You are on page 1of 16

INTEG.

AND

COMP. BIOL., 42:102117 (2002)

Maneuvering Hydrodynamics of Fish and Small Underwater Vehicles1


PROMODE R. BANDYOPADHYAY2
Propulsion, Hydrodynamics and Silencing Division,
Naval Undersea Warfare Center, Newport, Rhode Island 02841

INTRODUCTION
The engineering community generally believes that
man made vehicles and machines are matured in development for steady state operation; biology-inspired
further improvements might not be cost effective.
However, the following two products are recent examples where induced drag of wing-tip vortices has
been reduced with devices inspired by winglets of
soaring eagles: the Spiroid Wing Tip of Gulfstream II
aircraft developed at Boeing, and the propellor with

tip-less round blades, by Bannasch (2000). Moreover,


the following developments have opened up opportunities in unsteady operations, namely in maneuvering,
in addition to, perhaps, in propulsion: our new understanding of the mechanisms of lift enhancement via
unsteady vortex dynamics, the advances in digital control, control theory and active materials technology
(Ellington, 1984; Dickinson et al., 1999; Bandyopadhyay, 1999; Bar-Cohen, 2001; Madden et al., 2001).
This paper reviews the related progress made at
NUWC. The science of maneuvering and stealth in
swimming and flight in nature is distilled and applied
to conventional rigid bodied and generic underwater
vehicles. To name a few, the present work rests on the

1 From the Symposium Stability and Maneuverability presented at


the Annual Meeting of the Society for Integrative and Comparative
Biology, 37 January 2001, at Chicago, Illinois.
2 Present address: Code 342 Cognitive & Neural S&T Division,
Office of Naval Research, 800 N. Quincy St., Arlington, Virginia
22217-5660. E-mail: bandyop@onr.navy.mil

FIG. 1.

Definition of length scales of a fish.

FIG. 2.

102

Morphology of dorsal fins of fish families.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

SYNOPSIS.
The understanding of fish maneuvering and its application to underwater rigid bodies are
considered. The goal is to gain insight into stealth. The recent progress made in NUWC is reviewed. Fish
morphology suggests that control fins for maneuverability have unique scalar relationships irrespective of
their speed type. Maneuvering experiments are carried out with fish that are fast yet maneuverable. The
gap in maneuverability between fish and small underwater vehicles is quantified. The hydrodynamics of a
dorsal fin based brisk maneuvering device and a dual flapping foil device, as applied to rigid cylindrical
bodies, are described. The role of pectoral wings in maneuvering and station keeping near surface waves is
discussed. A pendulum model of dolphin swimming is presented to show that body length and tail flapping
frequency are related. For nearly neutrally buoyant bodies, Froude number and maneuverability are related.
Analysis of measurements indicates that the Strouhal number of dolphins is a constant. The mechanism of
discrete and deterministic vortex shedding from oscillating control surfaces has the property of large amplitude unsteady forcing and an exquisite phase dependence, which makes it inherently amenable to active
control for precision maneuvering. Theoretical control studies are carried out to demonstrate the feasibility
of maneuverability of biologically inspired bodies under surface waves. The application of fish hydrodynamics to the silencing of propulsors is considered. Two strategies for the reduction of radiated noise are developed. The effects of a reduction of rotational rate are modeled. The active cambering of blades made of
digitally programmable artificial muscles, and their thrust enhancement, are demonstrated. Next, wake
momentum filling is carried out by artificial muscles at the trailing edge of a stator blade of an upstream
stator propulsor, and articulating them like a fish tail. A reduction of radiated noise, called blade tonals, is
demonstrated theoretically.

MANEUVERING HYDRODYNAMICS

Reciprocal of aspect ratio of dorsal fin.

theoretical foundations of fish locomotion laid by Taylor (1952), Lighthill (1975) and Wu (1971ad, 1972),
the numerical studies by Webb (1975) and the scaling
laws of Triantafyllou and Triantafyllou (1995).
LOW-SPEED MANEUVERING DYNAMICS OF FISH AND
SMALL UNDERWATER VEHICLES
Morphology of control surfaces
While biologists tend to attach importance to variations of a theme, engineers have a need to simplify,
that is to minimize the variations, to arrive at a robust
application. This probably is a reflection of the fact

FIG. 4.

FISH

AND

SUVS

103

that, in general, in a comparable environment, emerging materials, actuators and their control are far from
being as dynamically competent as those in living animals. It is reasonable to assume that even the static
morphology of a fish can provide clues to its locomotion, habitat and ecology (see among others, Lighthill, 1975; Aleyev, 1977; Bandyopadhyay et al.,
1997). As a starting point, the simplified morphology
of various fish families is examined to determine what
makes some families more maneuverable than others.
Maneuverability is defined as the minimum or commonly observed turning radius at a given normal acceleration. As suggested here, internal Froude number
may be interpreted as a measure of maneuverability.
Ability to make a sudden stop or start is also a measure
of maneuvering ability although this is not considered
here.
Length scales of the body and fins of a fish defined
in Figure 1 are examined. Twenty-eight species of fish
are considered. They are classified into three categories: low-speed highly maneuverable, high speed poorly maneuverable, and an overlapping category. The relationship between fin morphology and the characteristics of maneuvering is shown in Figures 2 and 3.
Several definite trends are observed. The result in Figure 3 agrees with observation that a cylinder, whose
length to diameter ratio is less than 10, tends to be
unstable. In the next section, a dorsal fin device for
brisk maneuvering based on the result in Figure 2, is
examined in the context of a rigid cylinder having a
large aspect ratio.

The example of a digitized bluefish trajectory in the rectangular tank with pipe and cinder block maze.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 3.

OF

104

(a) Coefficient of normal acceleration. Solid lines: Eq. (1). (b) Reduced coefficient of normal acceleration. Solid lines: Eq. (3).

Gap in maneuverability
An experiment was carried out comparing the maneuverability of fish and small underwater vehicles to
quantitatively establish the gap. Bluefish and mackerel,
which are oceanic fast swimmers and yet are maneuverable, were selected for an experiment on maneuverability. Swimming tanks with baffles, as shown in
Figure 4, were designed to photograph their trajectories. The turning dynamics were determined from the

digitized trajectories. The results of coefficient of normal acceleration Cg versus turning radius r/L were
compared with two small underwater vehicles as in
Figure 5a. Here, Cg 5 (V2/r)/g is acceleration perpendicular to the path, V is total velocity, r is the radius
of curvature in trajectory, and g is acceleration due to
gravity. There is a large gap between the maneuvering
capability of fish and the vehicles. There is a universal
trend, namely,

TABLE 1. Relationship between lengths of dolphins and their natural frequencies. From Bandyopadhyay et al. (2000a).
Author of Data (in Rohr et al., 1998)

Dolphin Specie
Range of Length L (cm)
Average Length L (cm)
Natural Frequency n (Hz) due to Pendulum
Model

Lang and Daybell

Rohr et al.

Fish

Lagenorhynchus
Obliquidens
209
209

Tursiops
Truncatus
182283
240

Tursiops
Truncatus
251270
260

0.32

0.31

0.34

Rohr et al.

Pseudorca
Crassidens
365
365
0.26

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 5.

PROMODE R. BANDYOPADHYAY

MANEUVERING HYDRODYNAMICS

OF

FISH

AND

SUVS

105

Cg 5 (r/L)21,

(1)

which is followed but in large turn radii only. Compared


to underwater vehicles, fish can make the same radius
turn at a normal acceleration that is lower by a factor
that can be as large as 10. Lower speed and laminar flow
could allow a fish to make stealthy maneuvers.
Froude number and maneuverability
The scaling law in Figure 5a can be improved so
that the wide Reynolds number range between natural
and man-made bodies is covered. Propose that the coefficient of normal acceleration during a turn is a func-

tion of inertia forces, viscous forces and gravity forces.


Define Re 5 VL/y, and an internal Froude number Fr
5 V/gL, where V is speed, L is length, y is kinematic
viscosity, and g is acceleration due to gravity. These
two ratios can be combined as:
Fr 4
V3 y
5 3 2.
Re
L g

(2)

This combined parameter can be used to rescale the


coefficient of normal acceleration shown in Figure 5a,
and the result is shown in Figure 5b. The solid lines
in Figure 5b are:

FIG. 7. Comparison of side thrust between abruptly cambered (left) and permanently cambered (right) fins at a steady tow speed of 3.6 m/
sec. The ordinates are in arbitrary volt scale.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 6. Photograph of cylinder and dorsal fin assembly. The solenoid-cam arrangement inside the cylinder used to deploy the fins is also
visible in the middle.

106

PROMODE R. BANDYOPADHYAY

Photograph of an euthynnid (due to Magnuson). The pectoral fins are swept back to change planform area and lift.

21

12

Cg
r
;
n V3
L
g 2 L3

(3)

The inverse power trend (3) is now followed over a


greater range of turn radius in both fish and vehicle
data. The significance of internal Froude number is as
follows. Rewrite Fr as:
Fr 5

V
gL

1!g 2 L 5 2p L .
L V

1 T



V

(4)

In other words, this internal Froude number is a ratio


of the time period of natural oscillation T like a pendulum, and the time it takes for the body to travel a
distance equal to its length. This is reminiscent of
Froude number that is relevant to wave drag of surface
ships. It is interesting that the jets produced by fish,
dolphins or whales for propulsion is similar to the recent flow visualization of fluttering of long and light
strips in a liquid by Belmonte et al. (1998) (also see

FIG. 9. Dimensions of a reference plain cylinder and two sets of


wings used to model the maneuvering ability of winged bodies. The
tail wings are always retained for stability. The dimensions are similar to the tow tank model described later.

Aleyev, 1977; Videler, 1993). (Lightness can be considered equivalent to near neutral buoyancy for submerged bodies.) On the other hand, heavy and short
strips tumble. Note that they also similarly define
an (internal) Froude number, which is the reciprocal
of the one given above. According to Belmonte et al.
(1998), Fr defined as in the present work, have a high
value for long and light strips which flutter. Thus, fish,
dolphins and whales have high Froude numbers. On
the other hand, short heavy animals like the beetle of
Fish (1999) have a low value of Fr. Thus, we reinterpret the conclusion of Fish that flexibility of a humpback whale gives it higher maneuverability than that
of a whirligig beetle which is rigid. We conclude that
internal Froude number is related to maneuverability.
The commonality of the wake pattern and the terms
in the Froude number for fluttering objects suggests

FIG. 10.

Reduction of required angles of attack in winged bodies.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 8.

MANEUVERING HYDRODYNAMICS

OF

FISH

AND

SUVS

107

with the calculated values. This length consideration


accounts for the seeming scatter in the data. Thus,
(buoyancy or lightness), length (slenderness) and
Froude number appear to play important roles in maneuvering.
Strouhal number of dolphins

that a dolphin can be modeled as a simple pendulum.


The dolphin swimming data due to Rohr et al. (1998)
was examined (Bandyopadhyay et al., 2000a). Table
1 is a summary of the data sets based on lengths. It
was proposed that the jet responsible for dolphin propulsion is analogous to the jet due to the predominantly side to side motion of fluttering long or light
strips when dropped freely in air or water. The natural
frequencies of dolphins calculated based on the following pendulum model is shown in Table 1:

@1 ! 2

n51

2p

L
.
g

(5)

When dolphin-swimming data based on their length


are extrapolated to zero swimming speed, they agree

St 5 0.22.

(6)

Thus, dolphins also have a similar and constant Strouhal number as fish, although their Reynolds number is
much higher. This provides another layer of evidence
that fish, dolphins and whales have similar Froude
numbers and mechanisms of maneuvering and propulsion, irrespective of their Reynolds numbers.
FISH-INSPIRED CONTROL FOR MANEUVERING:
DORSAL FIN DEVICE
The maneuverability of normally, stable cylinders
(length/diameter $ 10) is considered here. Their maneuverability is a slow process when it involves the
pitching of small fins near the boat tails and a subsequent large-scale separation of the entire cylinder. A

FIG. 12. Schematic of a surface wave propagating right to left over a submerged body. The velocity field and elliptic particle motion due to
finite depth of water and time signature of horizontal (Fx) and vertical (Fz) forces acting on the body are shown.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 11. Compared to their body lengths, winged cylinders make


a shorter radii turn.

It is known indirectly that fish propulsion takes


place predominantly in the Strouhal number range of
0.25 , St , 0.35, where St 5 fA/U (Fish and Rohr,
1999). Here, f and A are frequency and amplitude of
tail oscillation, and U is speed of fish motion. For dolphins, measurements due to Rohr et al. (1998) indicate
that the amplitude of oscillation is given by A/L 5 0.2
6 0.02 and it is independent of speed. When speed is
expressed as body length traversed per second, then
frequency can be expressed as: f 5 1.1(U/L). This is
supported by Rohr et al.s (1998) data, the mean trend
of Kayan and Pyatetskiy (introducing f 5 0 at U 5 0)
and the accounting of natural frequency (see Fig. 5 in
Bandyopadhyay et al., 2000a). With the above two
relationships, for dolphins, we then have

108

PROMODE R. BANDYOPADHYAY

76 mm diameter and 754 mm long cylinder model


with a dorsal fin device that is abruptly deployable,
and shown in Figure 6, was constructed. The model
was towed in a tank and a typical result of the production of side thrust is shown in Figure 7. It is clear
that the dorsal fin when cambered abruptly produces
large levels of side thrust practically immediately.
The theoretical study of biologically inspired control
surfaces described later showed that the above result
is of general significance. While man made vehicles,
like aircraft, have a moment-based control, biologically based maneuvering of engineering bodies would be
force based. One consequence of the latter is the faster
under water response allowing a greater agility.

FISH-INSPIRED CONTROL FOR MANEUVERING:


SMALL AGILE VEHICLES
Pectoral fins and turning. Several species of fast yet
agile fish (Fig. 8) have large pectoral fins. They do not
have a gas bladder and retract these fins to control lift
force. A detailed hydrodynamic coefficients based
modeling was carried out to determine the effectiveness of these pectoral fins in low speed maneuvering
in an engineering context. The computational modeling of cruise and turn was carried out on three configurations of the cylinder shown in Figure 9: a reference
plain cylinder and two others where pairs of wings, of
the order of cylinder diameter, are attached. All three
cylinders are provided with a pair of tail planes for
stability. The results are shown in Figures 10 and 11.
The wings allow the cylinder to be sustained at lower
angles of attack. They also allow the cylinder to make
lower radii turns.
Damping due to pectoral fins near surface waves

FIG. 14. Variation of the absolute values of the maximum or minimum of the coefficients of axial forces and pitching moments with
towing speed. Note reduction due to hydrofoil.

The effect of the pectoral wings was examined in


presence of travelling surface waves. This geometry
and the forces on a body due to linear theory are
shown in Figure 12. The model shown in Figure 13
was constructed, where the wing was offset from the
cylinderbelow it and not above. The maximum and
minimum values of the periodic coefficients of axial
force and pitching moments are compared in Figure
14 between the hydrofoil-cylinder and the plain cylinder case. Here, coefficient of axial force Cfx 5 Fx/
(rgbAf), and coefficient of pitching moment CTy 5 Ty/
(rgbAp D), where Fx is axial force, Ty is pitching moment, r is fluid density, g is acceleration due to gravity,
b is peak-to-trough wave height, Af is frontal area, Ap
is planform area, and D is the offset of the hydrofoil
leading edge from the model axis (511.43 cm for the
hydrofoil-cylinder model and D/2 for the plain cylin-

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 13. Photograph of instrumented laboratory scale model for evaluation of fish inspired control surfaces on a rigid cylinder. This control
surface is passive and is an euthynnid-inspired offset pectoral winglet for pitch stabilization near surface waves.

MANEUVERING HYDRODYNAMICS

der). The pectoral wings have a stabilizing damping


effect.
Theoretical control study of biologically-inspired
control surfaces. The control system synthesis of a
small cylinder equipped with a pair of dorsal and caudal fins, and in the presence of surface waves, as
shown in Figure 15, was examined. Closed loop control laws were derived using the dorsal and caudal fins
for depth and pitch control, respectively. The system
is shown in Figure 16. For the typical cylinder geometries considered here, Figure 17 shows an example
of a simulation where precise depth control and pitch
regulation were achieved using the dorsal fin only.
FISH-INSPIRED CONTROL FOR MANEUVERING:
DUAL FLAPPING FOILS
Unsteady vortex mechanism
The remaining gap in turning ability depicted in
Figure 5 between fish and rigid bodies can be attri-

FIG. 16. Closed-Loop System (Including the Caudal and Dorsal Fin
Controllers).

FISH

AND

SUVS

109

FIG. 17. Dorsal Fin Control: Sinusoidal Disturbance: (a) Depth error ze 5 z 2 yr; (b) Depth z and reference command yr; (c) Camber
d 5 u; (d) Pitch angle.

buted primarily to the absence of sufficient control surfaces and perhaps to flexibility of the main body in
the latter category (Fish, 1999). Fish-inspired control
surfaces can therefore be a partial solution for maneuverability for rigid bodies. The hydrodynamic mechanism of maneuvering and stealth of the moving surfaces of a fish was examined in an engineering experiment. The head and tail oscillations of a fish were
simulated on a rigid cylinder that also removed the
added complication of body undulation. Several experiments were carried out on the strut mounted floating cylinder shown in Figure 18. A six-component dynamic load cell, three actuators and two displacement
sensors for measuring the phase of the tail actuators
were housed within the model. Dye visualization and
phase-matched laser Doppler velocimetry measurements of the vortex shedding due to the oscillation of
the tail flaps, and time histories of the forces and moments on the entire model assembly were carried out

FIG. 18. Schematic of model for studying the following effects: (1)
the precision maneuvering ability of dual flapping foils attached to
the tail of a rigid cylinder, and (2) the interference between simulated
head motion and tail flapping of a fish on a rigid cylinder. A software
operated digital controller is used to select the phase lag of the nose
slider actuator relative to the two flap actuators that operate in phase,
called waving mode here (as opposed to clapping mode where they
operate in anti-phase).

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 15. Schematic of the maneuvering devices (Dorsal and Caudal


Fins) and axisymmetric cylinder. Dorsal fins, shown uncambered,
are mounted in the horizontal plane. The caudal fins, mounted in
the horizontal plane, are akin to flukes in whales. XI 2 ZI 5 Inertial
Coordinate System (Origin at the Calm Surface). X9I 2 Z9I 5 Translation of Inertial Frame (Origin at Geometrical Center). XB 2 ZB 5
Body Fixed Coordinate System.

OF

110

PROMODE R. BANDYOPADHYAY

for an uniform freestream and various tail oscillation


Strouhal numbers. In the first configuration, only the
pair of flapping foils attached to the tail was oscillated
and the slider in the nose was absent. In the second,
the interaction of both the nose and tail oscillations
was examined. Figures 19 and 20 show the photographs of the instrumented model in these two configurations, respectively.

Figure 21 shows the measurements of time integrated axial (thrust) force coefficient (ca) versus Strouhal number (St) of tail flap oscillation, They are defined as: ca 5 F/(1/2 rU`2D2) and St 5 fA/U`. Here, F
is axial force, r is fluid density, U` is freestream velocity, d is length scale of flap, f and A are flap oscillation frequency and amplitude, repectively. The data
asymptotes to Lighthills two-dimensional inviscid the-

FIG. 20. Photograph of cylinder model for studying the interference effects between head and tail oscillations, and digital controller of
actuators.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 19. Photograph of water tunnel model of the dual flapping foil device. Dual flapping foils and divider plate are shown at the right end;
to the left of foils lie the actuators, two phase transducers, and actuator control circuits. The six-component load cell is located at the junction
of the strut and cylinder.

MANEUVERING HYDRODYNAMICS

AND

SUVS

111

Comparison of measurements of axial force (thrust) components.

ory at Strouhal numbers that are below the range of


fish. We believe that Lighthills line indicates the natural shedding symptote. As St is increased, axial force
generated becomes oscillation frequency dependent,
approaching the forced shedding asymptote of Bandyopadhyay (1996).
The pairs of tail flaps (Figs. 18 and 19) were oscillated in two modes termed waving and clapping. Their

FIG. 22.

FISH

phase is the same and opposite in them, respectively,


mimicking the motion the names indicate. The respective vorticity-velocity vector maps of the vortex shedding process in the axial plane are shown in Figures
22 and 23, where t* 5 tU`/D is phase and the locations
of the flap trailing edges are indicated by two small
filled squares on the vertical axis. The large crossstream forces in waving mode (Fig. 22), which have

Vorticity-velocity vector maps in the axial plane (x/D versus 6 y/D) in the waving mode for phase t*.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 21.

OF

112

PROMODE R. BANDYOPADHYAY

Vorticity-velocity vector maps in the axial plane (x/D versus 6 y/D) in the clapping mode for phase t*.

maneuvering and axial components, owe their origin


to the formation of a staggered vortex train. On the
other hand, the clapping mode which produces a pure
axial thrust only (Fig. 23), owes its origin to symmetric vortex trains which are mirror images of each other.

FIG. 24.

The resulting induced velocity between the successive


vortices in the waving (maneuvering) mode is inclined
to the streamwise axis. On the other hand, it is a perfectly aligned streamwise jet in the clapping mode.
The general conclusion is that the mechanism of dis-

Decay of circulation distribution in the wake of flapping foils: waving mode of flapping. The horizontal axis represents flap phase.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 23.

MANEUVERING HYDRODYNAMICS

crete and deterministic vortex shedding from oscillating control surfaces has the property of large amplitude unsteady forcing and an exquisite phase dependence, which makes it inherently amenable to active
control for precision maneuvering.
Decay of wake. The vorticity-velocity vector maps
were used to compute the circulation in the shed vortices by two methods: velocity line and vorticity area
integrals. The circulation distributions are compared
immediately after formation and after a short travel in
Figure 24, for the waving mode. The maximum value
of circulation (2G) drops by a factor of 3 within a
mere distance of half the body diameter or flap width
(D). The effect is a rapidly dissipating wake. Such a
rapid drop is attributed to a transverse-to-freestream
orientation of the vortex, rather than a streamwise orientation.
In the second configuration (Fig. 20), a small obstruction, a maximum of 34 mm, was alternately pro-

FISH

AND

SUVS

113

truded at the port and starboard sides of the nose to


generate small vortices and simulate the effect of the
head swaying of a fish. The vortices interacted with
those shed from the dual flapping foils in the tail. The
phase between the nose and tail flap motion was varied
and the axial force signature on the entire cylinder assembly was measured. The time-integrated values are
shown in Figure 25. It is remarkable that a fine thrust
regulation within 65% can be achieved by phased
seeding of vortices between spatially distributed control surfaces.
A biologically-inspired maneuvering vehicle. Measurements with the flapping foil cylinder model shown
in Figure 19 indicate the following. At say, 20 cm/sec
of flow speed, the steady drag levels are 1/100th of the
peak unsteady forces and 1/50th of the time mean values due to the dual flapping foils. Thus, a remarkable
feature of flapping foil locomotion is the production
of large unsteady forces. However, in man-made systems, unsteady mechanisms are rare.
While their peak and mean values are large, the
forces produced by flapping foils are inherently periodic with large differences between the minimum and
maximum values. To generate large forces practically
at all phase, pairs of flapping foils may therefore be
operated out of phase. Figure 26 shows such a small
vehicle. The dual flapping foils are arranged in a crucifix form. For pure thrust, the horizontal pairs operate
out of phase with respect to the vertical pairs. Independent operation of the dual flapping foils could provide precision maneuvering. The main body has a laminar low drag profile. The low-speed swimming of the
tethered neutrally buoyant vehicle in Figure 26 has
been demonstrated in a small tank (Bandyopadhyay et
al., 2000b).
BIOMIMETIC PROPULSOR: ACTIVE NOISE CONTROL
In a predator-prey environment, the ideal underwater animal or vehicle should produce no noise and
leave no wake signature. Thus the primary motivation
for biological studies might be stealth. Maneuvering

FIG. 26. A small agile vehicle. The ruler is 30 cm long. The four pairs of flapping foils are arranged in a cross configuration. The horizontal
and vertical pairs oscillate in the opposite phase producing high and uniform thrust. The main body is the so-called low drag laminar B1 body.
The flap actuators are located inside the main body.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 25. Variation of time-averaged axial thrust with phase lag of


nose slider. Speed, Frequency & Tail Flap Strouhal number St: 16.2
21.0 cm/sec, 3.77 Hz & 0.60.46.

OF

114

PROMODE R. BANDYOPADHYAY

Modeling of the effects of reduction in rotational speed of a propulsor on radiated noise.

involves operation at off-design condition. Two strategies for propulsor noise reduction are examined here:
reduction of rotational speed and blade tonals, where
science distilled from biology may be applicable to
engineering problems.
Reduction of rotational speed
A modeling was carried out to determine the effect
of reduction of rotational rate of a propulsor on these
sources of radiated noise: blade rate tonal due to wake
deficit; trailing edge singing; and, ingested turbulence.

They are expressed as rotational rate (RPM) to the


power of 4, 5 and 6, respectively. The result is shown
in Figure 27. A RPM reduction of 5% can give a 3
5 dB reduction in noise.
In principle, we can propose that the application of
unsteady lifting mechanisms of fruit fly (Ellington,
1984; Dickinson et al., 1999), namely rotation lift, delayed stall and wake capture, might be an avenue for
lift enhancement, leading to reduction of RPM. However, the implementation of these mechanisms to practical propulsors is fraught with difficulties. An important step in this direction involves the development of
programmable cambering of blades using electro-active polymeric artificial muscles. A small two-bladed
propulsor, shown in Figure 28, was built. A controller
was built that could power the muscle electrodes by
means of square waveforms: positive, negative or bipolar. The peak volts, frequency and duty cycle could
be varied. Figure 29 shows the types of blade cambering achieved. A low frequency pulse (O(1 Hz)) led
to cambering and oscillation, while a high frequency
pulse (O(100 Hz)) led to cambering only. Figure 30
shows an example of thrust enhancement of about 15%
at a RPM of 520 due to blade cambering.
Reduction of blade tonals

FIG. 28. Details of the Biomimetic Propulsor model: (a) end view
of blade with artificial muscle, (b) plan view of blade in (a), and (c)
experimental setup of propulsor assembly in water-filled box, with
motor drive and load cell.

A rotor blade traversing the wake of stator blade


experiences a time-dependent load due to vertical gust,
which gives rise to radiated noise, called blade tonals.
Lighthills equation relating the load derivative to fluctuating pressure describes this overall process. Now, if
the trailing edge of the upstream stator blade is oscillated like a fish tail, then the momentum deficit in the
wake can be filled. This can be expected to reduce the

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 27.

MANEUVERING HYDRODYNAMICS

OF

FISH

AND

SUVS

115

vertical gust on the rotor, and thereby the noise radiated. This flow, shown in Figure 31, was modeled hydrodynamically. The controller, of the dynamic inversion type, is shown in Figure 32, which produces the
circulation necessary to cancel the derivative of the

lift. Sears function is used to account for the phase


difference between gust and rotor lift. The result of a
controller canceling the derivative of the lift is shown
in Figure 33. The peak-peak amplitude of the radiated
noise is reduced by at least 40 dB.

FIG. 30. Comparison of thrust signature during active cambering. Two-bladed propulsor. RPM 5 520. Negative Fx indicates positive thrust.
Muscle: MS-417 cloth backing; Power: 3 V, 1 Hz; No bubbles are produced during these measurements. Pulse: Negative Unipolar producing
positive cambering.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 29. Cambering of blade muscle. Non-rotating propulsor. Input pulse forms and nature of cambering are as follows. (a) Bipolar: Muscle
is Flat, (b) Positive Unipolar: Muscle is Cambered Negatively; Negative Lift; Reversed Thrust, and (c) Negative Unipolar: Muscle is Cambered
Positively; Positive Lift; Forward Thrust. Power supply to muscles: 100 Hz and 7 V; High frequency leads to cambering without oscillation
and bubbles; 1 cm 3 1 cm Grid.

116

PROMODE R. BANDYOPADHYAY

CONCLUDING REMARKS
The progress made at NUWC in bridging biology
and the hydrodynamics of small underwater vehicles
is reviewed. Maneuvering hydrodynamics is examined, the goal being the understanding of mechanisms
of noise reduction. The approach is not to build robotic
replicas of swimming or flying animals. Instead, the
scientific principle is distilled and applications are conceived for retro-fitting to, or for modification of existing engineering vehicles or components.
The general conclusion is that, swimming and
flight in nature are characterized by oscillating control
surfaces, which could sometimes include the main
body. The mechanism of discrete and deterministic
vortex shedding from oscillating control surfaces has
the property of large amplitude unsteady forcing and
an exquisite phase dependence, which makes it suitable for active control for precision maneuvering,
stealth and lift enhancement. The knowledge base is
new but advanced. Several examples of applications
to rigid cylinders and propulsors are discussed. However, the applications are not widely explored yet.
Rich possibilities remain. Advances in actuator technology, like artificial muscles, are required for new
and matured application of the recently understood

FIG. 32. Block diagram of control strategy: choose actuator circulation G(t), so that L in rotor blade is reduced.

unsteady vortex dynamics principles of swimming


and flight in nature.
ACKNOWLEDGMENTS
The author gratefully acknowledges the sponsorship of the Office of Naval Research (Codes 342 and
333), Program Managers Dr. Teresa McMullen, and
Mr. James Fein, for early support. Collaboration
with the following are acknowledged: Professor Anuradha Annaswamy of MIT, William P. Krol, Jr.,
William H. Nedderman, John Castano, James Dick,
James Q. Rice and Daniel P. Thivierge of NUWC,
Dr. William Macy of URI, Professor Sahjendra
Singh of UNLV and Dr. Mehran Mojarrad of BPI of
New Mexico.

FIG. 33. Comparison of radiated noise due to gust from stator


wake. Upper: control off; lower: control on.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

FIG. 31. Schematic representation of blade tonal generation: a rotor blade traversing a stator wake. The trailing edge of the stator blade
oscillates like a fish tail to fill the wake momentum deficit and reduce the vertical gust on the rotor blade, which is the source of blade tonal
noise.

MANEUVERING HYDRODYNAMICS
REFERENCES

FISH

AND

SUVS

117

Dickinson, M. H., F. Lehmann, and S. P. Sane. 1999. Wing rotation


and the aerodynamic basis of insect flight. Science. 284:1,954
1,960.
Ellington, C. P. 1984. The aerodynamics of hovering insect flight.
IV. Aerodynamic mechanisms. Phil. Trans. R. Soc. London Ser
B 305:79113.
Fish, F. E. 1999. Performance constraints on the maneuverability of
flexible and rigid biological systems. Proceedings of the 11th
International Symposium on Unmanned Untethered Submersible Technology. Autonomous Undersea Systems Institute, Durham, New Hampshire: pp. 394406.
Fish, F. E. and J. Rohr. 1999. Review of dolphin hydrodynamics and
swimming performance. SPAWARS System Center Technical
Report 1801, San Diego, California.
Madden, J. D., P. G. Madden, and I. W. Hunter. 2001. In Proceedings
of SPIE 8th Annual Symposium on Smart Structures and Materials: Electroactive polymer actuators and devices. Yoseph
Bar-Cohen, Ed., SPIE, Bellingham, Washington.
Rohr, J. J., E. W. Hendricks, L. Quigley, F. E. Fish, J. W. Gilpatrick,
and J. Scardina-Ludwig. 1998. Observations of dolphin swimming speed and Strouhal number. Tech. Rept. 1,769, U.S. Navy
Space and Naval Warfare Systems Center, San Diego, California.
Lighthill, J. 1975. Mathematical biofluiddynamics. Soc. Ind. Appl.
Math., Philadelphia.
Taylor, G. I. 1952. Analysis of the swimming of long and narrow
animals. Proc. R. Soc. London. Ser. A Biol. Sci. 214:158183.
Triantafyllou, G. S. and M. S. Triantafyllou. 1995. An efficient
swimming machine. Sci. Amer. 272:6470.
Videler, J. J. 1993. Fish swimming. Chapman & Hall, London.
Webb, P. W. 1975. Hydrodynamics and energetics of fish propulsion.
Bull. Fish. Res. Bd. Can. 190:1158.
Wu, T. Y. 1971a. Hydrodynamics of swimming propulsion. Part 1.
Swimming of a two-dimensional flexible plate at variable forward speeds in an inviscid fluid. J. Fluid Mech. 46:337355.
Wu, T. Y. 1971b. Hydrodynamics of swimming propulsion. Part 2.
Some optimum shape problems. J. Fluid Mech. 46:521544.
Wu, T. Y. 1971c. Hydrodynamics of swimming propulsion. Part 3.
Swimming and optimum movements of a slender fish with side
fins. J. Fluid Mech. 46:545568.
Wu, T. Y. 1971d. Hydromechanics of swimming fishes and cetaceans. Adv. Appl. Mech. 11:163.
Wu, T. Y. 1972. Extraction of flow energy by a wing oscillating in
waves. J. Ship Res. 14:6678.

Downloaded from http://icb.oxfordjournals.org/ at University of the West of England on June 25, 2015

Aleyev, Y. G. 1977. Nekton. Junk, The Hague.


Annaswamy, A., W. P. Krol, Jr., and P. R. Bandyopadhyay. 2000. A
biomimetic propulsor for active noise control. Part 2: Theory.
Bull. of the American Physical Society, 45, No. 9:95.
Bandyopadhyay, P. R. 1996. A simplified momentum model of a
maneuvering device for small underwater vehicles. NUWCNPT Technical Report 10,552, Naval Undersea Warfare Center,
Newport, Rhode Island.
Bandyopadhyay, P. R. 1999. Emerging approaches to flow control
in hydrodynamics. Paper CDC99-INV0134, The 38th IEEE
Conf. On Decision and Control, Control Systems Society, Dec.
710, 1999, Tempe, AZ, Vol. 3 of 5:2,8452,850.
Bandyopadhyay, P. R., J. M. Castano, J. Q. Rice, R. B. Philips, W.
H. Nedderman, and W. K. Macy. 1997. Low-speed maneuvering
hydrodynamics of fish and small underwater vehicles. ASME
J. Fluids Eng. 119:136144.
Bandyopadhyay, P. R., J. M. Castano, W. H. Nedderman, and M. J.
Donnelly. 2000a. Experimental simulation of fish-inspired unsteady vortex dynamics on a rigid cylinder. ASME J. Fluids
Eng., 122, No. 2:219238.
Bandyopadhyay, P. R. et al., 2000b. Biomimetics research at
NUWC. Video (Edited), 14 Mts., Naval Undersea Warfare Center, Newport, Rhode Island.
Bandyopadhyay, P. R., W. H. Nedderman, and J. Dick. 1999a. Biologically-inspired bodies under surface waves. Part 1: Load
measurements. ASME J. Fluids Eng., 121, No. 2:469478.
Bandyopadhyay, P. R., S. Singh, and F. Chockalingam. 1999b. Biologically-inspired bodies under surface waves. Part 2: Theoretical control of maneuvering. ASME J. Fluids Eng., 121, No.
2:479487.
Bandyopadhyay, P. R., W. P. Krol, Jr., D. P. Thivierge, W. H. Nedderman, and M. Mojarrad. 2000c. A biomimetic propulsor for
active noise control. Part 1: Experiments. Bull. of the American
Physical Society, 45, No. 9:95.
Bannasch, R. 2001. New wing and propellor constructions inspired
by techniques learned from wing-propelled animals. Technical
University, Berlin (due to appear).
Bar-Cohen, Y. (ed.) 2001. Electroactive polymer actuators as artificial muscles: Reality, potential and challenges. SPIE Press,
Bellingham, Washington.
Belmonte, A., H. Eisenberg, and E. Moses. 1998. From flutter to
tumble: Inertial drag and Froude similarity in falling paper.
Physical Review Letters 81:345348.

OF

You might also like