamics of swimming: From fish fins to flexible propulsors for autonomous underwater vehicles. Adv. Sci. Technol. 2008, 58, 193–202. [Google Scholar] [CrossRef]
  110. Georgiades, C.; Nahon, M.; Buehler, M. Simulation of an underwater hexapod robot. Ocean. Eng. 2009, 36, 39–47. [Google Scholar] [CrossRef]
  111. Long, J.H., Jr.; Schumacher, J.; Livingston, N.; Kemp, M. Four flippers or two? Tetrapodal swimming with an aquatic robot. Bioinspiration Biomim. 2006, 1, 20–29. [Google Scholar] [CrossRef] [Green Version]
  112. Sitorus, P.E.; Nazaruddin, Y.Y.; Leksono, E.; Budiyono, A. Design and implementation of paired pectoral fins locomotion of labriform fish applied to a fish robot. J. Bionic Eng. 2009, 6, 37–45. [Google Scholar] [CrossRef]
  113. Cai, Y.; Bi, S.; Zheng, L. Design and experiments of a robotic fish imitating cow-nosed ray. J. Bionic Eng. 2010, 7, 120–126. [Google Scholar] [CrossRef]
  114. Punning, A.; Anton, M.; Kruusmaa, M.; Aabloo, A. A biologically inspired ray-like underwater robot with electroactive polymer pectoral fins. In Proceedings of the IEEE Conference on Mechatronics and Robotics, Aachen, Germany, 13–15 September 2004; pp. 241–245. [Google Scholar]
  115. Chen, Z.; Um, T.I.; Bart-Smith, H. Ionic polymer-metal composite enabled robotic manta ray. Electroact. Polym. Actuators Devices (EAPAD) 2011, 7976, 797637. [Google Scholar] [CrossRef]
  116. Clarck, R.P.; Smits, A.J. Thrust production and wake structure of a batoid-inspired oscillating fin. J. Fluid Mech. 2006, 562, 415–429. [Google Scholar] [CrossRef] [Green Version]
  117. Low, K.H. Locomotion simulation and system integration of robotic fish with modular undulating fin. Int. J. Simul. 2008, 7, 64–77. [Google Scholar]
  118. Zhang, Y.H.; Song, Y.; Yang, J.; Low, K.H. Numerical and experimental research on modular oscillating fin. J. Bionic Eng. 2008, 5, 13–23. [Google Scholar] [CrossRef]
  119. Zhang, Y.H.; He, J.H. Research on influence on fin ray motion pattern on the propulsion of bionic undulating fins. Chin. J. Eng. Des. 2017, 24, 89–99. [Google Scholar] [CrossRef]
  120. Low, K.H.; Willy, A. Biomimetic motion planning of an undulating robotic fish fin. JVC/J. Vib. Control 2006, 12, 1337–1359. [Google Scholar] [CrossRef]
  121. Low, K.H. Design, development and locomotion control of bio-fish robot with undulating anal fins. Int. J. Robot. Autom. 2007, 22, 88–99. [Google Scholar] [CrossRef]
  122. Low, K.H. Maneuvering of biomimetic fish by integrating a buoyancy body with modular undulating fins. Int. J. Hum. Robot. 2007, 4, 671–695. [Google Scholar] [CrossRef]
  123. Hu, T.; Shen, L.; Low, K.H. Bionic asymmetry: From amiiform fish to undulating robotic fins. Chin. Sci. Bull. 2009, 54, 562–568. [Google Scholar] [CrossRef]
  124. Low, K.H. Modelling and parametric study of modular undulating fin rays for fish robots. Mech. Mach. Theory 2009, 44, 615–632. [Google Scholar] [CrossRef]
  125. Epstein, M.; Colgate, J.E.; Maciver, M.A. Generating thrust with a biologically-inspired robotic ribbon fin. In Proceedings of the 2006 IEEE/RSJ International Conference on Intelligent Robots and Systems, Beijing, China, 9–15 October 2006; pp. 2412–2417. [Google Scholar] [CrossRef]
  126. MacIver, M.A.; Sharabash, N.M.; Nelson, M.E. Prey-capture behavior in gymnotid electric fish: Motion analysis and effects of water conductivity. J. Exp. Biol. 2001, 204, 543–557. [Google Scholar]
  127. MacIver, M.A.; Fontaine, E.; Burdick, J.W. Designing future underwater vehicles: Principles and mechanisms of the weakly electric fish. IEEE J. Ocean. Eng. 2004, 29, 651–659. [Google Scholar] [CrossRef]
  128. Siahmansouri, M.; Ghanbari, S.; Fakhrabadi, M.M.S. Design, implementation and control of a fish robot with undulating fins. Int. J. Adv. Robot. Syst. 2011, 8, 61–69. [Google Scholar] [CrossRef]
  129. Conte, J.; Modarres-Sadeghi, Y.; Watts, M.N.; Hover, F.S.; Triantafyllou, M.S. A fast-starting mechanical fish that accelerates at 40 m s−2. Bioinspiration Biomim. 2020, 5, 035004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Epps, B.P.; Techet, A.H. Impulse generated during unsteady maneuvering of swimming fish. Exp. Fluids 2007, 43, 691–700. [Google Scholar] [CrossRef]
  131. Dabiri, J.O. Optimal vortex formation as a unifying principle in biological propulsion. Annu. Rev. Fluid Mech. 2009, 41, 17–33. [Google Scholar] [CrossRef] [Green Version]
  132. Gharib, M.; Rambod, E.; Shariff, K. A universal time scale for vortex ring formation. J. Fluid Mech. 1998, 360, 121–140. [Google Scholar] [CrossRef] [Green Version]
  133. Krueger, P.S.; Gharib, M. The significance of vortex ring formation to the impulse and thrust of a starting jet. Phys. Fluids 2003, 15, 1271–1281. [Google Scholar] [CrossRef] [Green Version]
  134. Krueger, P.S. An over-pressure correction to the slug model for vortex ring circulation. J. Fluid Mech. 2005, 545, 427–443. [Google Scholar] [CrossRef]
  135. Krueger, P.S. Measurement of propulsive power and evaluation of propulsive performance from the wake of a self-propelled vehicle. Bioinspiration Biomim. 2006, 1, S49. [Google Scholar] [CrossRef]
  136. Ruiz, L.A.; Whittlesey, R.W.; Dabiri, J.O. Vortex-enhanced propulsion. J. Fluid Mech. 2011, 668, 5–32. [Google Scholar] [CrossRef] [Green Version]
  137. Guo, S.; Shi, L.; Ye, X.; Li, L. A new jellyfish type of underwater microrobot. In Proceedings of the International Conference on Mechatronics and Automation (ICMA 2007), Harbin, China, 5–8 August 2007; pp. 509–514. [Google Scholar] [CrossRef]
  138. Yang, Y.; Ye, X.; Guo, S. A new type of jellyfish-like microrobot. In Proceedings of the IEEE International Conference on Integration Technology (ICIT’07), Shenzhen, China, 20–24 March 2007; pp. 673–678. [Google Scholar] [CrossRef]
  139. Shi, L.; Guo, S.; Asaka, K. A novel jellyfish-like biomimetic microrobot. In Proceedings of the IEEE/ICME International Conference on Complex Medical Engineering (CME), Gold Coast, Australia, 13–15 July 2010; pp. 277–281. [Google Scholar] [CrossRef]
  140. Yeom, S.W.; Oh, I.K. A biomimetic jellyfish robot based on ionic polymer metal composite actuators. Smart Mater. Struct. 2009, 18, 085002. [Google Scholar] [CrossRef]
  141. Villanueva, A.; Smith, C.; Priya, S. A biomimetic robotic jellyfish (Robojelly) actuated by shape memory alloy composite actuators. Bioinspiration Biomim. 2011, 6, 036004. [Google Scholar] [CrossRef] [PubMed]
  142. Najem, J.; Sarles, S.A.; Akle, B.; Leo, D.J. Biomimetic jellyfish-inspired underwater vehicle actuated by ionic polymer metal composite actuators. Smart Mater. Struct. 2012, 21, 094026. [Google Scholar] [CrossRef] [Green Version]
  143. Krieg, M.; Mohseni, K. Thrust characterization of a bioinspired vortex ring thruster for locomotion of underwater robots. IEEE J. Ocean. Eng. 2008, 33, 123–132. [Google Scholar] [CrossRef] [Green Version]
  144. Krieg, M.; Mohseni, K. Dynamic modeling and control of biologically inspired vortex ring thrusters for underwater robot locomotion. IEEE Trans. Robot. 2010, 26, 542–554. [Google Scholar] [CrossRef]
  145. Serchi, F.G.; Arienti, A.; Laschi, C. A biomimetic, swimming soft robot inspired by the Octopus vulgaris. Lect. Notes Comput. Sci. 2012, 7375, 349–351. [Google Scholar] [CrossRef] [Green Version]
  146. Serchi, F.G.; Arienti, A.; Laschi, C. Biomimetic Vortex Propulsion: Toward the New Paradigm of Soft Unmanned Underwater Vehicles. IEEE/ASME Trans. Mechatron. 2013, 18, 484–493. [Google Scholar] [CrossRef]
  147. Van Buren, T.; Floryan, D.; Brunner, D.; Senturk, U.; Smits, A.J. Impact of trailing edge shape on the wake and propulsive performance of pitching panels. Phys. Rev. Fluids 2017, 2, 014702. [Google Scholar] [CrossRef]
  148. Shen, L.; Zhang, X.; Yue, D.K.P. Turbulent flow over a flexible wall undergoing a streamwise travelling wave motion. J. Fluid Mech. 2003, 484, 197–221. [Google Scholar] [CrossRef] [Green Version]
  149. Liu, H.; Kawachi, K. A numerical study of undulatory swimming. J. Comput. Phys. 1999, 155, 223–247. [Google Scholar] [CrossRef]
  150. Huera-Huarte, F.J. Propulsión acuática bio-inspirada basada en aleteo: Revisión y últimos avances. DYNA Ingeniería e Industria 2016, 91, 560–563. [Google Scholar] [CrossRef] [Green Version]
  151. Gazzola, M.; Argentina, M.; Mahadevan, L. Scaling macroscopic aquatic locomotion. Nat. Phys. 2014, 10, 758–761. [Google Scholar] [CrossRef] [Green Version]
  152. Taylor, G.K.; Nuds, R.L.; Thomas, A.L.R. Flying and swimming animals cruise at a Strouhal number tuned for high power efficiency. Nature 2003, 425, 707–711. [Google Scholar] [CrossRef] [PubMed]
  153. Rohr, J.; Fish, F. Strouhal numbers and optimization of swimming by odontocete cetaceans. J. Exp. Biol. 2004, 207, 1633–1642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Triantafyllou, G.S.; Triantafyllou, M.S.; Grosengaugh, M.A. Optimal thrust development in oscillating foils with application to fish propulsion. J. Fluids Struct. 1993, 7, 205–224. [Google Scholar] [CrossRef]
  155. Floch, F.; Phoemsatphawee, S.; Laurens, J.M.; Leroux, J.B. Porpoising foil as a propulsion system. Ocean. Eng. 2012, 39, 53–61. [Google Scholar] [CrossRef]
  156. Katz, J.; Weihs, D. Hydrodynamic propulsion by large amplitude oscillations of an airfoil with chordwise flexibility. J. Fluid Mech. 1978, 88, 485–497. [Google Scholar] [CrossRef]
  157. Jones, K.D.; Lai, J.C.S.; Tuncer, I.H.; Platzer, M.F. Computational and experimental investigation of flapping-foil propulsion. In Proceedings of the 1st International Symposium on Aqua Bio-Mechanisms/International Seminar on Aqua Bio-Mechanisms, Tokai University Pacific Center, Honolulu, HI, USA, 27–30 August 2000. [Google Scholar]
  158. Lan, C.E. The unsteady quasi-vortex-lattice method with applications to animal propulsion. J. Fluid Mech. 1979, 93, 747–765. [Google Scholar] [CrossRef]
  159. Isogai, K.; Shinmoto, Y.; Watanabe, Y. Effects of dynamic stall on propulsive efficiency and thrust of flapping airfoil. AIAA J. 1999, 37, 1145–1151. [Google Scholar] [CrossRef]
  160. Triantalyllou, M.S.; Triantafillou, G.S.; Yue, D.K.P. Hydrodynamics of fishlike swimming. Annu. Rev. Fluid Mech. 2000, 32, 33–53. [Google Scholar] [CrossRef]
  161. Lu, X.Y.; Yin, Z.X. Propulsive performance of a fish-like travelling wavy wall. Acta Mech. 2005, 175, 197–215. [Google Scholar] [CrossRef]
  162. Blake, R.W. The mechanics of labriform locomotion II. An analysis of the recovery stroke and the overall fin-beat cycle propulsive efficiency in the alglefish. J. Exp. Biol. 1980, 85, 337–342. [Google Scholar]
  163. Mattheijssens, J.; Marcel, J.P.; Bosschaerts, W.; Lefeber, D. Oscillating foils for ship propulsion. In Proceedings of the 9th National Congress on Theoretical and Applied Mechanics, Brussels, Belgium, 9–11 May 2012. [Google Scholar]
  164. Kim, J.Y.; Ghajar, A.J.; Tang, C.; Fouth, G.L. Comparison of near-wall treatment methods for high Reynolds number backward-facing step flow. Int. J. Comput. Fluid Dyn. 2005, 19, 493–500. [Google Scholar] [CrossRef]
  165. Tytell, E.D.; Hsu, C.Y.; Williams, T.L.; Cohen, A.H.; Fauci, L.J. Interactions between internal forces, body stiffness, and fluid environment in a neuromechanical model of lamprey swimming. Proc. Natl. Acad. Sci. USA 2010, 107, 19832–19837. [Google Scholar] [CrossRef] [Green Version]
  166. Tytell, E.D.; Borazjani, I.; Sotiropoulos, F.; Baker, T.V.; Anderson, E.J.; Lauder, G.V. Disentangling the functional roles of morphology and motion in the swimming of fish. Integr. Comp. Biol. 2010, 50, 1140–1154. [Google Scholar] [CrossRef] [Green Version]
  167. Bhalla, A.P.S.; Griffith, B.E.; Patankar, N.A. A forced damped oscillation framework for undulatory swimming provides new insights into how propulsion arises in active and passive swimming. PLoS Comput. Biol. 2013, 9, e1003097. [Google Scholar] [CrossRef] [Green Version]
  168. Cherukat, P.; Na, Y.; Hanratty, T.J.; McLaughlin, J.B. Direct numerical simulation of a fully developed turbulent flow over a wavy wall. Theor. Comput. Fluid Dyn. 1998, 11, 109–134. [Google Scholar] [CrossRef]
  169. Maass, C.; Schumann, U. Numerical Simulation of Turbulent Flow over a Wavy Boundary; Fluid Mechanics and Its Applications: Dordrecht, The Netherlands, 1994; pp. 287–297. [Google Scholar]
  170. You, D.; Wang, M.; Mittal, R.; Moin, P. Large-Eddy simulations of longitudinal vortices embedded in a turbulent boundary layer. AIAA J. 2006, 44, 3032–3039. [Google Scholar] [CrossRef]
  171. Hasse, C.; Sohm, V.; Wetzel, M. Hybrid URANS/LES turbulence simulation of vortex shedding behind a triangular flameholder. Flow Turbul. Combust. 2009, 83, 1–20. [Google Scholar] [CrossRef]
  172. Balaras, E. Modeling complex boundaries using an external force field on fixed Cartesian grids in large-eddy simulations. Comput. Fluids 2004, 33, 375–404. [Google Scholar] [CrossRef]
  173. Fauci, L.; Peskin, C.S. A computational model of aquatic animal locomotion. J. Comput. Phys. 1988, 77, 85–108. [Google Scholar] [CrossRef]
  174. Van Rees, W.M.; Gazzola, M.; Koumoutsakos, P. Optimal shapes for anguilliform swimmers at intermediate Reynolds numbers. J. Fluid Mech. 2013, 722, R3. [Google Scholar] [CrossRef] [Green Version]
  175. Tytell, E.D.; Leftwich, M.C.; Hsu, C.Y.; Griffith, B.E.; Cohen, A.H.; Smits, A.J.; Hamlet, C.; Fauci, L.J. Role of body stiffness in undulatory swimming: Insights from robotic and computational models. Phys. Rev. Fluids 2016, 1, 073202. [Google Scholar] [CrossRef]
  176. Griffith, B.E.; Hornung, R.D.; McQueen, D.M.; Peskin, C.S. An adaptive, formally second order accurate version of the immersed boundary method. J. Comput. Phys. 2017, 223, 10–49. [Google Scholar] [CrossRef]
  177. Griffith, B.E. An accurate and efficient method for the incompressible Navier–Stokes equations using the projection method as a preconditioner. J. Comput. Phys. 2009, 228, 7565–7595. [Google Scholar] [CrossRef]
  178. Akhtar, I.; Mittal, R.; Lauder, G.V.; Drucker, E. Hydrodynamics of a biologically inspired tandem flapping foil configuration. Theor. Comput. Fluid Dyn. 2007, 21, 155–170. [Google Scholar] [CrossRef]
  179. Gazzola, M.; Chatelain, P.; van Rees, W.M.; Koumoutsakos, P. Simulations of single and multiple swimmers with non-divergence free deforming geometries. J. Comput. Phys. 2011, 230, 7093–7114. [Google Scholar] [CrossRef]
  180. Borazjani, I.; Sotiropoulos, F. Numerical investigation of the hydrodynamics of carangiform swimming in the transitional and inertial flow regimes. J. Exp. Biol. 2008, 211, 1541–1558. [Google Scholar] [CrossRef] [Green Version]
  181. Dong, H.; Bozkurttas, M.; Mittal, R.; Madden, P.; Lauder, G.V. Computational modelling and analysis of the hydrodynamics of a highly deformable fish pectoral fin. J. Fluid Mech. 2010, 645, 345–373. [Google Scholar] [CrossRef]
  182. Borazjani, I.; Sotiropoulos, F.; Tytell, E.D.; Lauder, G.V. Hydrodynamics of the bluegill sunfish c-start escape response: Three-dimensional simulations and comparison with experimental data. J. Exp. Biol. 2012, 215, 671–684. [Google Scholar] [CrossRef] [Green Version]
  183. Mittal, R.; Iaccarino, G. Immersed boundary methods. Annu. Rev. Fluid Mech. 2005, 37, 239–261. [Google Scholar] [CrossRef] [Green Version]
  184. Eldredge, J.D. Numerical simulatons of undulatory swimming at moderate Reynolds number. Bioinspir. Biomim. 2006, 1, S19–S24. [Google Scholar] [CrossRef] [PubMed]
  185. Zhu, Q.; Yuming, L.; Yue, D.K.P. Dynamics of a three-dimensional oscillating foil near the free surface. AIAA J. 2006, 44, 2997–3009. [Google Scholar] [CrossRef]
  186. De Silva, L.W.A.; Yamaguchi, H. Numerical study on active wave devouring propulsion. J. Mar. Sci. Technol. 2012, 17, 261–275. [Google Scholar] [CrossRef] [Green Version]
  187. Filippas, E.S.; Belibassakis, K.A. Hydrodynamic analysis of flapping-foil thrusters operating beneath the free surface and in waves. Eng. Anal. Bound. Elem. 2014, 41, 47–59. [Google Scholar] [CrossRef]
  188. Filippas, E.S.; Papadakis, G.P.; Belibassakis, K.A. Free-surface effects on the performance of flapping-foil thruster for augmenting ship propulsion in waves. J. Mar. Sci. Eng. 2020, 8, 357. [Google Scholar] [CrossRef]
  189. Lauder, G.V.; Tytell, E.D. Hydrodynamics of undulatory propulsion. Fish Biomech. 2006, 23, 425–468. [Google Scholar] [CrossRef]
  190. Babu, M.N.P.; Krishnankutty, P.; Mallikarjuna, J.M. Experimental study of flapping foil propulsion system for ships and underwater vehicles and PIV study of caudal fin propulsors. In Proceedings of the IEEE/OES Autonomous Underwater Vehicles, Oxford, MS, USA, 6–9 October 2014. [Google Scholar] [CrossRef]
  191. Drucker, E.G.; Lauder, G.V. Locomotor forces on a swimming fish: Three-dimensional vortex wake dynamics quantified using digital particle image velocimetry. J. Exp. Biol. 1999, 202, 2393–2412. [Google Scholar] [PubMed]
  192. Drucker, E.G.; Lauder, G.V. A hydrodynamic analysis of fish swimming speed: Wake structure and locomotor force in slow and fast labriform swimmers. J. Exp. Biol. 2000, 203, 2379–2393. [Google Scholar]
  193. Drucker, E.G.; Lauder, G.V. Locomotor function of the dorsal fin in teleost fishes: Experimental analysis of wake forces in sunfish. J. Exp. Biol. 2001, 204, 2943–2958. [Google Scholar]
  194. Videler, J.J.; Muller, U.K.; Stamhuis, E.J. Aquatic vertebrate locomotion: Wakes from body waves. J. Exp. Biol. 1999, 202, 2430–3423. [Google Scholar]
  195. Videler, J.J.; Stamhuis, E.J.; Maijller, U.K.; van Duren, L.A. The scaling and structure of aquatic animal wakes. Integr. Comp. Biol. 2002, 42, 988–996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Muller, U.K.; Stamhuis, E.J.; Videler, J.J. Hydrodynamics of unsteady fish swimming and the effects of body size: Comparing the flow fields of fish larvae and adults. J. Exp. Biol. 2000, 203, 193–206. [Google Scholar] [PubMed]
  197. Stamhuis, E.J.; Videler, J.J. Quantitative flow analysis around aquatic animals using laser sheet particle image velocimetry. J. Exp. Biol. 1995, 198, 283–294. [Google Scholar] [PubMed]
  198. Wen, L.; Wang, T.M.; Wu, G.H.; Liang, J. Quantitative thrust efficiency of a self-propulsive robotic fish: Experimental method and hydrodynamic investigation. IEEE/ASME Trans. Mechatron. 2013, 18, 1027–1038. [Google Scholar] [CrossRef]
  199. Wen, L.; Wang, T.M.; Wu, G.H.; Liang, J. Hydrodynamic investigation of a self-propulsive robotic fish based on a force-feedback control method. Bioinspir. Biomim. 2012, 7, 036012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Anderson, E.J.; McGillis, W.R.; Grosengaugh, M.A. The boundary layer of swimming fish. J. Exp. Biol. 2001, 204, 81–102. [Google Scholar]
  201. Buckles, J.; Hanratty, T.J.; Adrian, R.J. Turbulent flow over a large-amplitude wavy surface. J. Fluid Mech. 1984, 140, 27–44. [Google Scholar] [CrossRef]
  202. Kuzan, J.D.; Hanratty, T.J.; Adrian, R.J. Turbulent flows with incipient separation over solid waves. Exp. Fluids 1989, 7, 88–98. [Google Scholar] [CrossRef]
  203. Hudson, J.D. The Effect of a Wavy Boundary on Turbulent Flow. Ph.D. Thesis, University of Illinoids Urbana, Champaign, IL, USA, 1993. [Google Scholar]
  204. Hudson, J.D.; Dykhno, L.; Hanratty, T.J. Turbulence production in flow over a wavy wall. Exp. Fluids 1996, 20, 257–265. [Google Scholar] [CrossRef]
Figure 1. Examples of fish swimming modes.
Figure 1. Examples of fish swimming modes.
Jmse 08 00479 g001
Figure 2. Main propulsion procedures from the view point of hydrodynamics; (a) Drag-based; (b) Added mass; (c) Momentum injection.
Figure 2. Main propulsion procedures from the view point of hydrodynamics; (a) Drag-based; (b) Added mass; (c) Momentum injection.
Jmse 08 00479 g002
Figure 3. Schematic representation of non-swimming locomotion.
Figure 3. Schematic representation of non-swimming locomotion.
Jmse 08 00479 g003
Figure 4. Velocity field overlaid with pressure field [50]; (a) BCF; (b) MPF.
Figure 4. Velocity field overlaid with pressure field [50]; (a) BCF; (b) MPF.
Jmse 08 00479 g004
Figure 5. Time variation of the average total, total, instantaneous, and viscous forces for anguilliform swimming [50].
Figure 5. Time variation of the average total, total, instantaneous, and viscous forces for anguilliform swimming [50].
Jmse 08 00479 g005
Figure 6. Average forces against cruising velocity for anguilliform swimming [50]. Reproduced or adapted from [50], with permission from Elsevier, 2020.
Figure 6. Average forces against cruising velocity for anguilliform swimming [50]. Reproduced or adapted from [50], with permission from Elsevier, 2020.
Jmse 08 00479 g006
Figure 7. Reynolds number for several organisms [151]. Reproduced or adapted from [151], with permission from Elsevier, 2020.
Figure 7. Reynolds number for several organisms [151]. Reproduced or adapted from [151], with permission from Elsevier, 2020.
Jmse 08 00479 g007
Figure 8. Reynolds against swimming number for several organisms [151].
Figure 8. Reynolds against swimming number for several organisms [151].
Jmse 08 00479 g008
Figure 9. Velocity field at the boundary layer and flow separation (in the last profile).
Figure 9. Velocity field at the boundary layer and flow separation (in the last profile).
Jmse 08 00479 g009
Figure 10. (a) Lagrangian treatment of the moving boundary; (b) Eulerian treatment of the moving boundary.
Figure 10. (a) Lagrangian treatment of the moving boundary; (b) Eulerian treatment of the moving boundary.
Jmse 08 00479 g010
Table 1. Summary of body and/or caudal fin (BCF) swimming mechanisms.
Table 1. Summary of body and/or caudal fin (BCF) swimming mechanisms.
AuthorInstitutionPropulsion ModeSource of Inspiration
Cohen [32]Cornell University, USAAnguilliformLamprey
Grillner and Matsushima [33]; Grillner et al. [34]Karolinska Institute, SwedenAnguilliformLamprey
Ayers et al. [35,36,37]; Wilbur et al. [38]; Westphal et al. [39]Northeastern University, USAAnguilliformLamprey
Liu et al. [40]Zhejiang University, ChinaAnguilliformLamprey
Xu et al. [41]National University of Singapore, SingaporeAnguilliformEel/lamprey
Crespi et al. [42]Swiss Federal Institute of Technology, SwitzerlandAnguilliformSea snake
Ijspeert et al. [43]; Ijspeert et al. [44]; Crespi et al. [45]Swiss Federal Institute of Technology, SwitzerlandAnguilliformSalamander
McIsaac and Ostrowski [46,47,48]University of Western Ontario, CanadaAnguilliformEel
Lamas et al. [49,50,51]University of A Coruña, SpainAnguilliformEel
Kumph [52]Massachussets Institute of Technology, USACaranguiformPike
Watts et al. [53,54,55]University of Glasgow, United KingdomCaranguiformSalmon
Hirata et al. [56]Ship Research Institute Shinkawa, JapanCaranguiformSea bream
Heo et al. [57]
Li et al. [58]
Konkuk University, Korea
Beihang University, China
Caranguiform
Hybrid
Mackerel
Mackerel
Barrett [11]; Barrett et al. [59]; Anderson and Chabra [60]Massachussets Institute of Technology, USAThunniformTuna
Liang et al. [61]; Liang et al. [62]; Liang et al. [63]; Wang et al. [64]; Liu et al. [65]Beihang University, ChinaThunniformTuna
Suleman and Crawford [66]
Kruusmaa et al. [67]
University of Victoria, Canada
Tallinn University of Technology, Estonia
Thunniform
Thunniform
Tuna
Tuna
Nakashima et al. [68]; Nakashima et al. [69]; Nakashima et al. [70]; Nakashima et al. [71]Tokyo Institute of Technology, JapanThunniformDolphin
Yu et al. [72]; Yu et al. [73]; Yu et al. [74]; Shen et al. [75]; Shen et al. [76]; Liu et al. [77]Chinese Academy of Sciences, ChinaThunniformDolphin
Dogangil et al. [78]; Dogangil et al. [79]Istanbul Technical University, TurkeyThunniformDolphin
Ho and Lee [80]Konkuk University, South KoreaThunniformDolphin
Aghbali et al. [81]College of Engineering, IranThunniformShark
Long et al. [82]Vassar College, USAThunniformShark
Kodati et al. [83]; Kodati et al. [84]University of Delaware, USAOstraciiformBoxfish
Gordon et al. [85]University of California, USAOstraciiformBoxfish
Hu et al. [86]Peking University, ChinaOstraciiformBoxfish
Anderson et al. [87]Massachussets Institute of Technology, USAOstraciiformFlapping foil
Read et al. [88]Massachussets Institute of Technology, USAOstraciiformFlapping foil
Yamamoto et al. [89]Mitsubishi Heavy Industries, JapanOstraciiformFlapping foil
Streitlien and Triantafyllou [90]City College of New York, USAOstraciiformFlapping foil
Paterson and Stern [91,92]University of Lowa, USAOstraciiformFlapping foil
Karpouzian et al. [93]Univ of Southern California, USAOstraciiformFlapping foil
Yamaguchi and Bose [94]University of Tokyo, JapanOstraciiformFlapping foil
Saimek and Li [95]University of Minnesota, USAOstraciiformFlapping foil
Herr and Dennis [96]MA Institute of Technology, USAOstraciiformFlapping foil
Guo [97]Kagawa University, JapanOstraciiformFlapping foil
Belibassakis and Politis [98]; Belibassakis and Filippas [99]; Filippas et al. [100]; Koutsogiannakis et al. [101]National Technical University of Athens, GreeceOstraciiformFlapping foil
Table 2. Summary of median and paired fin (MPF) swimming mechanisms.
Table 2. Summary of median and paired fin (MPF) swimming mechanisms.
AuthorInstitutionPropulsion ModeSource of Inspiration
Kato et al. [102]Tokai University, JapanLabriformBlack bass
Kato et al. [103]Tokai University, JapanLabriformSea turtle
Litch et al. [104,105]Massachussets Institute of Technology, USALabriformSea turtle
Kim et al. [106]
Salumae et al. [107]
Seoul National University, South Korea
Tallinn University of Technology, Estonia
Labriform
Labriform
Sea turtle
Sea turtle
Bozkurttas et al. [108,109]Franklin W. Olin College of Engineering, USALabriformSunfish
Georgiades et al. [110]McGill University, CanadaLabriformFlapping foil
Long et al. [111]Vassar College, USALabriformFlapping foil
Sitorus et al. [112]Bandung Institute of Technology, IndonesiaLabriformFlapping foil
Cai et al. [113]Beihang University, ChinaRajiformRay
Punning et al. [114]Tartu Univeristy, EstoniaRajiformRay
Chen et al. [115]University of Virginia, USARajiformRay
Clark and Smiths [116]Princeton University, USARajiformRay
Low, K.H. [117]; Zhang et al. [118,119]Nanyang Technological University, SingaporeRajiformRay
Low and Willy [120]Nanyang Technological University, SingaporeRajiform/gymnotiformRay/knifefish
Low [121]; Low [122]; Hu et al. [123]; Low [124]Nanyang Technological University, SingaporeGymnotiformKnifefish
Epstein et al. [125]Northwestern University, USAGymnotiformKnifefish
McIver et al. [126,127]Northwestern University, USAGymnotiformKnifefish
Siahmansouri et al. [128]University of Tabriz, IranGymnotiformKnifefish
Table 3. Summary of non-swimming mechanisms.
Table 3. Summary of non-swimming mechanisms.
AuthorInstitutionPropulsion ModeSource of Inspiration
Gharib et al. [132]; Krueger and Gharib [133]; Krueger [134], Krueger [135], Ruiz et al. [136]California Institute of Technology, USAJetPiston-cylinder
Guo et al. [137]; Yang et al. [138]; Shi et al. [139]Kagawa University, JapanJetJellyfish
Yeom and Oh [140]Chonnam National University, South KoreaJetJellyfish
Villanueva et al. [141]; Najem et al. [142]Virginia Tech, USAJetJellyfish
Krieg and Mohseni [143]; Krieg and Mohseni [144]University of Colorado, USAJetSquid
Serchi et al. [145]; Serchi et al. [146]Scuola Superiore Sant’Anna, ItalyJetOctopus
Table 4. Reference and non-dimensional parameters.
Table 4. Reference and non-dimensional parameters.
DimensionReference ParameterNon-Dimensional Parameter
Length L r e f = L L * = L / L r e f
Velocity u r e f = U u * = u / u r e f
Pressure p r e f = ρ u r e f 2 p * = p / p r e f
Time t r e f = L r e f / u r e f = L / U t * = t / t r e f
Gravity g r e f = g g * = g / g r e f

Share and Cite

MDPI and ACS Style

Lamas, M.I.; Rodriguez, C.G. Hydrodynamics of Biomimetic Marine Propulsion and Trends in Computational Simulations. J. Mar. Sci. Eng. 2020, 8, 479. https://doi.org/10.3390/jmse8070479

AMA Style

Lamas MI, Rodriguez CG. Hydrodynamics of Biomimetic Marine Propulsion and Trends in Computational Simulations. Journal of Marine Science and Engineering. 2020; 8(7):479. https://doi.org/10.3390/jmse8070479

Chicago/Turabian Style

Lamas, M. I., and C. G. Rodriguez. 2020. "Hydrodynamics of Biomimetic Marine Propulsion and Trends in Computational Simulations" Journal of Marine Science and Engineering 8, no. 7: 479. https://doi.org/10.3390/jmse8070479

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop