You are on page 1of 135

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI
films the text directly from the original or copy submitted. Thus, some
thesis and dissertation copies are in typewriter face, while others may
be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the


copy submitted. Broken or indistinct print, colored or poor quality
illustrations and photographs, print bleedthrough, substandard margins,
and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete
manuscript and there are missing pages, these will be noted. Also, if
unauthorized copyright material had to be removed, a note will indicate
the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by


sectioning the original, beginning at the upper left-hand corner and
continuing from left to right in equal sections with small overlaps. Each
original is also photographed in one exposure and is included in
reduced form at the back of the book.

Photographs included in the original manuscript have been reproduced


xerographically in this copy. Higher quality 6" x 9" black and white
photographic prints are available for any photographs or illustrations
appearing in this copy for an additional charge. Contact UMI directly
to order.

University Microfilms International


A Bell & Howell Information C om p an y
3 0 0 North Z e e b R oad. Ann Arbor, Ml 4 8 1 0 6 -1 3 4 6 USA
3 1 3 /7 6 1 -4 7 0 0 8 0 0 /5 2 1 -0 6 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Order Num ber 9237863

Mechanics of flapping fin locomotion in the cownose ray,


Rhinoptera bonasus (Elasmobranchii: Myliobatidae)

Heine, Carlton E., Ph.D.


Duke University, 1992

UMI
300 N. ZeebRd.
Ann Arbor, MI 48106

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
M echanics o f Flapping Fin Locomotion in the Cownose Ray,
Rhinoptera bonasus (Elasmobranchii: M yliobatidae)

by
Carlton Heine
Department of Zoology
Duke University

Date: /Ip ril 3 ^ . ( ci c('2-


Approved:

Steven,Vogel, Supervis
isor
UUm Jj}
V Of _ ________________

Dissertation submitted in partial fulfillment of


the requirements for the degree of Doctor
of Philosophy in the Department of
Zoology in the Graduate School
of Duke University
1992

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
ABSTRACT
(Zoology)

M echanics o f Flapping Fin Locom otion in the Cownose Kay,


Rhinoptera bonasus (Elasmobranchii: M yliobatidae)

by
Carlton Heine
Department of Zoology
Duke University

Date: ^ f >n ^
Approved:

Steven Vogel# Supervisor

Wa*~-

An abstract, of a dissertation submitted in partial


fulfillment of the requirements for the degree
of Doctor of Philosophy in the Department of
Zoology in the Graduate School of
Duke University
1992

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
I j
fitch
P^-D.
HU2M
m z-
Abstract

Myliobatoid rays swim by flapping their large pectoral fins dorso-


ventrally. The motion of the fins is a predominantly flat heaving pattern
with the chord remaining nearly perpendicular to the fin's motion. This
type of ray swimming has been compared to bird flight, although no
quantitative studies have been published on the kinematics or mechanics
of swimming in these rays. In this thesis I document the kinematics,
and present a hypothesis for the mechanics of this large-amplitude,
flapping type of ray locomotion.
Two species, Rhinoptera bonasus and Myliobatis freminvillei,
were video-taped swimming in place against a current, and through still
water, over a range of swimming speeds from 0.3 to 0.9 m s‘l. From the
video tapes, the frequency, amplitude, and twist of the fin, and angle of
the body were measured over time. Pectoral fin-beat frequencies were
clustered around 1 Hz and were not correlated with swimming speed.
The amplitude of the fin-tip motion was between 0.6 and 0.9 times the
half wing span and also was not correlated with swimming speed. The
only variable strongly correlated with swimming speed was the
maximum speed of the wing tip on the upstroke. These kinematic data
were used to calculate the angle of attack of a flapping pectoral fin.
A ray was cast in rigid urethane foam, and this model was cut into
spanwise sections. The force on each section was measured in a wind
tunnel, and the lift and drag coefficients were calculated. The force
produced by a flapping ray wing was calculated from these coefficients
using the quasi-steady assumption. The calculated force shows no net

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
forward component; instead the resultant force points vertically,
oscillating up and down. These results suggest that thrust is not
produced by the flapping fins of these neutrally-buoyant anim als. My
hypothesis is that this imbalanced vertical force pushes on the broad,
pitching, airfoil-shaped body, causing the animal to glide forward.
Using unsteady wing theory the thrust produced by the oscillating
body was calculated to be approximately 1 N. Attempts to test the
pitching body hypothesis experimentally have proved supportive, yet not
conclusive.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Acknowledgements

A dissertation is, on the one hand, a scholarly project but it is also


a learning experience, and credit must be accorded to those individuals
who have guided this education. The members of my dissertation
committee, Drs. John Lundberg, Edward Shaughnessy, Vance Tucker,
Steven Vogel, and Stephen Wainwright, all played significant roles
throughout this project. Without their advice and guidence this
dissertation would have fallen short of its goals.
My comrades, graduate students Hugh Crenshaw, Olaf Ellers,
M att Healy, Ira Katz, John Long, Anne Moore, Ann Oliver, Ann Pabst,
Kathy Reinsel, Mark Westneat, and the other Zoology graduate students,
were always there to lend a hand, and to share my crazy ideas,
triumphs, and failures.
This dissertation not only faced intellectual hurdles but also many
practical and mechanical challenges. I need to thank the Harris
brothers, Rick Monaghan, and Joe Andrews for their essential role in
helping me acquire animals. The large flow tank would not exist
without the efforts of Julie Parrish, the advice and tools of the Marine
Lab's maintenance staff, and funds from Stephen Wainwright. Carl
Mills and Doug Karger were patient and helpful with my electronics
education. The energy and talents of Chuck Pell and the Biodesign
studio were indispensable in making the ray models.
Throughout all the ups and downs, providing unconditional
support and assistance, has been my family and my friend Dania.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Financial support was provided by the Zoology department and the
Cocos Foundation morphology fellowship.
Four individuals were so critical to my success that they deserve a
special mention; Dania for all your help, Hugh and Vance for all your
advice, and.most importantly my advisor, Steve Vogel, for your patience
with my frustrations and your light handed guidance.
A final note of gratitude must be awarded the rays who inspired
me, put up with my experiments, and tried to teach me how they swam.

cownose ray, Rhinoptera bonasus

Reproduced with permission of the copyright owner. Further reproduction prohibited w ithout permission.
Table o f Contents

C hapter 1. Introduction 1

C hapter 2. Study Animals 5


Morphology of cownose rays 7
Aspect ratio 7
Muscle morphology 10
Fin skeleton 11
Relative muscle mass 14
Density of rays 15

C hapter 3. Kinematics of Flapping Ray Locomotion 17


Materials and methods 17
Tanks 18
Video and computer digitization 20
Training the animals to swim 22
Analysis of the tapes 23
A lateral view of swimming 24
A posterior view of swimming 26
Results 26
Discussion 47

C h ap ter 4. Quasi-Steady Fluid Force Approximations 51


Materials and methods 54
Model ray 54
Wind tunnel 55

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Flight balance 56
Calibration of the balance 57
Experimental protocol 58
Angle of attack 00
Tunnel boundary corrections 61
Results 62
Force coefficients vs. angle of attack 62
Calculating thrust 71
Discussion 73
Forward motion from a vertical force 76

C hap ter 5. Unsteady Analysis of the Heaving,


Pitching, Ray Body 79
Theory 81
Calculation of thrust 85
Materials and methods 87
Results 90
Discussion 92

C h ap ter 6. Conclusions 96

A ppendix A. Flow Visualization 101

Appendix B. Angle Data 108

References. 113

Biography. 121

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
C hapter 1

Introduction

Rays are biologically successful animals using any of a number of


criteria for success. They are diverse, with over 400 species described
(Nelson, 1984). They have a long fossil record, dating back to the Upper
Jurassic (Compagno, 1977). They are abundant, vdth observations of
schools containing millions of individuals (Clark, 1963). This success, as
with any large aquatic vertebrate, requires rays to be strong swimmers.
Batoids as a group exhibit a wide range of swimming styles, from
the undulations of a shark-like caudal region, seen in the Pristidae, to
the almost bird-like flapping of the pectoral fins in the Mobulidae. Rays
that generate propulsion from their pectoral fins locomote in a unique
way. While the body is held rigid, the flexible pectoral fins contort to
oscillate dorso-ventrally, producing lift and thrust. Their swimming can
be described as graceful and powerful.
The patterns of pectoral fin movement in rays can be divided into
two broad categories, undulations of the fin's distal margin and large
amplitude flapping motions. Most species use only one of these types of
fin movements to swim; although, some rays that normally swim using
distal edge undulations can turn to large amplitude flapping as an
escape response. The division among the species, based on the two

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
patterns of fin movement is consistent with differences in the overall
shapes of the animals and aspects of their natural history.
Rays that swim by undulating the edges of their fins, such as the
Dasyatidae and Rajidae, are generally sedentary. They spend a great
deal of their time motionless on the bottom of the ocean, often part ia lly
buried in sand. Their seasonal migrations are unimpressive or entirely
absent. The profile of these animals varies from a circle to a diamond;
and, generally, the ratio of the wing span to the length of the body has a
value of one or less.
In contrast, rays that swim using the large amplitude flapping
motion of the pectoral fins, the Myliobatidae and the Mobulidae, have
very active lifestyles. Even though the Myliobatidae are tied to the ocean
bottom by their benthic diet, their time is spent swimming in the water
column, and they regularly undergo long migrations. Species in the
family Mobulidae, with their pelagic diet, are open-ocean fish and can be
found world-wide in tropical and sub-tropical waters (Bigelow, 1953).
The rhombic-shaped profile of these flapping rays is not as varied as
those of the rays using undulating motion. In these rays, the distal
margin of the pectoral fins end in distinct, pointed tips, and the ratio of
the wingspan to body length is greater than one. Within these two
families the shape of the pectoral fin is remarkably similar from one
species to the next.
Swimming mechanics and performance capabilities of rays have,
to date, been only minimally documented. A few observations were made
by Marey (1893) on the undulating type of ray swimming. The fluid

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
dynamics of the undulating type of ray locomotion were more recently
studied using a combination of unsteady airfoil theory and blade element
analysis (Daniel, 1987).
Even less information exists on the large amplitude flapping type
of locomotion. The studies that have been done all make superficial
comparisons to bird flight (Klausewitz, 1964; Oehmichen, 1958). These
studies are not quantitative but merely describe the motions. They do not
include an analysis of the mechanics involved. Although rays and birds
do oscillate pectoral appendages dorso-ventrally, they have very different
types of pectoral appendages. Birds that fly underwater, such as
penguins, have stiff, long and narrow wings and a rotating shoulder
joint. They rotate and swing their wings such that the major portion of
the stroke involves a non-accelerating motion which can be analyzed by
comparing the wings to a propeller (Hui, 1988). Rays have broad, flexible
fins and no joint that will allow an anterior-posterior rotation of the fins;
therefore the fin is constantly changing shape to produce the required
motion. The motion is dominated by a remarkably flat, heaving pattern,
with the chord nearly perpendicular to the fin's motion. The rays' stroke
involves very little twisting or pitching. These differences in both the
morphology and the details of the wing motion indicate that rays use a
different mechanism for producing the thrust required for locomotion.
The aim of this thesis is to analyze the mechanics of the large
amplitude flapping type of ray locomotion and to produce a model
describing how these rays produce thrust.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
This study focuses on the mechanics of locomotion of rays in the
family Myliobatidae. Chapter 2 introduces the study an im a ls and
includes a brief description of their general locomotor morphology. Also
included are measurements of their overall shape, locomotor muscle
mass, and density.
Chapter 3 is a quantitative study of myliobatoid swimming
kinematics. It includes measurements of beat frequency and amplitude
as a function of swimming speed, as well as fin and body angle changes.
Using this information the angle of attack is computed.
Analyzing the fluid mechanics of this type of locomotion is
extremely complicated. The assumptions required for conventional
approaches need to be extended, possibly beyond their limits, to be used
for an analysis of the large amplitude flapping type of ray locomotion.
An investigation of some of these assumptions is included as Appendix
A. The simplest approach for the force analysis uses a quasi-steady
approximation of the fluid forces based on empirically determined force
coefficients. This method is developed for flapping rays in Chapter 4.
A complete unsteady analysis is hampered by the requirement that
oscillations be small in amplitude, which is probably not true for the
pectoral fins. The body motion does meet this requirement and can be
analyzed using unsteady airfoil theory. These calculations and the
results from an oscillating device for measuring the forces from the
body's movements are included in Chapter 5.
Chapter 6 presents some concluding remarks from this study on
the large amplitude flapping type of ray locomotion.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
C hapter 2

Study Anim als

To collect data on the large amplitude flapping type of ray


locomotion, I have chosen to use two species in the family Myliobatidae,
the cownose ray, Rhinoptera bonasus (Mitchill), and to a lesser extent the
bullnose ray, Myliobatis freminvillei (Lesueur). They are common off the
Carolina coast, and are easy to transport to and work with at the Duke
Marine Lab. In the fall, many pound nets are set in Core Sound, fishing
for flounder. During the first two weeks of October there is typically a
large side catch of cownose rays. I established contact with local
fishermen who helped me acquire animals from their nets. All of the
rays I encountered were female.
Cownose rays survive well in captivity and are easy to train and to
work with. This species is one of the smallest in the family Myliobatidae.
They are bom at a wingspan of about 35 cm and reach about 100 cm when
full grown (Smith, 1987). It is easy to find individuals with wingspans of
45 cm, which are large enough to make observations on but small
enough to fit in the available tanks. Up to six rays were maintained
simultaneously, in either two 1500 litre tanks or in the 5600 litre raceway.
After much trial and error they were found to feed on the razor dam
Tagelus sp. They were allowed to satiate, with each animal eating

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
approximately one dozen 6 cm clams per day. The water supply in all
holding tanks used a flow-through system, and the temperature was not
controlled. The water temperature ranged from 15°C to 25°C. At the
lower temperatures the animals were sluggish and did not feed.
In addition, cownose rays are one of only a few species of batoids
that have received attention from scientists, so information is available
about their general biology (see Smith, 1985,1986,1987; Schwartz, 1965,
1967). They are thought to be a dominant predator on the commercially
valuable oyster population in the Chesapeake bay, and this has motivated
several studies of their life history. Tagging of cownose rays released off
the Carolina coast has shown that they migrate as far south as northern
South America for the winter months and north to the Chesapeake bay
for the summer (Schwartz 1965). Observations done in Chesapeake bay
have indicated, using point to point times, that they swim at speeds
between .2 and .7 m s 'l when feeding (Ben Blaylock personal comm). I
have observed cownose rays in the wild swimming at speeds of greater
than 1 m s"l. While more data is needed to determine what speeds and
ranges are possible for this species of ray, they seem to be excellent
swimmers, capable of high speeds for long distances.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Morphology of cownose rays

All rays have the common characteristic of being dorso-ventrally


flattened. This feature enables the main aspects of their overall shape to
be described using the dorsal projection. Both cownose and bullnose rays
have a generally rhombic profile with their wingtips ending in distinct,
swept-back points, typical of the flapping rays; see figure 2.1.
Prom a lateral perspective these rays are thin and streamlined
with the trailing edge being relatively sharp. The fineness ratio, defined
as the chord divided by the maximum thickness, ranges from 5 to 10 for
myliobatoid rays. From an anterior view these animals are deepest at
the center of the body with the thickness smoothly tapering off toward the
wing tips. The overall morphology of rays that employ the large
amplitude flapping type of locomotion, the Myliobatidae and the
Mobulidae, is remarkable consistent from one species to the next.

Aspect ratio
The most common measure of the projected wing shape used in
fluid mechanics is the aspect ratio, defined as the wing span squared,
divided by the wing area (Mises, 1945). The wing span is the distance
between the wing tips, with the wing tip defined as the point on the wing
most distal to the axis of the body. Defining the wing area is not simple
for rays, since one of their characteristics is that the anterior edge of the
pectoral fin is continuous with the head. It is difficult to define where the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 2.1 Cownose ray, Rhinoptera bonasus

V-

Bullnose ray, Myliobatis freminvillei

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
wing stops and the body begins. The convention in aerodynamics is that
the area of the body transected by the wings is included in the wing area.
In trying to keep within this convention while using a tractable
definition, I have defined the wing area as the area of the dorsal
projection of the entire animal minus the pelvic fins and tail. This
definition can be applied to all species of rays, allowing comparisons to be
made.
To measure aspect ratios, freshly killed animals were laid on a
piece of posterboard and the margins of the animals were traced, not
including the pelvic fins and tail. The tracings were cut out and passed
though a LI-COR model 3100 area meter. The area meter was calibrated
using a calibrated disc indicating a systematic error of less than 0.1% of
the reported values. The tracings were run through the meter 3 times
and the average area was recorded. The imprecision, defined as the
standard deviation of repeated measurements expressed as a percentage
of the average value, was less than 0.25%. One individual was traced
three times and three separate cut outs produced. The variation of these
three tracings was less than 1%. In addition the cast of a cownose ray,
used in Chapter 4, was also measured.
The wing span was measured on the animal using a meter stick.
The span measurement was repeated 3 times and the average taken.
The measurement variation in span was less than 0.25%. The aspect
ratios of standard shapes were calculated from their equations and are
also included in Table 2.1. The results are presented in the following
table with an accuracy of better than ± 0.05 of the reported values.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Table 2.1 Aspect Ratios

Aspect Ratio

circle 1.3
diamond (square) 2.0
Dasyatis sabina 1.3
Raja eglanteri 1.8
Myliobatis freminvillei
individual #1 3.2
#2 3.5
#3 3.2
Rhinoptera bonasus
individual #1 3.3
#2 3.5
#3 3.4
#4 3.5
#5 3.3
#6 3.4
cownose cast 3.3

Muscle morphology
I dissected four individuals, one bullnose and three cownose rays,
for information on their internal structure. These dissections were not
meant to produce a detailed description of the anatomy; rather they were
meant to provide a general picture of the motor propelling the animals.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
These dissections revealed many odd structures which warrant further
attention, but which are beyond the scope of this dissertation.
With the skin removed the muscles in the pectoral fins were
arranged in parallel bundles running from the body out to the wing tip.
The muscles looked like neatly arranged strands of spaghetti with some
bundles wrapped like candy canes with red muscle fibers. This pattern
was observed on both the dorsal and ventral sides of the fins. The skin
was well attached to the muscle and was inseparable for the most distal
10% of the fin. The anatomy of the fins of these rays is consistent with
the general features described by Calvert (1983) and Bone (1989). The
detailed structure of the muscle arrangement has not been investigated
in the species used in this study.

Fin skeleton
Occupying the plane in the middle of the fin is the array of
cartilaginous radials which, with the three pterygia at the base of the fin,
make up the fin skeleton; see figure 2.2. The radials run perpendicular
to the anterior-posterior axis of the body, and are jointed spanwise. The
radial segments range in length from 0.5 to 1.3 cm in animals with a
wingspan of approximately 50 cm. There is approximately a one to one
correlation between fin ray and muscle bundle.
At the base of the fin there are three basal elements. They are,
anteriorly to posteriorly: the propterygium, the mesopterygium, and the
metapterygium. The propterygium is fused with the cranium and

11

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
articulates with the scapula, coracoid, and the mesopterygium at its
posterior end. The metapterygium lies between the viscera and the fin
and articulates anteriorly with the scapula and mesopterygium. These
joints do not permit large ranges of motion.
The pattern of joints between the radials is highly organized in the
fin. If it is assumed that bending occurs at the joints between the radials
and not in the radials themselves then this complex pattern will have
implications in the orientation of the flexibility of the whole pectoral fin.
The radials are longest near the fin tip and at the trailing edge. In the
center of the fin the radials alternate long and short, and the pattern of
joints in successive fin rays form a sweeping curve with the joints
generally lying farther from the body toward the leading edge of the fin;
see figure 2.2.

12

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
05

Figure 2.2 is a diagram of a dorsal view of the right fin skeleton with all
of the soft tissues removed. The leading edge is to the left.

13

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Relative muscle masses
With the pectoral fin structure being roughly symmetrical about
the fin rays, it is reasonable to ask about the symmetry of the thrust
production of the upstroke versus the downstroke. A crude answer to
this question can be provided by comparing the mass of the dorsal
musculature to that of the ventral musculature. If the muscles are of the
same type and their arrangement similar (as suggested by Calvert, 1983),
then the ratio of the power produced should be similar to the ratio of the
muscle masses.
To measure the ratio of muscle masses, I filleted the right and left,
dorsal and ventral fin muscles (for 4 separate fillets per animal). The
fillets included the skin and all material above or below the plane of the
fin rays. The three pterygial elements formed the proximal border, and
as much tissue as possible was included from the distal ends of the fins.
The weights were recorded using a triple beam balance to the nearest
gram, and the ratios were calculated for each fin. The percent of total
mass is the sum of the right and left dorsal and ventral muscle masses
divided by the total mass of the animal.

14

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Table 2J2 Relative Muscle Mass

Ventral / Dorsal
Right______ Left Total mass %Total mass

cownose #1 .73 .72 1.946 kg 28%


#2 .62 .64 2.102 30.
#3 .61 .69 1.412 24
#4 .62 .65 1.482 23
#5 .79 .76 1.318 26
bullnose #1 .53 .53 .845 21

Prom the results shown in table 2.2 it is reasonable to hypothesize


that the upstroke plays a more important role in producing thrust than
the downstroke. These results emphasize the importance of locomotion
for these rays, since such a significant portion of the animal is locomotor
muscle. There may also be a tendency for larger animals to have a
higher percentage of their body mass devoted to locomotor muscle
although more data is required to test this hypothesis.

Density of rays

It is curious that the upstroke would be the dominant power stroke


since rays, when motionless, sit on the bottom of the tank and are slightly
negatively buoyant. Rays have no swim bladder, and to my knowledge

15

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
have no way of rapidly adjusting their density. I measured the density of
three individual Rhinoptera bonasus. I weighed, to the nearest gram,
live, anesthetized, motionless, an im a ls both fully submerged and
emersed with excess water removed. I then weighed a measured volume
of water to determine its density. I calculated the density of the an im als
using the following relationship:

The density of the rays were 1062,1058, and 1052 kg m‘3, with an
accuracy estimated to be better than ± 5 kg m*3. The density of the
seawater was 1018 kg m“3.
While it is clear that a detailed examination of the morphology of
these animals would be a fruitful study, I have chosen to concentrate on
the interaction between the ray and the surrounding water instead. The
information presented here is restricted to that required for a better
understanding of the hydrodynamics. I do hope that in the future
someone studies the internal functional morphology of these animals.

16

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
C hapter 3

K inem atics o f F lapping


R ay L ocom otion

The foundation for studying the mechanics of flapping ray


locomotion is a quantitative analysis of the kinematics. Kinematics is the
study of motion without including aspects of mass or force. There have
been, to date, no studies of the kinematics of the Myliobatidae and only a
single qualitative study of the Mobulidae kinematics (Klausewitz, 1963).
Although the two families of rays exhibit similar general patterns of
movement, the details appear to be different. This chapter presents the
first quantitative analysis of the kinematics of the large amplitude
flapping type of ray locomotion. This study is restricted to non­
accelerating cruising locomotion of two species in the family
Myliobatidae.

M aterials and m ethods

To measure the kinematics of flapping ray locomotion I analyzed


the movements of cownose rays, Rhinoptera bonasus, and bullnose rays,
Myliobatis freminvillei, swimming in two tanks at the Duke Marine Lab.

17

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
In one tank the animals swam in still water past a window, and in the
other tank they swam in place against a current. Animals were filmed
from a lateral view, a dorsal view and, using an underwater housing,
from behind. The animals' movements were recorded on videotape and
the tapes later analyzed for the relevant data. All of the rays were
females collected from pound nets in Core Sound.

Tanks
One tank is a large still-water tank referred to as the raceway. The
other tank is a recirculating flow tank (see Vogel, 1981). The raceway is
composed of two 1.83 m. square sections connected by a 6.10 m. by 0.91 m.
straight section forming a two dimensional dumb-bell shape. The t a n k is
0.61 m. deep throughout and holds approximately 5700 liters. It has two
windows on one of the long walls in the straight section. The top is
completely open. Wallpaper with a 3 cm square grid was applied to the
wall opposite the windows and to the floor of the straight section. Water
is supplied by a flow-through sea water system at a rate of approximately
375 liters per hour. The water in the tank can be filtered using a
recirculating diatomaceous earth pool filter. All filming was done with
the water supply and filtration systems turned off. Video was recorded
through the window in the middle of the long straight section or through
a floating Plexiglass lid from above the tank.
The flow tank has a rectangular glass and PVC test section 2.29 m.
long by 0.61 m. wide by 0.61 m. deep. The return section is a C-shaped

18

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
construction of PVC pipe 0.53 m. in diameter. The total volume of the
tank is approximately 2800 liters, with approximately 850 liters in the test
section. The test section was bordered at the front and rear by ceiling
light panels made of a regular array of approximately 1 cm square
channels. A single panel was used in the rear; and two stacks, each
approximately 5 cm thick, were used as a flow straightening honeycomb
at the front. These grates prevented the animals from entering the
return section. The water is circulated by a propeller, located in the
return section, driven by a 2 HP variable speed, DC universal motor.
This flow tank is capable of generating flow velocities, in the test
section, of up to 0.9 m s'*. The flow speeds were measured using a
Marsh-McBimey model 511 electromagnetic flow probe calibrated
against a dye bolus traveling a given distance in a measured time. The
flow meter was used to map the water speeds, on a three-dimensional 150
point grid, in the working section. The water speed varied by less than
4% in the direction of the flow, and by less than 12% with depth, and less
than 17% across the test section. 70% of the variation in speed occurred
within 10 cm of a surface.
The flow speed was then measured at the middle of the cross
section, 10 cm behind the grid at the front of the test section, and these
results calibrated against the dial on the control box. Each control box
setting was tested in increasing and then decreasing increments for a
total of 6 independent times, and the results were averaged. The range
for the 6 trials was less than 4%. These measurements were checked by
digitizing the path of a dye bolus traveling 70% of the length of the test

19

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
section at each of the controller box settings. Linear regressions were
calculated for the horizontal motion of the dye stream versus time, with
the speed of the flow defined as the slope of the regression. All of the
regressions for water speed had correlation coefficients of at least .999.
During filming the control box settings were used to set the water speed,
and the averaged value at that setting was recorded as the water speed.
In the analysis only data from animals swimming within 10 cm of the
point where the speed was measured were used, and thus the variation
in speed over the area represented by the animal was less than 10%.
Using the patterns of dye movements, other characteristics of the
flow were calculated. The dye was released from a glass pipet and,
almost immediately after entering the flow, the dye stream spread to
have a cross section of approximately 1 cm2; approximately the size of the
individual channels in the honeycomb. The diffused dye streak
remained intact for the first 90% of the test section. At the back of the
working section the turbulence generated by the transition of square test
section to round return section spread the dye streak beyond recognition.
There was no detectable vertical motion of the dye stream except at the
highest speed where the dye stream rose 1 cm in 60 cm, forming an
angle of 1° with the floor of the tank.

Video and computer digitization


The animals were videotaped swimming using either a Magnavox
S-VHS CHQ model VR9260AV01, or a Canon VHS model VC-30. AH

20

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
filming was done using available daylight. Films in the flow tank were
shot from a distance of 5 m and films in the raceway at 3 m. At the
beginning of each filming session, a square grid of known dimensions
was filmed and later digitized for calibration purposes. Initially the grid
was digitized over the entire screen and the results used to calculate the
optical distortion. The optical distortion for the entire system, from the
object, through the glass of the tank, through the camera lens, onto the
video tape, through the tape player, and back out to the image on the
computer screen, was less than 0.5% in the horizontal and 0.9% in the
vertical direction. The maximum parallax error in the flow tank from
the back wall to the front wall was 5%. All measurements in the flow
tank were taken assuming the an im al was in the center of the tank. In
the raceway the maximum parallax error was 14% and all distance
measurements were corrected by the ratio of the length of the body as
measured on the animal to the calculated length from the video.
The videotape was played back using a Panasonic AG-6300 video
editing deck, which was capable of playing the tape frame by frame. All
the filming and analysis was done at the standard video frequency of 30
frames per second. Video technology requires a time signal with an
accuracy of better than .001 sec for the signal to be synchronized. The
frame count was therefore used to calculate time for all analyses and
assumed to introduce no detectable error in the measurements.
The computer digitization was done using an Amiga 1000 and a
Commodore 1300 genlock. The genlock allows the graphics of a
computer program to be superimposed over the video signal. Software

21

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
written in AmigaBasic placed a mouse-positioned cursor which was
used to select individual pixel positions, on the screen, and recorded the
screen coordinates of the pixels. The screen was a Commodore 1080 RGB
monitor with a resolution of 640 pixels wide by 200 pixels high. Pixel
positions were converted to real distances using the vertical and
horizontal calibration constants. These constants were calculated using
the number of pixels between a pair of points digitized from the square
grid, divided by the real distance between the grid points. To test for the
precision of the digitizing technique a single grid point was digitized for
90 frames and the pixel coordinates were identical for 88 of the frames,
the other 2 being off by a single pixel. It was therefore assumed that the
signal was stable and repeatable and introduced no detectable error in
the measurements.

Training the animals to swim


In the raceway, rays were recorded when they occasionally
volunteered to swim past the window, and no training was required.
Training them to swim in a closed box with current required more effort.
A nim als were introduced to the flow tank after at least a full day to
adjust to captivity. Training consisted of putting the animal in the tank,
and gradually turning up the current speed while the animal was facing
into the flow. When the animal turned to swim with the current the
motor was turned off, and the animal was assisted off the back grid. The
animals were coaxed by my hands to face into the flow and generally

22

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
swam when in current. The animals occasionally came to rest on the
bottom but were inspired to swim when they slid back too far. This
process required several days of intermittent effort before the rays would
swim in place for extended periods. Not all of the animal a responded
successfully to the training. Once the animals were trained they swam
mostly in the front of the tank, approximately 5-10 cm from the
honeycomb. The camera was therefore positioned to cover the front half
of the test section.

Analysis of the tapes


Since this study focuses on normal, non-accelerating locomotion,
the following criteria were used to select footage for analysis. In the data
from the raceway the horizontal position of the nose versus time was
required to have a linear relationship with constant slope and a
correlation coefficient of better than .90, and the vertical position of the
nose oscillated about a value that changed by less than 2 cm. The
duration of segments used for analysis lasted as long as the above
criteria held or until the animal passed from view through the window.
Segments from the flow tank lasted a full 3 sec. or 90 frames. In the flow
tank the same criterion was used for the vertical position of the nose, and
the mean horizontal position of the nose was also required to change by
less than 2 cm. Only footage where the animal was more than 10 cm
from a surface was used, and furthermore the animal had to be within
10 cm of the point where the flow speeds were recorded. The an im als did

23

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
not roll or yaw by a visually detectable amount. When filming from
behind, the position of the nose was estimated. This set of criteria form
my definition of steady, non-accelerating swimming.

A lateral view of swimming


Once a section of tape had been selected for analysis, three points
and three slopes were digitized in each frame of the sequence. The three
points were: (a) the most anterior point on the rostrum, (b) the wing tip of
the near wing, and (c) the most posterior point of the dorsal fin. This last
point is the most clearly definable point near the back margin of the body.
The profile of the posterior 2/3 of the ray's body forms a wedge with
approximately straight slopes on both the dorsal and ventral surfaces.
The slopes were measured for the profile of the (d) dorsal and (e) ventral
surfaces of the body, and (f) for the wing at a position 60% of the distance
from the center line of the body toward the near wing tip (see figure 3.0).
(A line had been drawn at this point on the ventral surface of the wing,
parallel to the long axis of the body, using a laundry marker. This line
made identifying this angle unambiguous.) Each slope was recorded as
the difference in vertical pixel position divided by the difference in
horizontal pixel position. These values were later multiplied by the ratio
of the vertical to horizontal calibration coefficients to yield the corrected
slope.

24

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.0
b

The data describing the position of the wing tip was examined and
the maximum and minimum values of the vertical component and the
corresponding frame count of those values were recorded. From these
data the wing beat frequency and amplitude could be calculated. The
vertical distance between sequential maximums and minimums
averaged over a 3 sec. sequence was recorded as the amplitude of the
wing motion. The frequency was calculated as the reciprocal of the
duration of each full stoke cycle, (maximum to maximum or minimum
to minimum), then averaged over the 3 sec. sequence.
Identification of the same point on the animal in each frame
provided the largest source of error in this digitization technique. This
error was measured by digitizing the exact same segment of tape three
independent times then calculating the standard deviation at each of the
three points in each frame. The average of these standard deviations
indicated the precision of the digitization. The precision for all three
points was ± 2 mm vertically and ± 2 mm horizontally. The slopes had a
precision of ± 1.1° for the body surfaces and ± 3.1° for the wing angle.

25

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A posterior view of swimming
To measure the spanwise curvature and the twist of the wing the
animals were filmed from behind in the flow tank. The distance from
the center of the body to the right wing tip was divided into 5 equal
sections and the points at the leading and trailing edges, on the proximal
side of each section, were digitized. La the two-dimensional image, from
the posterior view, the chord is projected onto the vertical plane and this
projected length was measured at all five locations. The distances
between these points were measured on the animal, and the arctangent
of the ratio of the projected length to the actual chord was calculated.
The precision of these angle measurements is: 2.3° for the center of the
body, 1.7° for the outer edge of the body, 2.6° for the proximal wing
section, 4.6° for the central wing section, and 10.9° for the distal wing
section. The results were transformed onto a coordinate system moving
with the body (the body angle is fixed at zero degrees). The changes in
these angles over time and over the span of the wing gave a dynamic
measure of wing twisting.

Results

Two types of graphs are used to display the results. The first type
(figures 3.1,3.2, 3.3, 3.6,3.9,3.12,3.15, 3.16, 3.17, 3.18) are line plots of
different variables against time. These graphs illustrate the typical
pattern present in the data. The important information is in the shape of

26

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
the curves and in the phase relationships with the wing tip motion. All
data presented in these graphs comes from an a single cownose ray, with
a wing span of 46 cm, swimming in the flow tank at a speed of 0.6 m s"l.
This sequence was chosen because the ray's movements were most
uniform, and thus it presented results with maximal clarity. The errors
from digitizing were smoothed by averaging three independent
digitizations. The coordinate system for all of these types of graphs is
centered in the body with the X axis along the wing span, the Z axis
aligned with the long axis of the body, and the Y axis projecting dorsally.
The animal is swimming in the positive Z direction.
The second type of graph (figures 3.4,3.5,3.7, 3.8, 3.10,3.11,3.13,
3.14) presents summaries of the data. The variables are plotted over the
range of swimming speeds investigated. Each data point represents the
measured variable averaged over the duration of a single three second
filming sequence. The data for these graphs were taken from, a cownose
ray in the still-water raceway, a single bullnose ray in the flow tank, and
four different cownose rays swimming in the flow tank. Only one
animal, a cownose ray swimming in the flow tank, cooperated
adequately enough to produce enough data for statistical analysis. All of
the regression results come from this individual, cownose A, with the
other cases plotted using different symbols for comparison by inspection.

27

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.1 comes from a posterior view, looking along the long a-sHg
of the body from behind, and plots the horizontal (X) and vertical (Y)
position of the wing tip over time.

Figure 3.1

Wing Tip Position


(posterior view)

O 20 ■

X(t)
Y(t)

10 -

0 1 2 3
Time (sec)

The vertical motion of the wing tip appears as a smooth sinusoidal wave
while the horizontal position shows the wing tip moving toward the
midline of the body at the top of the stroke. Viewed from this angle, the
wing is bent into a smooth arc at the top of the stroke, and is out straight
to the side level with the body at the bottom of the stroke.

28

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.2 shows, from a lateral perspective, the Y-Z plane, the
horizontal position (Z) of the nose and wing tip over time for the animal
swimming in place against a current.

Figure 3.2
Horizontal Displacem ent
(lateral veiw)

Nose Z(t)

Wingtip Z(t)

5- i

0 1 2 3
Time (sec)

The motion of the nose is very uniform with no detectable oscillations in


its position over time, or forward velocity. The wing tip undergoes only
small motions parallel to the direction that the animal is swimming in
and does not seem to move in a repeatable pattern in this direction.

29

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.3 shows the vertical displacement of the wing tip, nose,
and tail over time again from a lateral perspective.

Figure 3.3

V ertical Displacment
(lateral view)

— Nose Y(t)
-•— Wing Tip Y(t)
■o— Tail Y(t)

0 1 2 3
Time (sec)

Notice that the fin is not brought below the plane of the body on the
downstroke. Also note that the nose oscillates 180° out of phase with the
wing tip and out of phase with the tail. The odd shape of the top of the
wing tip stroke reflects the wing curling with the tip moving laterally
away from the cameras in this view.

30

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.4, the first of the summary graphs, shows the frequency
of the wing beat plotted over the range of swimming speeds recorded.

Figure 3.4
Stroke Frequency

□ Bullnose
♦ Stillwater
■ Cownose A
1.0
• Cownose B
■ Cownose C
□ Cownose D

Speed (m/s)
y = 0.65828 + 0.60105x RA2 = 0.378

The range of frequencies exhibited is narrow, from .8 to 1.4 Hz, and the
correlation with swimming speed is not strong.

31

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The next figure, 3.5, presents the amplitude of the wing beat
versus the swimming speed. The amplitude is expressed as the
dimensionless ratio of the vertical displacement of the wing tip to the h a lf

wing span.

Figure 3.5
Stroke Amplitude

Q Bullnose
♦ Stillwater
■ Cownose A
* Cownose B
■ Cownose C
° Cownose D

”»----- 1------1----- 1----- •— r


0.3 0.4 0.5 0.6

Speed (m/s)
y = 0.45645 + 0.11145x RA2 = 0.071

Each individual ray seems to have a preferred amplitude which it holds


relatively constant over this range of swimming speeds. There may be
individual differences in the magnitude of this preferred amplitude.

32

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.6 shows the angle between the dorsal surface of the body
and the horizontal plane over time, measured from the lateral view.

Figure 3.6
Body Angle

10 - .
bo
3« *• VA
I*
w
fa f1a< OL
i hi
i1
1a a fa
4l a1
* 1.
A I • A -5
w
f i *n i L
A

o-
%Vt
a i a '
• • I I
fL va a
b• a
*
P
r
* kW.m
F
/9 *a !

Xr
*a• **a
» a at
b:
•.*
Body Angle

Tip Position

The body pitches almost 180° out of phase with the wing tip motion.

33

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Taking the average maximum and minimum values of the
pitching angles of the dorsal surface of the body and plotting them
against the range of swimming speeds yields figures 3.7 and 3.8.
Figure 3.7 shows the maximum angles of the body, or the angle of
the body pitched maximally nose up.

Figure 3.7
Maximum Body Angle
20

Q
bn
£3 ■
s Bullnose
* m B i S B B
10 ' -------------- ------- *
1 ° ■ Cownose A
■ o 9 • 0 d Cownose C
IPQ ® a
■ Cownose D
a

0.3 0.4 0.5 0.6 0.7 0.8 0.9

Speed (m/s)
y = 11.667 - 2.0725x RA2 = 0.036

34

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.8 shows the minimum angle of the body, or the angle
when the body is maximally pitched nose down.

Figure 3.8

Minimum Body Angle

4-

0 Bullnose
• Stillwater
Cownose A
• Cownose C
■ Cownose D
- 6-

- 8-

-10
0.3 0.4 0.5 0.6 0.7 0.8 0.9
Speed (m/s)
y = 1.0364 - 2.4456x RA2 = 0.070

There appears to be no systematic variation in the magnitude of the


body's pitching over this range of swimming speeds.

35

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.9 shows the angle between the wing and the horizontal
plane, at the measured position, over time.

Figure 3.9
W ing Angle

Wing Angle
Tip Position

1 2
Time (sec)

The wing twists up at the beginning of the upstroke, then down slightly
at the beginning of the downstroke. The pattern of wing twist is dearly
not sinusoidal with the duration of positive twist, leading edge up, only
25% of the period of the cyde. The flapping, or heaving of the wing, does
show a phase lag to the pitching of the wing.

36

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The average maximum and m inim um angles of the wing's
twisting plotted against swimming speed yields figures 3.10 and 3.11.
Figure 3.10 shows the maximum wing angles or the angle when
the fin is twisted maximally leading edge up.

F igure 3.10
Maximum W ing A ngle

13 Bullnose
• Stillwater
■ Cownose A
* Cownose C
■ Cownose D

0.3 0.4 0.5 0.6 0.7 0.8 0.9


Speed (m/s)

y = 16.317+ 8.3618* RA2 = 0.134

37

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.11 shows the minimum wing angles or the angle when
the wing is maximally twisted leading edge down.

F igure 3.11
M inimum W ing Angle

0 Bullnose
3
♦ Stillwater
33
M -10 - B Cownose A
• Cownose C
Cownose D

-20
0.3 0.4 0.5 0.6 0.7 0.8 0.9

Speed (m/s)
y = -4.4709-8.8477x RA2 = 0.424

The wing may increase the angle through which it twists with
increasing swimming speed, but the change is only a few degrees and it
is not strongly correlated with swimming speed.

38

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.12 shows the speed of the wing tip and was generated by
taking the derivative with respect to time of the wing tip position. At each
point in time the vertical displacement between the preceding point and
the following point was divided by the elapsed time and that value
recorded as the speed of the wing tip at that point. The upstroke speed is
recorded as positive and the downstroke negative.

Figure 3.12
Wing Tip Speed

>*
I
%
£
cn -0.2" it
“ Tip Speed

*' Tip Position

Time (sec)

Note that the speed of the wing tip changes constantly over a stroke cycle
resulting in large accelerations.

39

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.13 shows the plot of the average m axim um upstroke
speed against the swimming speeds.

Figure 3.13
Upstroke Tip Speed
0.9

0.8 -

! 0.7 -
3 Bullnose
• Stillwater
0.6 - ■ Cownose A
° Cownose B
Cownose C
n Cownose D
% 0 .4 -

0.3
0.3 0 .4 0.5 0.6 0 .7 0.8 0 .9
Speed (m/s)
y = 0.25613 + 0.58762x RA2 = 0.904

The average maximum speed of the wing tip, on the upstroke, has a
strong correlation with swimming speed. The slope of this relationship
was statistically different from 0 at the .001 level.

40

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.14 plots the average maximum downstroke tip speed
against the swimming speed.

Figure 3.14
Downstroke U p Speed
0.7

0.6 -

0 Bullnose
0.5 - ♦ Stillwater
i> ■ Cownose A
° Cownose B
0 .4 -
■ Cownose C
a 0 Cownose D
0.3 -

0.2
0.3 0.4 0.5 0.6 0.7 0.8 0.9

Speed (m/s)
y = 0.22152 + 0.21387x RA2 = 0.234

The strong correlation between tip speed and swimming speed is not
repeated for the downstroke.

41

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 3.15 shows the profile of the right wing, as seen from a
posterior view, X-Y plane, during a stroke cycle. Each line represents
the profile of the fin at a different time during the upstroke.

Figure 3.15

Wing Curvature

2/30 sec
10 -
4/30 sec
I
6/30 sec

5- 8/30 sec

10/30 sec

0 10 20 30

X(cm )

The vertical and horizontal excursions of each position on the wing can
be described as a percentage of the wing tip motion. These percentages
were calculated using constant arc-lengths and used to translate the tip
motion into descriptions for the motion of the entire wing.
With the flapping velocity known at several positions along the
wing and the forward velocity of the whole animal known, the relative
fluid angle can be calculated at those positions. The relative fluid angle
describes the direction from which a local element of the wing

42

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
encounters the fluid. It is calculated by vector addition of the forward
velocity and the vertical and horizontal components of the flapping
velocity.
Figure 3.16 plots the angle between the direction of the animal’s
propagation and the direction from which the midspan of the four wing
sections encounter the oncoming fluid.

Figure 3.16
Relative Fluid Angle

40-

toA Distal
20- Central
3
Proximal
Base
01

-20 -

-40
0 1 2 3
Time (sec)

The amplitude of the relative fluid angle increases with each more distal
section. Even out toward the wing tip the relative fluid angle is always

43

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
less than 45°. This is due to the slow flapping speeds and the fast
forward speed of these swimming rays.
The change in angle of the fin along its span describes the twist
pattern of that fin. These angles were computed from the chord lengths
projected onto the vertical plane, X-Y, in the posterior view.
Figure 3.17 presents the calculated values of the angle at four
points along the span of the fin. The angles are expressed relative to the
angle of the body measured with the same method.

Figure 3.17
W ing T w ist

48 20

1 2
Time (sec)

The third position is the same location that was measured as the wing
angle from the lateral view, shown in figure 3.9. There is agreement

44

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
between the results for the magnitude of twist from the two independent
methods of measuring the angle of the wing. From these data the
average percentage of twist of each section was calculated with respect to
the third section with the following results: body, 12%; base of the wing,
16%; proximal wing section, 34%; central wing section, 100%; distal wing
section, 132%.
With the data on the relative fluid angle and a description of the
twist of the wing the angle of attack can be calculated. The angle of
attack is defined as the angle between the chord of the wing and the
relative fluid angle. The data on average twist of each wing section was
applied to the data on wing angle over time, shown in figure 3.9, and the
angle formed between this result and the relative fluid angle for each
section was computed over time.

45

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The angle of attack is plotted in figure 3.18 for the 4 wing sections.
Again the amplitude increases with each more distal section, with the
maximum value found at the most distal position.

Figure 3.18
Angle of Attack
25 -

15 -

5-
Distal
‘So
0) •
Central
« -5 “ Proximal
. Base
I -15 -

-25 -

-35 -

Time (sec)

With the exception of the most distal section, where the angles of attack
becomes quite large, the angles calculated fall between +15° and -15°.
This range of angles of attack is reasonable for a lift producing wing.
The angles of attack have not been corrected for the effect of the
downwash in the wake altering the local flow pattern.

46

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
D iscussion

Watching rays swim leaves the impression that they are


expending little effort to move forward at a significant speed. This
impression is due to the low frequencies and speeds with which they flap
their pectoral fins. Other animals move their locomotor appendages
much faster relative to their forward speeds. A classic series of results
from fish swimming is that most teleosts beat their tails from around 1
Hz to about 20 Hz, and that the tail beat frequency is strongly correlated
with swimming speed (Bainbridge, 1958; Hunter and Zweifel, 1971).
Rays oscillate their wings at about 1 Hz, and variation is not correlated
with swimming speed.
Rays do not vary the frequency of their stroke cycle to swim faster.
Neither do they vary the amplitude of the stroke, or the angle of the
wing's twist, or angle of the body's pitch in a way that correlates with
swimming speed. A multiple regression using all of these parameters
also fails to show a correlation with swimming speed. Instead rays vary
the maximum speed at which the fin moves through the water. To
maintain a constant frequency and amplitude while changing the
maximum speed of the wing's motion, the time spent during the parts of
the cycle must change. When the ray swims faster the fin moves faster
through the water but takes a correspondingly longer time to change
direction. This is an idea that was proposed by Vanderplank (1950) for
insects but was never supported by data on insect kinematics.

47

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Physiological control of locomotion would be very easy in a system
like that of these rays. A simple pattern generator can. be used to set the
frequency and a simple stretch receptor system could set a constant
amplitude for the stroke. When the animal wants to swim faster a
mechanism which recruits more muscle fibers without changing any
other parameter could produce the increase in forward velocity.
The fact that neither the frequency nor amplitude of the stroke
varies substantially with a threefold change in forward swimming speed
may be attributable to optimum values for the lift force on the wing with
respect to these two variables. The lift force generated by the wing
(discussed in greater detail in Chapter 4) is proportional to the speed of
the wing squared and to the lift coefficient.
The amplitude of the stroke affects only the speed of the wing with
a larger amplitude over the same period resulting in higher wing speeds
and therefore greater lift. The amplitude should be as large as possible
with the limits set by aspects of the fin's morphology. Physical
manipulation of the pectoral fin of anesthetized rays resulted in a
maximum dorsal displacement similar to that seen in free swimming
animals. The fin is capable of flexing below the plane of the body, a
motion that would produce a larger amplitude stoke, but rays rarely
brought their fins below the body while swimming.
The frequency of the stroke cycle affects both the speed of the wing
and the lift coefficient. An increase in frequency with a constant
amplitude will result in an increased wing speed and greater lift. An
increased frequency also results in a shorter period for each stoke, and

48

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
due to unsteady effects a decreased lift coefficient. The lift on a wing is
due to circulation around the wing, a flow pattern that takes a short but
finite time to fully develop, the Wagner effect (Wagner 1925). The shorter
the period the less time to develop the lift producing flow pattern anti
consequently the less lift produced. Ideally if the unsteady aerodynamic
characteristics of ray flight were fully understood the optimum stroke
frequency could be calculated. This however is not yet possible.
This information does not explain how rays swim so effortlessly.
Rays are close to neutrally buoyant and do not need to produce much
vertical force to maintain a position in the water column. They only need
to produce a forward thrust force. Figure 3.2 shows very little motion of
the fin parallel to the direction that the animal is moving, and the
calculated angles of attack are quite small. This rules out any type of a
drag-based locomotion mechanism. Since the motion of the fin is
perpendicular to the direction of the animal's propagation and the angles
of attack are within the range that airfoils can operate without stalling,
rays must be using a lift-based mechanism for locomotion.
A system that generates thrust through a lift based mechanism is
essentially a propeller. Propeller performance is described by a
dimensionless number called the advance ratio. The advance ratio is the
ratio of the relative forward velocity to the angular velocity of the
propeller (Mises, 1945) and is given by the formula:

J = U2%l cod (3.1)

49

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
where U is the forward velocity, COis the angular velocity and d is the
diameter, or in this case the wingspan. Values of the advance ratio for
typical propellers range from .5 to 2.5. Advance ratios for biological
propellers are: penguins 1.0 (Hui, 1988), fruit flies .77 (Vogel, 1966), and
desert locust 1.7 (Jensen, 1956). I calculated the advance ratio for rays
using the maximum tip speed to calculate the angular velocity. This is
an overestimate of this velocity and therefore gives an underestimation of
the advance ratio. The m inim um value for the advance ratio for
swimming rays is 4.0. In other words, the ray is moving forward faster
than the maximum speed at which its wing tip is flapping. When the
ray swims faster the advance ratio increases. This slow flapping of the
fins relative to the forward swimming speed is why rays appear to swim
so effortlessly. How the ray accomplishes this feat is the subject of the
next chapter.

50

with permission of the copyright owner. Further reproduction prohibited without permission.
C hapter 4

Q uasi-Steady F luid
F orce A pproxim ations

Transforming the results from the last chapter, the kinematic


data, into information on the forces produced during those motions
requires knowledge of the aerodynamic characteristics of the ray over the
range of motions observed. There are several methods for acquiring this
aerodynamic information. Unfortunately all of these methods are based
on assumptions or approximations that may hold true for airplanes and
propellers but not for rays.
The easiest method for analyzing the aerodynamics of flapping
animal locomotion is to use the quasi-steady assumption, transforming a
dynamic problem into a series of static situations. The quasi-steady
assumption implies that the forces on an oscillating system are
equivalent to those that would result from steady motion with the same
instantaneous velocities and angles of attack (Ellington, 1984). Although
this assumption is probably not true for the large amplitude flapping
rays, invoking it will provide a first approximation of the forces involved
and illuminate some of the important characteristics of this unique type
of swimming.
A full unsteady analysis of flapping ray locomotion is hampered by
the complex kinematics, involving continuous deformations, and by the

51

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
requirement, stipulated by the available techniques of flutter analysis,
that oscillations be small in amplitude. The only technique for
calculating unsteady forces from flapping animal flight that has been
successful uses information found in the vorticity of the wake of an
animal (Ellington, 1984; Rayner, 1979; Spedding, 1984). This technique
requires a well behaved vortex wake and some prior understanding of the
general mechanics of the system, both of which are unavailable for
flapping ray locomotion (see Appendix A).
Animal locomotion has also been studied by direct force
measurements (Buckholz, 1981; Cloupeau, 1979; Vogel, 1966; Weis-Fogh,
1956). A fundamental difficulty with this approach is that neutrally
buoyant systems moving at constant velocity produce no net force. The
thrust produced exactly balances the drag, and we are left with nothing
to measure. Insects, birds, and some types of fish separate the thrust-
producing apparatus from some of the drag-producing anatomy,
enabling the parts to be measured independently and then summed to get
at the forces of locomotion. Rays unfortunately, do not separate the
thrust producing elements from non-thrust-producing elements. The
wings do not attach to the body at distinct hinges but rather the
amplitude of the oscillations increases continuously through a transition
zone, and the body with its broad wing-like morphology and pitching
motion is probably not simply dragged passively through the water. Also
the fact that rays are essentially neutrally buoyant eliminates the
possibility of setting up a force balance where the animal's weight is
balanced by aerodynamic forces.

52

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
This list of reasons as to why it is difficult to produce a more
accurate and elegant solution to the problem of the mechanics of flapping
ray locomotion defends invoking the, less than ideal, quasi-steady
assumption to calculate the forces from ray swimming.
The fluid force resulting from a ray's motions can be divided into
three components: (1) a lateral component acting along the span of the
wing, (2) lift defined as perpendicular to the velocity and the span of the
wing, and (3) drag defined as parallel to the velocity. Since this study
concerns only steady locomotion, and the animals have paired wings
with symmetrical motion about the body midline, the lateral force will be
neglected. The components of lift and drag can be calculated using the
following equations:

L ^ C Lp S U* (4 .!,

D ^ C d P S U 2 (4 2 )

where p is the density of the fluid, S the projected surface area of the
element, and U is the velocity of that element relative to the water. The
velocity is calculated by taking the square root of the sum of the squares of
the forward swimming speed and the lateral and vertical speeds of the
element in the flapping fin. The force coefficients CLand CDcontain all of
the messy fluid information and can be defined for steady and unsteady

53

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
motion. The challenge in using this method is to accurately estimate
these force coefficients.
To measure the force coefficients a ray was cast in rigid urethane
foam and this cast mounted on a flight balance capable of measuring lift
and drag simultaneously. lif t and drag were measured in steady flow
and these forces were transformed into force coefficients using equations
(1) and (2). The use of the quasi-steady assumption allows these
coefficients to be measured in a steady motion and then applied to the
unsteady motion of the flapping ray. In this chapter the dynamic forces
are calculated using these empirically derived force coefficients
combined with the kinematic data from the last chapter.

M aterials and methods


Model ray
A female cownose ray, Rhinoptera bonasus, with a wing span of 50
cm., was preserved in a 10% formalin solution with the wings in a flat,
relaxed, untwisted, natural position. In this position the ray is fairly flat
and the surface follows the contours of the uncontracted underlying
tissue. I made a two part plaster-of-paris mold of this preserved ray.
The mold was clamped together and two casts were made of orthopaedic
foam (Pedilen rigid foam 200, Otto Bock, Duderstadt, Germany). (I have
since learned of a better source for castable plastics: BJB Enterprises,
Inc. 13912 Nautilus Dr. Garden Grove, CA 92643, (714) 554-2758.) The
casts duplicated the dimensions of the ray to better than 1% except at the

54

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
very thin trailing edge of the wings where the cast is up to 2 mm thicker.
Each cast had a mass of approximately 350 g and an aspect ratio of 3.21,
as defined in Chapter 2.
The aerodynamic force on the model was measured using a two
component flight balance mounted in a wind tunnel. In using the wind
tunnel I invoked the law of dynamic similitude, which states that for
steady flow situations force coefficients will be equal if shape, orientation,
and Reynolds number are equal (Vogel, 1981). Practically this means
that the force coefficients measured in a wind tunnel are equivalent to
those of the animal swimming in water as long as the Reynolds numbers
match.

Wind tunnel
The wind tunnel used is described in Tucker and Parrott (1970)
and Tucker and Heine (1990). iUr speed was measured using a pitot-
static tube connected to an electronic manometer (Datametrics Barocel
1174) calibrated against a water manometer. The variation in air speed
over the region of the tunnel occupied by the model was less than 2%.
Force coefficients were measured for the complete cast at air speeds of
3.5, 7.6,11.3,14.6,and 17.3 m s‘l. These correspond to speeds in water
ranging from 0.25 to 1.2 m s'l. This range of speeds represents Reynolds
numbers ranging from 120,000 to 580,000 based on a wing span of 0.5 m.
This range of Reynolds number exceeds, at both extremes, the range over
which I have observations for swimming animals. There were no
detectable changes in the force coefficients over this range of Reynolds

55

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
number at any angle of attack; therefore the force coefficients were
assumed to be independent of Reynolds number over this range and all
subsequent measurements were done at an air speed of 10.4 m s*l.

Flight balance
The flight balance consists of two orthogonal parallelograms,
similar to the strain gauge balances described in Chapter 6 of Gorlin and
Slezinger (1966). It was mounted inside the wind tunnel and shielded
from the wind by a streamlined shroud. The balance was constructed of
aluminum beams and struts that could be assembled in numerous
configurations. The following configuration was used in this study.

Figure 4.1 Flight Balance

a l mounting
point

drag
| lift
- Strain Gauge

56

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The aluminum beams were 1.27 cm by 0.159 cm thick and 7.62 cm long.
Two strain gauges (Omega, HBM 6/120 LY 63) were mounted on each
beam for a total of four gauges per parallelogram. This arrangement
provided the minimum temperature sensitivity and m ayim nm strain
sensitivity. The struts were made from 1.27 cm aluminum square stock.
The four strain gauges formed the four arms of a Wheatstone bridge and
the output voltage was amplified through a circuit designed around an
Intersil (ICL7605) instrumentation amplifier chip. The amplified
voltages from the imbalance of the four gauges were read on a digital volt
meter. There were separate circuits for the two components of the
balance.

Calibration of the balance


The balance was calibrated for the drag component by hanging
weights from a string attached to the back of the model and passing over
a pulley on a special, low starting torque, ball bearing ( New Hampshire
Bearing SR 168). The lift component was calibrated by placing weights
on the model directly over the mounting attachment. The sensitive axes
of the balance were located and found to form an angle of 89.5°. No
correction was applied for the 0.5° deviation from orthogonal. Loads
applied parallel to the sensitive axis were undetectable on the other
component to better that 1 part in 1000. Loads applied obliquely produced
less than a 1% error from the interactions of the two components. The
wind tunnel was set so that the sensitive axis of the lift component was

57

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
parallel to the gravitational force for all measurements. The calibrations
were linear with correlation coefficients of better than .999, and they were
repeated for each set of measurements. The relative bias of the balance
was less than 1% of the reported values and the imprecision from
repeated measurements was less than 4%.
The model was attached to the balance by a streamlined steel strut
which attached flush to the ventral surface of the model. The mounting
strut had a thickness of 3 mm and a chord of 12 mm, resulting in a
fineness ratio of 4. Approximately 10 cm of the strut was exposed to the
airflow between the model and the shroud covering the balance. The
drag of the exposed section of the mounting strut was measured over the
range of angles used during the experiment and these values subtracted
from the measured values of the model. The strut was undetectable on
the lift component of the balance. In addition, the interference drag of
the mounting strut was investigated following the procedure in Tucker
(1990). The interference drag was below the resolution of the balance.

Experimental protocol
Measuring the force coefficients on the whole model will average
the results for all of the parts but, since the ray does not move uniformly
through the water, an analysis using these coefficients will be
inaccurate. The local force coefficients will vary due to the shape of the
ray changing dramatically from the body to the wing tip. The body is
streamlined and positively cambered in cross section, while proximal

58

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
sections of the wing are uncambered, and distal sections of the wing are
negatively cambered; see figure 4.6. This variation in the shape of the
ray's parts requires that the force coefficients be measured independently
for the varying sections.
Two problems arise when trying to measure the sections
independently. The first is how to divide the animal into sections. The
ray's shape changes continuously with no obvious dividing points. The
second problem is that an isolated section would produce a very different
airflow pattern than that same section would as part of the whole
animal.
To solve the first problem I arbitrarily divided the model's total
wing span into 10 segments of equal length. This arrangement was
chosen because it compliments the measurements of wing twist and
kinematics presented in Chapter 3. I define the body as spanning the
central 4 segments with the wings being the two distal sets of 3 segments
each. Since the body moves uniformly and does not vary as much in
shape, it was measured intact. The wings were separated from the body
along lines parallel to the anterior-posterior axis of the body, and each
was further divided into the 3 separate sections along parallel lines. This
produced 6 wing sections and the isolated body which were each
measured independently.
To solve the problem of maintaining the appropriate overall flow
pattern, the adjacent sections of the model were mounted next to, but not
connected to, the section being measured. In this way the whole model
was mounted in the wind tunnel at each angle of attack, but only one

59

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
section at a time was instrumented. To test the effect of the gap between
the test section and the adjacent sections I measured the force on a
section while varying the distance separating it from the adjacent
sections in increments of 1, 2, 3, 5, 10, 30mm. The results were identical
for 1 and 2 mm, the measured drag force increased by 2% at 3mm, by 7%
at 5mm, by 15% at 10mm and finally by 34% at 30 mm. The lift force did
not change until a gap of 10mm was present. All subsequent
measurements were done with a gap of less than 2 mm, and this gap
was assumed to have no effect on the overall flow pattern and hence on
the results.

Angle of attack
The angle of attack is defined as the angle between the chord of the
segment and the direction of the local fluid velocity. The chord is defined
as the line joining the anterior-most point to the posterior-most point of
the cross section of the segment. The angle between the chord and the
floor of the wind tunnel was measured using a Craftsman automatic
protractor. The angle of the airflow was measured with respect to the
floor of the tunnel by tilting a yawmeter until the dynamic pressure from
the two (over and under) openings were balanced, then inverting the
yawmeter and rebalancing the pressures. The average angle between
the two recordings was used as the direction of local flow. The
imprecision of the angle measurements was less than the resolution of
the protractor's 1° increments. All angle measurements were therefore

60

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
assumed to be accurate to within 1°. The angle of attack was varied by
rotating the mounting strut about the mounting point on the flight
balance. Each section of the model was tested over a range of angles of
attack that covered those measured from swimming an im als (Chapter3).

Tunnel boundary corrections


There are differences between a model moving through still air
and air being blown over a model fixed in a box. These difference are due
to the proximity of the walls of the box and a series of boundary
corrections must be applied to the values of forces measured before
accurate coefficients can be calculated. I checked for the following
effects: solid blocking by the model, horizontal buoyancy from an
increasing boundary layer thickness on the tunnel walls, downwash
effect from the walls intercepting the deflected airflow, and the other
relevant boundary correction factors presented in chapter six of Pope and
Harper (1966). With my small model in a large wind tunnel the only
condition that was detectable was the downwash effect on the measured
drag. The drag coefficients reported have all been corrected for
downwash by the adding the following term to the derived drag
coefficients:

61

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
where C is the cross section of the wind tunnel (1.5 m2), S is the projected
area of the model, CLis the lift coefficient, and 5 is a factor determined by
the configuration of the test, taken to have a value of 0.125 for this study.
All of the model's sections were investigated, and the results from
the corresponding sections from right and left wings were averaged.
There was a good match between the results from the two wings, with
the differences being less than 8% of reported values. The relationships
between the derived force coefficients and the angles of attack were fitted
with third order polynomials using the commercial software package
Cricket Graph (ver. 1.3 by Cricket Software).

Results

The following table summarizes the force coefficients and is


included only for comparison with results from measurements on other
types of wings. The values reported were all taken straight from the data
set with the lift-to-drag ratio calculated at every angle of attack. The
maximum values reported are those with the maximum absolute
magnitude.

62

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Table 4.1 Force Coefficients

Cq .
urrm a@CD .
umm Clumwt {CJCQ}m*x a @{CJ

intact model .022 3° .43 9.0 7°


body only .029 2° .41 6.2 10°
whole wing .024 4° -.77 10.7 7°
proximal wing .017 3° -.50 -10.9 -3°
central wing .029 3° -.85 -6.4 -7°
distal wing .013 2° -.80 -10.8 -6°

These results are based on projected areas of 441.6 cm2 for the body, 181.8
cm2 for the whole left wing, 180.1 cm2 for the whole right wing, 87.4 cm2
for the left proximal wing section, 87.1 cm2 for the right proximal wing
section, 52.1 cm2 for the left central wing section, 53.5 cm2 for the right
central wing section, 39.2 cm2 for the left distal section, and 36.6 cm2 for
the right distal section. The body had a span of 20 cm and each wing
section had a span of 5 cm.
In the coordinate system that I have used upward lift is positive
and downward lift is negative. It is interesting to note that the wing is
better at producing downward lift while the body produces more upward
lift. Another puzzling observation is that the drag coefficient is larger for

63

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
the central wing section than for the distal section, which is subject to
effects of the tip vortex.

Force coefficients vs. angle of attack


The goal of this study is to produce a set of mathematical
relationships between force coefficients and angles of attack so that total
forces can be calculated. These relationships are shown in the sieries of
graphs shown in figures 4.2 through 4.5.

64

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4J2
Body Coefficients
0.12

0. 10 -

0.08-

0.06-

0.04-

0.02
-20 -10 0 10 20

Angle
y = 3.3111e-2 - 2.9473e-3x + 4.8844e-4xA2 RA2 = 0.970

0.6

0.4-

0. 2 -

o.o-

- 0. 2 -

-0.4
-20 -10 0 10 20
Angle
y = - 4.6405e-2 + 3.4964e-2x - 2.7972e-4xA2 RA2 = 0.997

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4.3
Proximal Wing Section
Coefficients
0.12

0. 10 -

0.08-

0.06-

0.04-

0. 0 2 -

0.00
-20 -10 0 10 20
Angle
y = 1.9781e-2 - 1.3405e-3x + 6.6299e-4xA2 - 9.7943e-6xA3 RA2 = 0.938

0.6

0.4-

0. 2 -

- 0. 2 -

-0.4-

- 0.6
-20 -10 0 10
Angle
y = - 3.3827e-2 + 4.3965e-2x - 1.2334e-4xA2 RA2 = 0.997

66

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4.4
Central Wing Section
Coefficients
0.14

0. 12 -

0. 10 -

0.08-

0.06*

0.04-

0.02
-20 -10 0 10 20
Angle
y = 4.2702e-2 - 3.2484e-3x + 5.6445e-4xA2 RA2 = 0.893

1
-20 -10 0 10 20

Angle
y = - 3.9552e-2 + 6.2016e-2x - 4.7737e-4xA2 RA2 = 0.996

67

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
F igure 4.5

Distal Wing Section


Coefficients
0.4

0.3

CD 0.2

0.1

0.0
-20 -10 0 10 20 30

Angle
y a 3.2562e-2 - 4.1755e-3x + 9.8664e-4xA2 RA2 . 0.956

-l
-20 -10 0 10 20 30

Angle
y = - 0.10312 + 8.7119e-2x + 3.4055e-4xA2 - 1.2176e-4xA3 RA2 = 0.

68

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Only the distal wing section was investigated over a large enough
range of angles of attack to observe stall. The drag coefficient of the
proximal wing section has a repeatable W shaped pattern, with the drag
coefficient being minimum at slightly positive and negative angles of
attack and increasing in between when the cord is parallel to the airflow.
These results for force coefficients are reasonable for airfoils
operating at Reynolds numbers and with cross sections similar to the
ray's wing. The following are tracings of the proximal faces from the
sections of the model ray used in this study:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4.6
Dorsal

distal wing section


4 cm

central wing section

proximal wing section

body

Ventral

70

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Calculating thrust
The purpose of measuring these coefficients is to be able to
approximate the thrust produced by flapping these large pectoral fins.
Using equations (1) and (2), the fitted relationship between the force
coefficients and the angles of attack, and the kinematic data (from
Chapter 3) on local velocity and angle of attack were combined to
calculate the lift and drag on each section. These lift and drag forces
were then resolved onto the axes of thrust and a force orthogonal to
thrust in the vertical plane. The spanwise component of the force was
neglected due to the fact that the spanwise force from the right wing
would cancel with that from the left wing. The direction of thrust is
defined as the average direction of the animal's propagation, in practice
horizontal. The orthogonal force is vertical. The forces from the three
wing sections were then summed resulting in an approximation of the
forces produced by a single wing. The results are presented in the
following graph:

71

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 47

Force Generated
by 1 Wing

*— Horiz Force
■s— Vert Force

-6
0 1 2 3
lim e (sec)

For the vertical force up is positive and down is negative. For the
horizontal force positive is forward, or thrust, and negative is drag. The
results reported are the two components of the resultant net force. The
kinematic data covers approximately three seconds and that is reflected
in the duration of these force calculations. The total area between the
line tracing the thrust and the line of 0 force was calculated and divided
by the duration of the measurement with the result being the average
thrust produced. The result is an average net thrust on the wings of
0.1 N.

72

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Discussion

According to the force calculations the ray is flapping its large


pectoral fins yet is producing very little net forward force. The thrust
produced barely overcomes the drag of moving the fins through the
water. This is a problem if the ray is trying to move forward using the
flapping fins as a propeller.
Most of the force produced is directed vertically. These rays are
close to being neutrally buoyant, which means that they do not need to
work to maintain their position in the water column. The only force
needed for swimming is a forward thrust, yet these calculations do not
show the production of much net forward force from the flapping fins.
These calculations of force are based on many measured and
modeled parameters. The results for all of the parts of this process are by
themselves quite reasonable. However, while each of these
measurements has a corresponding level of accuracy, it is difficult to see
how combining the magnitudes of these inaccuracies would affect the
calculated forces. To check for the sensitivity in the force calculations of
errors in the measurements, I have recalculated the forces using values
for the parameters that have all been shifted by their measured
inaccuracy in the direction that would maximize the thrust. The results
from this exercise are presented in the following figure:

73

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 4.8

Best Case Forces

11 H oriz Force
V ert Force

Time (sec)

This alteration gives a small improvement in the calculated thrust. The


fin is overcoming the drag of moving it through the water but is still not
providing a large forward force. The average thrust is 0.4 N. The
vertical force has, however, also increased substantially. This strange
result therefore cannot be explained by errors in the data upon which the
calculations of force are based.
The methods used to calculate the forces are based on several
shaky assumptions. The quasi-steady assumption implies that steady
force coefficients model the dynamic flow interactions accurately. Blade
element analysis assumes that the three dimensional flow patterns do

74

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
not effect the forces. While the implications of invoking these
assumptions cannot be measured or calculated, the direction in which
these assumptions affect the resulting forces can be examined.
The measurements of the force coefficients were done on a three
dimensional model and should contain some information on the loading
pattern of these low aspect ratio wings. There may still be some aspect
which a fully three dimensional analysis would illuminate. Such an
analysis would show an increase in the induced drag and would probably
not show a change in the total lift produced. This would not increase
thrust but would actually decrease the available forward force.
Determining the implications of invoking the quasi-steady
assumption is slightly more complicated. The standard test to determine
the validity of using a quasi-steady analysis is to calculate the reduced
frequency parameter; if its value is less than 0.5 then the unsteady effects
will be small compared to quasi-steady effects (Walker, 1925; Lighthill,
1970). The reduced frequency parameter is a dimensionless number that
relates the angular velocity to the forward velocity of an oscillating
airfoil;

G = (OC/U (4.4)

where 00 is the angular velocity, C is the chord, and U is the forward


velocity. The values of the reduced frequency parameter for rays range
from .4 to 2.5, and so unsteady effects are likely to be very important.
Unsteady effects, however, cannot change the direction of lift and drag,

75

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
only the magnitude and the phasing of those forces (McCroskey, 1982).
With a chord as large as that of a ray's wing, circulation around the
wing is probably not fully developed during each stroke, and therefore the
true value of the lift force is less than that calculated by the quasi-steady
method, the Wagner effect (Wagner, 1925). This as well will result in
decreased thrust.
The ray's wing changes direction twice during each stroke, and a
large mass of water must be accelerated at each change of direction.
This acceleration of fluid results in a force on the ray called the
acceleration reaction (Daniel, 1984). The force from the acceleration
reaction is in a direction that opposes this acceleration. The
accelerations of the ray's wing are vertical in direction and therefore this
force would increase the total vertical force but would not change the
results for thrust.
While these considerations put the magnitude of the reported
forces into serious question, they do not change the conclusion that the
ray’s wings are producing little net thrust and that most of the force
generated by their motion is directed vertically.

Forward motion from a vertical force


There are situations in which a force applied in one direction
results in motion at some angle to that force. Two examples are gliding
and sailing. In gliding the weight of the craft is vertical, towards the
earth, yet the craft descends at an angle determined by the arctangent of

76

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
the lift-to-drag ratio. In sailing the wind blows from one direction
causing a force on the sail in another direction and the balancing of this
aerodynamic force with the hydrodynamic forces on the hull and keel
results in the boat moving in yet another direction. I hypothesize that it
is with a similar scheme that the ray moves forward. The broad wing­
like body of the ray resists vertical motion and instead travels along some
glide path. The ray's wings are like sails, with the relative fluid flow
coming from the flapping motion rather than the wind, and the ray's
body is like a keel, balancing the vertical force and producing forward
locomotion.
The vertical force generated from flapping the fins is an
unbalanced force and, as such, will try to accelerate the whole animal in
a vertical direction. Like a sail or a glider wing the vertical force is
countered by a fluid force at some offset angle, resulting in forward
motion.
This vertical force from the fins, causing the body to pitch, is "the
tail that wags the head". In most cases the tail wagging the head
reduces the efficiency of locomotion. In fish that swim by the
carangiform mode numerous morphological adaptations are present to
minimize the lateral movement at the head of the fish. The reduced area
of the caudal peduncle and the increased depth of the anterior region of
the body are to prevent a side to side motion of the front of the fish. Such a
motion would increase the drag of the body and would create a flow
pattern that might reduce the efficiency of propulsion at the caudal fin.
Rays, on the other hand, use the tail wagging the head, or in this case

77

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
the fins wagging the body, to produce thrust. The important parameter
is the phase relationship between the heaving motion and the pitching
motion. If the movement of the body presents a chord that has a
backward facing component then thrust will be produced, but if the chord
has a forward facing component then the result is drag (Lighthill, 1975).
The ray's broad, wing-like, pitching body resists motion in the
direction of its breadth and instead moves off along some "glide path".
The ray's body pitches 180° out of phase with the wing motion and the
vertical forces from the wings, oscillating up and down, result in a net
forward motion for the swimming rays. Testing this hypothesis is the
subject of the next chapter.

78

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
C hapter 5

U n stead y A nalysis o f the


H eavin g, P itching, Ray Body

When an airfoil moves through a fluid it induces in the fluid a


characteristic velocity profile. The two-dimensional pattern of fluid
motion can be described as composed of a translational motion and a
circulation around the airfoil. Fluid does not physically circulate around
the wing, but without the component of circulation the fluid would be
required to flow from below the wing to above the wing around the sharp
trailing edge. This flow would result in an infinite velocity at the sharp
trailing edge, an impossible situation even in an inviscid fluid. The fluid
must flow smoothly and tangentially off the trailing edge, the Kutta
condition. For every two-dimensional airfoil, in steady motion, there is a
unique value of circulation that produces a velocity profile that satisfies
the Kutta condition. This circulation produces higher flow velocities
above the wing, where it adds to the translational velocity, and slower
flow velocities below the wing, where it subtracts from the translational
velocity. The result of the circulation around the wing is a force on the
wing known as lift that is perpendicular to the translational motion. The
lift due to this circulation can be theoretically calculated using the Kutta-
Joukowski theorem.

79

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Theodorsen (1934) combined the Kutta condition with potential flow
theory to derive the aerodynamic forces on an oscillating airfoil. This
classic analysis has been the foundation for many studies of unsteady
fluid dynamics (McCroskey, 1978). Unfortunately Theodorsen's analysis
involves linearizations that restrict the results to infinitely sm all
oscillations about the equilibrium motion. This restriction prevents me
from being able to use these methods to calculate the forces on the ray's
wings. However, the motion of the ray’s body may be considered to be of
small amplitude, which allows me to use these results to calculate the
forces on the ray's body.
Garrick (1936) used Theodorsen's results and the method
developed by Von Karman (1935) to derive the horizontal force on a
flapping, pitching airfoil. This propulsive force is due to the shedding of
a vortex wake and to leading edge suction. Garrick's analysis still
restricts the oscillations to small amplitude.
Many people have studied the fluid forces on oscillating airfoils in
attempts to better estimate the effects on real wings. Wu (1961) analyzed
the forces on a flexible plate. Ahmadi (1986) using an unsteady lifting-
line method calculated the force on a wing of finite span. Yates (1986)
studied the optimum pitching axis in an oscillating airfoil. Chopra
(1977) used unsteady lifting-surface theory to analyze rectangular and
swept-back wings. The most sophisticated approach for studying
oscillating wings uses a vortex-lattice method developed by Albano (1969)
and used by Lan (1979) to calculate the forces on two dimensional wings
of varying planform shape. All of these methods are based on potential

80

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
flow theory and use linearizations that restrict the results to small
amplitude oscillations. Garrick's (1936) method employs the fewest
assumptions with regard to the kinematics. Therefore I have used the
results presented in Garrick's paper to calculate the thrust from the
pitching heaving ray body.

Theory

The following analysis is taken from Garrick (1936). The motion of


the airfoil is described as a sinusoidal oscillation with two degrees of
freedom: vertical flapping or heaving, perpendicular to the translation,
and pitching about an arbitrary fixed axis (a) parallel to the span. The
airfoil is approximated by a thin flat plate of infinite span moving in an
ideal fluid, as shown in figure 5.1.

Figure 5.1 O scillating Airfoil

81

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The heave (h) and the pitch (a) are described by the following complex-
number equations.

h = h 0e i(Pt + <?2>
( 1)

a = a 0e *& + W

The constants h 0 and CX0 represent the maximum amphtude of the heave
and pitch, 9o and 92 are the phase angles, and p is the radian frequency
of the oscillations. The solutions for Theodorsen's and Garrick's work is
described by an important parameter k, the reduced frequency
parameter, defined as:

k= pb I V (2)

where b is the half chord and V is the free stream velocity.


Equations F, G, and D, Theodorsen's functions, are combinations
of standard Bessel functions of the first and second kind of argument k.

F _ j 1(e71+y 0)+ y 1( y 1-e/ 0)


“ (e71+y0)2+(y1-e/0)2
(3)
Q- y 1y 0+j 1j 0
(j1+y0)2+(y1-e/0)2

82

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
D = Wi + y „)2 + (Yi - J„)2

Garrick uses two methods to calculate the average horizontal force


on an airfoil oscillating with a motion that can be described by the above
equations. The energy formula:
W = E + PXF (4)
where W is the average work done in unit time, E is the average increase
in kinetic energy in the vortex wake in unit time, and P^V is the average
work done in unit time by the propulsive force P*. The other method uses
the force formula:
Px = T t p S 2 + OP (5)
where P* is the propelling force, a is from equation (1), and S and P are
derived in Garrick pages 421 and 427 respectively. The propulsive force
represents thrust when Px is positive and drag when Px is negative.
The propulsive force by the energy formula is:

Px = izpbp 2[A1A02+A2a2+2A4<x0A0] (6)

where Ax = B^Cj, A2 = B2-C2, A4 = B4-C4, b is again the half chord, p


determines the frequency, and where
B1=F

B, = 6s{i(l- a) - (a + 1)[F (1 - a) + &]}

B, = k{(J - 2aF + ^f“)cos((p2-(pQ) - (2 .. Q )sin(<nr<Po)]


&6 11 R

83

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
c 4= [- y sin((p2 - (po) + (i - a)cos(tp2 - (p0)]

The direct calculation of Px results in the following equation:

P x = ir p b p 2 [di-Jil + ( a 2 + b 2 ) a £ + 2 ( a 4 + b 4 ) a 0A 0]

where a i =F 2+ G 2

a2 = b2{(F2 + G 2)[-i- +(i-a)2] + I - ( i - a)F --k ? }


&2 2 4 2 k

a4 = 6 {(F2 + G2)[- f sin((p2 - (p0) + (h - a)cos(<p2 - (p0)]


re 2
- £ cos((p2 - (p0) + & sin((p2 - (p0)}

b4 = | [ ( | + )cos((p2 - <p0) + j sin((p2 - cp0)]

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Calculations of thrust
To use these equations to calculate the thrust from a pitching
heaving ray body I need to provide the kinematic data in the form
presented in equation (1 ). From the results presented in Chapter 3 we
have p ~ 2nf = 2n(l Hz), and the maximum pitching angle a 0 is 10° or
0.1745 rad. This data is from a ray with a body 0.28 m long, resulting in a
half chord b = .14 m. The choice of pitching axis (a) will uniquely define
the remaining terms needed. The easiest pitching axis to choose is the
mid-chord where a = 0. To measure the heave, ho, and the phase angles,
<po and <p2, at a = 0 , 1 plotted the body angle and the vertical displacement
of the midpoint on the line between the nose and the tail. An example of
these results is presented in figure 5.2.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 5.2
Heave and Pitch of the Body

4-
-2
a
2-

I -1 i
& £

Body Angle
M idpoint Y(t)
0 1 2 3

Time (sec)

The average maximum vertical displacement of the center of the body


was 0.03 m and the phase difference between the body pitching and the '
heave (<p2 *<Po) was estimated to be 72° or 1.25 rad. Using these values for
the kinematic parameters, the reduced frequency parameter, k, was
calculated to be 1.5 for an animal swimming at 0.6 m s*l. With this value
of k the functions F, G, and D were evaluated using Mathematica with
the following results:
F = 0.5210
G = -0.0736
D = 1.7384

86

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Plugging these values into both equations for Px, (6 ) and (7), yields the
following result: Px = 3.8 N m'1. The results from the two methods
exactly matched, providing a check on the calculations.
Pxis the force per unit span so to get the force on the body Px needs
to be integrated over the span of the body. If the central 3/5 of a ray with a
total span of 0.5 m is assumed to move as described by equation (1), and if
the change in chord is accounted for, then the thrust produced by the
ray's body is approximately 1 N. It is important to remember that this
method for calculating thrust assumes a perfect fluid and an infinite
span and therefore will not detect drag. The net thrust on the body equals
the total thrust minus the profile drag. Using the minimum drag
coefficient presented in Chapter 4 to calculate the profile drag of the ray's
body moving at the same 0.6 m s'l gives a profile drag of approximately
0.25 N. This results in an average net thrust from the pitching heaving
body of 0.75 N.

Materials and methods

An attempt to test the result from the above calculation was done
by oscillating the model ray body mounted on the flight balance in a flow
tank. These measurements are preliminary, for there is no described
standard method for making measurements on an oscillating model in a
water tunnel.

87

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The same body model that was vised in the experiments described
in Chapter 4 was modified for use in the present study. Mass in the form
of lead shot and steel rods was inserted into the model so that the model
was neutrally buoyant in fresh water. The mass was distributed in a
manner attempting to simulate a uniform density over the model. The
surface of the model was returned to its original shape by filling the holes
with epoxy. When completed the model body had a mass of 1.450 kg.
A streamlined sting with a fineness ratio of 1.5 was attached to the
model in such a way that the model could freely pitch about a pivot point
at the midspan and approximately on the chord of the model. This pivot
point could be set at various distances from the leading edge and was set
at 1/3 of the chord for this experiment. Stops limiting the range of pitch
to 10° were inserted in the mounting system. On a complete ray the
presence of the wings would prevent flow around the spanwise edges of
the model body. Since the wings have been removed from this model,
artificially creating spanwise edges of the body, rectangular pieces of
sheet metal were glued to these cut faces of the model body. The pieces of
sheet metal were the length of the chord by 8 cm high and positioned in a
plane parallel to both the oncoming flow and the motion of the
oscillations of the model.
The end of the sting away from the model was rigidly attached to
the flight balance described in Chapter 4. The flight balance was rotated
90° from the orientation presented in figure 4.1 and was mounted on a
precision slide, (Design Comp Inc.), which restricted the motion to a
linear vertical translation. The balance with the model was oscillated by

88

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
an air cylinder, (Bimba; 043-D-B), controlled by a solenoid air switch,
(MAC; 45A-AA1-DDHC-1BA), and flow control valves on the outlets of the
two ports of the air cylinder. The solenoid was triggered by a square
wave and would alternately send air to one side then the other of the air
cylinder. The vertical position was measured using a 10 turn 10 K£2
potentiometer that was rotated by a string attached to the flight balance at
one end, passing over a pulley wheel mounted on the potentiometer, to a
weight hanging on the other end of the string. With this apparatus the
air pressure, frequency, duty cycle, and speed of the up stroke and down
stoke of the vertical translation could all be controlled independently.
The data was collected using a computer program called Lab View
by National Instruments through a National Instruments A-D board
model NB-MIO-16. The data collected included the vertical force, the
horizontal force, and the vertical position. Calibration curves were
estimated by hanging known weights on the balance, as described in
Chapter 4, and by moving the balance a measured distance and
recording the resulting values with Lab View. These relationships were
used to translate the measurements into force and distance. The drag of
the isolated mounting sting was measured, and the drags of the pieces of
sheet metal were calculated using a drag coefficient of 0.004 (from Vogel,
1981). These extra drag forces were subtracted from the measured
horizontal force. I could not find a method for computing tunnel
boundary corrections for an oscillating model in a water tunnel. The
results have therefore not been corrected for effects due to the walls of the
flow tank. The dorsal surface of the model was at least 14 cm from the

89

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
water's surface, and the ventral surface was at least 10 cm from the
bottom of the tank during the experiment. With this arrangement there
were no obvious surface waves produced by the oscillating model.
The flow tank is described in Vogel (1981), page 298. The working
section of this tank is 35 cm wide and the water depth was 32 cm. During
the experiment the oscillating model was exposed to a flow of 0.45 m s*l
as measured by timing a dye bolus travelling a measured distance in a
measured time.

Results

The results are presented in figure 5.3. The vertical and


horizontal forces are plotted in N and the displacement is plotted in cm.
There are two reference lines on the figure. The solid line is the 0 force
line, and the dashed line is approximately the drag on the model
measured with the flow on but without the model oscillating. The
average drag without the model oscillating is about 0.25 N, which
matches the value calculated from the wind tunnel test. An exact value
for this drag could not be calculated since the model pitched freely
resulting in varying values of the measured drag.
The total area between the line tracing the thrust force and the line
of zero force was calculated and divided by the duration of the
measurement, the result being the average thrust produced. The result
is a net thrust of 0.04 N.

90

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure 5.3

0P) 0
(ura) nopxsojj S a
*
g
£
13 «
.a o oo
“i ’■§k 2h
rCt°l, > eH

It

(^)aoaoj

91

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
D iscussion

The measurements of the forces on an oscillating ray body reported


here are preliminary, and they must be viewed with skepticism. The
resulting curves varied greatly with small variations in the settings of
the input parameters. Many arrangements resulted in net drag.
However, the data presented in figure 5.3 do indicate that it is possible to
produce net thrust by small amplitude oscillations of pitch and heave of a
ray's body.
The vertical force oscillates up and down, similarly to the vertical
forces calculated in Chapter 4. The two vertical forces are, however, not
directly comparable. The results presented in Chapter 4 do not include
inertial effects and the added mass from accelerating nearby fluid. In
principle the unbalanced vertical forces from motion of the wings serves
as the input for the motion of the body, and the net vertical forces taken
over the entire animal cancels.
The data for thrust presented in figure 5.3 shows some odd
features. Thrust is maximal at the point where the vertical force is
maximally downward, which occurs at the beginning of the upward
heaving motion. The thrust also has a repeatable oscillation at about 5
Hz. This oscillation in the thrust was found in all of the data from this
test and does not correspond with any detectable motion. (By contrast the
frequency of oscillations in drag, with the model motionless, was less
than 1 Hz.) The oscillation in thrust may be due to fluid interactions
such as the shedding of a vortex wake or periodic separation of the flow at

92

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
the leading edge. The oscillations in thrust may also be an artifact of my
apparatus.
The results from both the theoretical calculation of thrust, (0.75 N),
and the experimental measurement of thrust agree, (0.04 N), at least
within order of magnitude, with the drag on a stiff ray calculated in
Chapter 4, about 0.5 N. These conclusions are encouraging in light of the
pitching body hypothesis presented in Chapter 4, but more work is still
needed before a thoroughly convincing argument can be produced.
Maximizing thrust is only one of the important parameters that
must drive the morphology and kinematics of myliobatoid rays. It is
equally important, and maybe more important, that the design of these
rays maximizes the hydromechanical efficiency (r|), defined as the mean
forward velocity times the mean thrust to overcome the drag divided by
the mean rate of doing work against the surrounding fluid. Most of the
recent studies on propulsion from oscillating airfoils end with numerical
calculations of the hydromechanical efficiency and the thrust coefficient.
These results are plotted with respect to the reduced frequency
parameter and the feathering parameter (0), introduced by Lighthill
(1970). For small amplitude motions the feathering parameter is the
ratio between the maximum angle of the wing (cO with the forward
velocity, (angle of attack), and the maximum angle achieved by the
pitching axis and the forward velocity:

Ph o

93

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
The maximum angle of the pitching axis is defined as the radian
frequency (p) times the maximum heave amplitude (ha) divided by the
forward velocity (V). The value of the feathering parameter for the
oscillating ray body, analyzed in the beginning of this chapter, is
approximately 0.6. It should be noted here that there are two different
definitions for the reduced frequency parameter commonly used in the
literature and they differ by a factor of 2 . Some authors calculate the
reduced frequency with respect to the half chord and others with respect
to the full chord. Careful reading of these papers is required before the
results can be used.
The range in values of thrust on this oscillating ray body, predicted
from the studies by Lighthill (1970), Chopra (1974,1977), Lan (1979), and
Ahmadi (1986) is from 0.5 to 2.5 N, and the predicted hydromechanical
efficiency ranges from about 0.7 to 0.9.
Recent studies on oscillating airfoils have included results for two
dimensional calculations, rectangular wings versus swept-back wings,
different values of the phase angle (<p) between pitching and heaving, and
different locations of the pitching axis (see Lighthill, 1970; Chopra, 1974,
1977; Lan, 1979; Ahmadi, 1986). All of the studies show two important
trends. Hydromechanical efficiency increases with decreasing reduced
frequency and with increasing values of the feathering parameter. At
the same time the thrust coefficient increases with increasing reduced
frequency and with decreasing values of the feathering parameter.
Feathering parameters above one result in net drag. The implications of
these results is that as unsteady effects increase, (increasing the reduced

94

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
frequency parameter), the amount of thrust produced increases but the
efficiency of producing that thrust decreases. This is an interesting
result in light of the fact that rays do muscular work in moving the
pectoral fins, with their shorter chord and subsequently lower reduced
frequency parameter, while producing thrust with the oscillations of the
body with its longer chord and higher reduced frequency parameter.
Based on these criteria rays appeared to be well designed for efficient
locomotion.

95

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Chapter 6

C onclusions

Do rays fly underwater in the same manner as do birds through


the air? Superficial observations indicate that they do fly like birds. Both
kinds of animals flap a pair of wing-like appendages, extending from the
pectoral girdle, in a motion that is perpendicular to the anterior-posterior
axis of the body. However, upon closer examination the dynamics of ray
locomotion rapidly diverges from that of bird flight. Certainly birds don't
possess the dorso-ventrally flattened body that is characteristic of rays
and a requirement for the pitching body hypothesis presented in this
thesis.
Rays are unique animals. They do not look like or behave as any
other creature. In spite of their many fascinating features, rays have
largely been ignored by the scientific community. On every occasion that
I dissected or observed one of these animals I was introduced to another
puzzling aspect of their life history. This study is far from exhaustive,
even with regard to steady locomotion. The current level of
understanding of the mechanics of flapping ray locomotion, and for that
matter of flapping bird flight, is far too primitive to be able to answer
thoroughly the question posed above.
In Chapter 4 ,1 concluded that the flapping pectoral fins do not
produce thrust. This conclusion is based on an analysis that treats the

96

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
fin as an airfoil in a series of steady-flow situations. This method is not
capable of detecting all aspects of the fluid forces; specifically it misses
unsteady effects. Leading edge suction probably plays a s ig n ificant, role
and can not be detected in a quasi-steady analysis. Another important
aspect of the hydrodynamics that is not accounted for is the shedding of a
starting and stopping vortex at the ends of each half-stroke. The
resulting complicated vortex wake may contain a net rearward
component of momentum, which would result in a reactive forward
thrust on the ray (Weihs, 1972). These considerations do not change the
conclusion that there are large unbalanced forces that try to accelerate
the body vertically and that the pitching body plays a significant role in
producing thrust. Significant leading edge suction or thrust from the
shedding of a vortex wake would merely indicate that the wings are
doing more than simply pushing the body up and down.
The idea that the entire animal is involved in producing thrust
indicates that these rays have come full circle from the primitive body-
undulation type of swimming. In fish there is a progression, based on
swimming performance, from a uniform body undulation over the entire
animal to a concentration of the motion so that only the high-aspect-ratio
caudal fin oscillates. In the later case the thrust-producing anatomy has
been separated from the rest of the body as with a ship and its propeller.
The hydrodynamics of the large-amplitude flapping type of ray
locomotion is more similar to that of the oscillating lunate tail. Rays are
a crescent shaped flying wing with the chord less than the span and with
a cross section similar in profile to the that of a lunate tail. In these

97

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
flapping rays the entire animal is the propeller or the "caudal fin", and
there is no parasite drag from a passive body. The implication is that
rays have a relatively larger trailing edge; therefore, for the same mass
of muscle, the ray can induce a larger mass of water to move than a
thunniform fish. The force produced by moving water is proportional to
the mass of water times the velocity of that water, and if more water is
moved then the same force will be produced with a slower water velocity.
There are benefits to moving a larger mass of water more slowly; The
energy input required to move fluid is proportional to the mass times the
velocity squared and less energy is required to move a larger mass of
water more slowly. In swimming, force is produced when energy is
expended in moving fluid, and producing the same amount of force for
less energy improves efficiency. Depending on the direction in which the
fluid is moved, flapping rays, by using their entire body, should be very
efficient swimmers.
Other aspects of improved efficiency result from the morphology of
these rays. Rays have a sharp trailing edge, preventing flow from below
the animal to above, and therefore inducing the circulation required for
lift. At the same time the leading edge of the ray is rounded which
allows flow to rapidly pass from below to above the ray without
separating. This flow pattern produces the leading edge suction force.
The general planform of these rays includes a swept-back leading edge.
Wu (1971b) and Chopra (1977) showed that swept-back leading edges
reduces the thrust due to leading edge suction while still producing the
same total thrust. A reduced dependence on leading edge suction limits

98

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
the chances for flow separation at the leading edge which would
drastically reduce the hydromechanical efficiency. The swept-back
crescent shaped wing has also been shown to have a higher lift to
induced drag ratio (Van Dam, 1987). The triangular shape of rays
maintain the swept-back leading edge while providing the largest span
and minimizes the area subject to the wing-tip effects. The similarities
in the overall morphology of all the myliobatoid rays indicates the
importance of this shape for hydromechanical efficiency.
In this thesis I have merely begun to explain steady swimming.
Aspects of maneuvering and acceleration are important untouched
questions that need to be addressed. Careful thought about my
hypothesized mechanism for flapping ray swimming reveals that
starting from rest on the bottom poses a difficult problem. Rays may be
using a fling type of mechanism (Weis-Fogh, 1973; Lighthill,1973) for
starting from the bottom, with the substrate acting as the opposing wing.
This is one idea that I would like to investigate further.
The evolution of this large-amplitude flapping type of ray
swimming is another interesting problem. In extant species of batoids
there is an almost continuous spectrum of methods for locomotion, from
the body undulations in sawfish, to the pectoral fin undulations of skates,
to the large amplitude flapping of the eagle rays. Mapping these modes
of locomotion onto a phylogenetic tree and imagining the mechanics of
intermediate forms of locomotion may reveal interesting insights.
My data for flapping rays comes primarily from the species
Rhinoptera bonasus, but I do have some data on Myliobatis freminvillei,

99

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
and I have watched films of individuals in the species Aetobatis narinari
swim. I believe that the conclusions presented in this study can be
generalized to the family Myliobatidae. I have also watched film s of
individuals in the family Mobulidae, and the details of their swimming
movements are different from those of the eagle rays. Manta rays flap
their fins in a motion which is symmetrical about the plane of the body.
In addition, the most proximal section of the pectoral fin initiates the flap
so that it appears as though there is an undulation progressing along the
span. These differences in the kinematics may produce significant
differences in the mechanics of Mobulidae versus Myliobatidae
locomotion. However, the general body plan for the two groups of rays is
so similar that I believe that the general mechanism for locomotion is
similar. Clearly there is a lot of work left to be done on the mechanics of
flapping ray locomotion.

100

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
HOW A RAY SWIMS

&&
PRESSURE
OTHER FISH

HEAVING
HOPE

TH RU ST

AILERONS
H ALYARD
WING

C IS T

SPERM ATOZOA

COSINE

PITCHING

Figure l. The Heaving Pitching Ray Body

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A ppendix A

Flow V isualization

The analysis of the fluid mechanics of animals locomoting using


paired appendages has relied on two approaches: blade element analysis
of the limb, and analysis of the momentum of the fluid in the wake of the
animal.
In blade element analysis, the wing is divided into a series of
chordwise elements and the two dimensional forces, on each element,
are calculated using the instantaneous velocity of the element relative to
the fluid (Osborne, 1951). The forces are then summed over the entire
wing (Mises, 1945). A fundamental assumption of this analyses is that
the chordwise elements act independently of each other; in other words,
there is no spanwise flow. Daniel (1987) used this method to analyze the
undulating fin type of ray locomotion, but maintained that the approach
was only valid for low amplitude undulations of the fin and probably did
not apply to the large amplitude flapping motion. To determine if blade
element analysis is valid for analyzing the large amplitude flapping type
of ray locomotion, I measured the surface flow pattern on a free
swimming cownose ray.
The wake from a body moving through a fluid contains all of the
information about the forces on that body. To use the momentum in the
wake to calculate the forces produced requires that the wake form

101

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
regular structures. Typically the wake from flapping, lift-producing
airfoils rolls up into vortex rings. The rate of momentum shed to the
wake, contained in these rings, is equal in magnitude and opposite in
direction to the force on the animal. These vortex rings are formed
because in generating lift circulation is developed around the wing, and
because vortices cannot end in the fluid this circulation must continue
into the wake forming a closed loop (Ellington, 1984; Rayner, 1979). To
test whether or not this approach could be used for analyzing ray
locomotion, I injected dye into the water near the surface of the animal
and observed the pattern formed in the wake.

Surface flow pattern


To determine the flow pattern over the surface of the animal, I
attached pieces of embroidery thread to the skin of a cownose ray.
Embroidery thread was chosen after experimenting with several types of
yam and finding that only the embroidery thread produced no detectable
deviation from orientating with the flow. These threads line up with the
streamlines, allowing me to visualize the surface flow pattern.
An individual ray was anesthetized with MS222. Using surgical
sutures, I attached 10 threads, each approximately 6 cm long, to the
dorsal surface of the animal. Each thread was attached at its midpoint
leaving the two ends free. After the animal recovered I filmed her from
above while she swam in still water in the raceway. The images were
recorded on videotape and analyzed following the procedure described in

102

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Chapter 3. The data recorded included the direction of the animal's
propagation, assumed to be represented by a line from the nose to the tail,
and the angles of the 10 strings. The direction of the animal's
propagation is used as the free stream direction for all calculations. The
threads were cut shorter and the observations repeated to test for the
effects of the thread length. No difference was detected. Also recorded
were the positions of the strings on the ray, and the frames that
represent the top and bottom of the stroke. The animal was assumed to
be swimming in a horizontal plane. I could not duplicate this
experiment on the ventral side of the animal.
The following table summarizes the data for a typical sequence of
one and a half strokes. Reported are the angles between the strings and
the defined free stream averaged over the sequence, and the standard
deviation of those angles during the sequence. The location of the
attachment of the string is reported as the percentage of the distance
from the midline to the wing tip, with the right wing being positive and
the left negative. A value of 100% refers to the right wing tip. In a
similar way, the position on the chord is presented as the percentage of
the distance from the leading edge to the trailing edge.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Table A.1 Surface H ow Pattern

Spanwise Cordwise Average Standard


Strinsr Position Position Ansrle fdesr) _Deviation

1 -90.5 31.0 -28.3 43.1


2 -64.6 45.5 5.8 7.6
3 -41.9 71.4 11.0 5.2
4 -32.9 16.7 2.7 2.7
5 -31.3 42.1 4.7 3.8
6 1.3 50.1 7.2 5.4
7 19.3 24.7 5.5 2.5
8 24.7 93.7 0.8 6.1
9 47.2 80.7 -3.9 6.2
94.4 50.1 33.5
00
10 1

The imprecision was measured by repeating the measurements


three times and the average standard deviation calculated. The
imprecision of the body angle was less than 1°, the imprecision for the
angles of all the strings, except the two closest to the wing tips, was less
than 3°, and the imprecision for the two strings near the wing tips was
less than 11°.
Figure A.1 represents the data from the same sequence. The
angles reported are between the string and the defined free stream and
they are plotted against frame number with the top and bottom of the
strokes indicated.

104

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Figure A.1

Surface Flow Pattern

-----o---- String 1
-----o---- String 2
-----□— String 3
bo
<u — ■— String 4
-----•— String 5
-----* — String 6
'bo
G -----A----
< String 7
-----■— String 8
- 0 String 9
\?W V?i
S A + y
A* $ String 10

Frame Count (30 frames/sec)

From the data it is clear that the flow near the wing tip does not
line up with the direction of the animal's propagation and therefore blade
element analysis will be inaccurate. I expected the results at the wing
tip to influence the flow over the whole wing since this is a low aspect
ratio wing. The flow over the rest of the animal does line up fairly well
with the defined free stream flow, which would indicate that blade
element analysis may be accurately used over most of the animal.

105

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Wake flow pattern
The flow pattern in the wake contains all of the information
regarding the forces on the animal, but it is difficult to get at this
information unless the wake forms recognizable structures. To see if the
wake formed systematic structures I tried several flow visualization
techniques, including swimming the animal through a cloud of dye or
particles, and injecting dye upstream of the animal swimming in the
flow tank. The technique that was most successful was to suture thin
plastic tubes to the ray and, using a syringe, fill them with dye. The dye
escapes from the tube and traces a line describing the path of a fluid
particle that passed through the point near the surface of the ray.
Experimenting with a number of tube types I found one from which the
dye would flow slowly as the animal swam. Four tubes, one near each
wing tip and one at the midpoint of each wing, were sutured to the dorsal
surface of an individual, with the tubes parallel to the long axis of the
body. The tubes were filled with dye, and the animal was photographed
as it swam. One camera was placed underwater behind the animal, and
another viewed the animal laterally through the window in the side of
the tank. The water in the tank was allowed to settle between each run,
minimizing any flow not due to the ray.
While the results of this experiment were inconclusive, a few
observations can be made. From the lateral view, with the dye
originating from the front end of the tube closest to the body, it appears
that the wake undulates with an amplitude greater than the movements
of the body. The pattern of the wake is a level section followed by an

106

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
mushroom shaped upward deflection, then another level section followed
by a mushroom shaped downward deflection. The posterior view of this
sequence produced no consistent pattern. In the posterior view, with the
dye originating from the back of the tube at the wing tip, it appears that
there are significant tip vortices. Analyzing this posterior view and the
corresponding lateral view of this sequence indicates that the dye flows
off the wing tip in a helix, probably changing handedness from the
upstroke to the downstroke. It is not clear from this data whether or not
the wake forms vortex rings. It is clear that learning the mechanics of
flapping ray swimming from an application of the vortex wake theory
would be difficult.
In addition to the flow visualization information presented above,
there is interesting information in the shape and location of the surface
wave, over the animal, seen in the lateral photographs. On the upstroke
the crest of the wave is posterior to the animal, and the trough is over the
anterior part of the animal. On the downstroke the crest of the wave
moves over the animal, and at the bottom of the downstroke the crest of
the wave has moved anterior to the animal. This pattern was seen in all
of the lateral sequences of photographs. If you use the height of the free
surface of the water as a manometer, this pattern implies that there is a
pressure gradient along the anterior-posterior length of the body, and it
changes dining a stroke cycle.

107

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
A ppendix B

A ngle D ata

Table B.1
Angle of Attack (deg)

Time (sec) Distal Central Proximal Wing Base

0.067 14.38 9.40 6.03 1.56


0.1 13.54 8.85 5.70 1.48
0.133 13.70 9.05 5.51 1.51
0.167 15.52 10.52 5.71 1.73
0.2 18.49 13.15 5.70 2.12
0.233 13.10 10.20 3.11 1.63
0.267 4.63 4.81 -0.79 0.71
0.3 -6.06 -3.46 -5.47 -0.80
0.333 -12.52 -8.56 -7.90 -1.69
0.367 -18.18 -12.61 -9.01 -2.25
0.4 -21.13 -14.41 -8.88 -2.37
0.433 -22.09 -14.78 -8.01 -2.30
0.467 -19.86 -13.16 -6.44 -1.99
0.5 -15.56 -10.60 -4.55 -1.61
0.533 -9.12 -6.62 -2.46 -1.04
0.567 -5.28 -4.23 -1.28 -0.70
0.6 -2.30 -2.36 -0.35 -0.42
0.633 -1.92 -2.23 -0.18 -0.41
0.667 0.37 -0.80 0.59 -0.20
0.7 4.41 1.86 1.95 0.20
0.733 7.35 3.84 3.08 0.52
0.767 8.61 4.74 3.68 0.68
0.8 10.34 6.10 4.39 0.93
0.833 14.54 9.34 5.95 1.50
0.867 17.75 11.91 7.29 1.98
0.9 17.95 12.18 7.54 2.06
0.933 17.06 11.56 7.14 1.97
0.967 18.04 12.33 7.18 2.08
1 17.15 11.72 6.24 1.93

108

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
1.033 12.59 9.09 3.68 1.46
1.067 1.30 2.11 -1.08 0.31
1.1 -8.15 -4.99 -5.71 - 1.02
1.133 -13.82 -9.63 -8.32 -1.90
1.167 -18.66 -13.23 -9.61 -2.44
1.2 -22.03 -15.44 -9.91 - 2.66
1.233 -25.76 -17.85 -10.03 -2.90
1.267 -24.46 -16.51 -8.43 -2.55
1.3 -17.49 -11.55 -5.47 -1.73
1.333 -6.84 -4.68 -1.96 -0.71
1.367 -0.55 -0.99 0.07 - 0.20
1.4 1.10 - 0.01 0.61 -0.07
1.433 -2.19 -2.28 -0.32 -0.41
1.467 -4.27 -3.68 -0.92 -0.62
1.5 -3.38 -3.30 -0.55 -0.58
1.533 I.73 -0.05 1.17 - 0.10
1.567 6.91 3.47 3.04 0.46
1.6 10.98 6.47 4.58 0.96
1.633 13.38 8.36 5.55 1.31
1.667 14.04 8.97 5.82 1.44
1.7 14.48 9.42 6.09 1.55
1.733 14.89 9.83 6.30 1.65
1.767 15.03 10.01 6.44 1.70
1.8 15.68 10.54 6.39 1.77
1.833 17.86 12.22 6.62 2.02
1.867 18.23 12.94 5.63 2.08
1.9 7.48 6.53 0.99 1.03
1.933 -5.24 -2.80 -5.01 -0.67
1.967 - 12.11 -8.65 -8.61 -1.87
2 -15.40 -10.94 -9.16 -2.15
2.033 -19.82 -13.82 -9.45 -2.42
2.067 -23.80 -16.34 -9.42 -2.65
2.1 -24.77 -16.77 -8.38 -2.59
2.133 -17.06 -11.37 -5.23 -1.71
2.167 -9.27 -6.47 -2.61 - 1.00
2.2 -1.25 -1.45 - 0.12 -0.27
2.233 0.67 -0.42 0.56 -0.14
2.267 -1.70 - 2.11 - 0.10 -0.40
2.3 -3.94 -3.77 -0.67 - 0.66
2.333 - 0.10 -1.30 0.62 -0.29
2.367 6.92 3.49 3.05 0.46
2.4 II.53 6.88 4.76 1.03
2.433 12.49 7.69 5.28 1.20
2.467 13.46 8.55 5.81 1.39
2.5 15.04 9.89 6.60 1.66
2.533 16.19 10.87 7.09 1.86

109

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.567 17.44 11.87 7.38 2.03
2.6 17.36 11.82 7.09 2.00
2.633 18.50 12.65 7.25 2.12
2.667 13.35 9.08 4.56 1.46
2.7 1.06 1.37 -0.29 0.22
2.733 -8.34 -4.96 -4.59 -0.89
2.767 -11.70 -7.69 -6.67 -1.45
2.8 -11.85 -7.89 -7.12 -1.53
2.833 -16.31 -11.47 -8.99 -2.17
2.867 -20.97 -14.78 -9.90 -2.61
2.9 -25.87 -18.00 -10.10 -2.95

Table R2
Relative Fluid Angle (deg)

Time (sec) Distal Central Proximal Wing Base

0.067 -28.93 -19.78 -10.77 -3.24


0.1 -27.26 -18.64 -10.18 -3.07
0.133 -24.74 -16.85 -9.20 -2.78
0.167 -22.00 -14.90 -8.11 -2.45
0.2 -13.28 -8.84 -4.76 -1.43
0.233 3.93 2.59 1.39 0.42
0.267 24.62 16.79 9.17 2.77
0.3 38.23 27.35 15.43 4.72
0.333 42.69 31.01 17.67 5.42
0.367 42.29 30.38 17.07 5.19
0.4 39.34 27.52 15.04 4.50
0.433 33.44 22.54 11.93 3.51
0.467 24.74 16.05 8.26 2.40
0.5 13.08 8.21 4.14 1.19
0.533 2.95 1.83 0.91 0.26
0.567 -2.38 -1.47 -0.73 -0.21
0.6 -6.47 -4.01 -2.01 -0.58
0.633 -8.14 -5.07 -2.54 -0.73
0.667 -11.88 -7.46 -3.76 -1.08

110

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
0.7 -17.41 -11.08 -5.65 -1.63
0.733 -22.31 -14.44 -7.45 -2.17
0.767 -24.89 -16.31 -8.50 -2.49
0.8 -26.45 -17.56 -9.27 -2.74
0.833 -29.67 -20.08 -10.79 -3.22
0.867 -32.53 -22.44 -12.26 -3.69
0.9 -32.99 -22.95 - 12.66 -3.83
0.933 -31.04 -21.56 -11.91 -3.62
0.967 -29.09 -20.13 - 11.11 -3.37
1 -23.11 -15.70 -8.56 -2.59
1.033 -6.53 -4.31 -2.31 -0.69
1.067 18.92 12.72 6.90 2.08
1.1 36.22 25.75 14.48 4.43
1.133 42.43 30.95 17.74 5.47
1.167 43.63 31.76 18.10 5.55
1.2 42.61 30.50 17.03 5.16
1.233 39.80 OP 7 ft 15.11 4.50
1.267 32.49 21.74 11.42 3.34
1.3 20.36 13.00 6.61 1.91
1.333 5.24 3.26 1.63 0.47
1.367 -5.92 -3.67 -1.83 -0.52
1.4 -8.82 -5.49 -2.75 -0.79
1.433 -6.47 -4.02 - 2.01 -0.58
1.467 -4.63 - 2.88 -1.45 -0.42
1.5 -7.95 -4.97 -2.50 -0.72
1.533 -15.79 - 10.02 -5.10 -1.48
1.567 -23.19 -15.05 -7.78 -2.27
1.6 -27.68 -18.33 -9.63 -2.83
1.633 -29.88 -20.09 -10.70 -3.17
1.667 -29.61 -20.05 -10.78 -3.22
1.7 -29.72 -20.30 - 11.02 -3.31
1.733 -29.58 -20.35 -11.14 -3.36
1.767 -29.71 -20.55 -11.32 -3.43
1.8 -27.77 -19.12 -10.52 -3.19
1.833 -24.94 -17.03 -9.32 -2.82
1.867 -13.34 -8.87 -4.77 -1.43
1.9 13.29 8.84 4.76 1.43
1.933 36.49 25.97 14.62 4.47
1.967 45.24 33.53 19.49 6.05
2 45.07 33.11 19.03 5.86
2.033 43.06 30.91 17.31 5.25
2.067 39.05 27.17 14.75 4.40
2.1 31.13 20.74 10.87 3.18
2.133 18.22 11.58 5.88 1.70
2.167 5.81 3.62 1.81 0.52
2.2 -5.25 -3.26 -1.63 -0.47

ill

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
2.233 -9.86 -6.15 -3.09 -0.89
2.267 -8.68 -5.41 -2.72 -0.78
2.3 -8.53 -5.33 -2.69 -0.77
2.333 -14.12 -8.95 -4.56 -1.32
2.367 -23.10 -15.00 -7.76 -2.26
2.4 -28.02 -18.59 -9.77 -2.88
2.433 -29.42 -19.76 -10.52 -3.12
2.467 -30.49 -20.75 -11.20 -3.35
2.5 -32.23 -22.26 -12.19 -3.68
2.533 -32.70 -22.79 -12.60 -3.82
2.567 -31.86 -22.23 -12.33 -3.75
2.6 -29.74 -20.62 -11.40 -3.46
2.633 -28.92 -19.97 -10.99 -3.33
2.667 -15.98 -10.64 -5.73 -1.72
2.7 8.31 5.49 2.94 0.88
2.733 28.58 19.71 10.84 3.29
2.767 36.46 25.96 14.62 4.47
2.8 39.03 28.02 15.86 4.86
2.833 43.33 31.55 18.00 5.53
2.867 43.31 31.24 17.59 5.35
2.9 39.36 27.53 15.04 4.50

112

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
R eferences

Ahmadi, A.R. and Widnall, S.E. (1986). Energetics and optimum motion
of oscillating lifting surfaces of finite span. J.Fluid Mech.
162,261-282.

Albano, E. and Rodden, W. (1969). A doublet-lattice method for


calculating lift distributions on oscillating surfaces in subsonic
flows. AIAA Journal 7(2), 279-285.

Bainbridge, R. (1958). The speed of swimming of fish as related to size


and to the frequency and amplitude of the tail beat. J. exp. Biol.
35,109-133.

Batchelor, G.K. (1967). An Introduction to Fluid Dynamics. Cambridge


University Press, Cambridge, U.K.

Bigelow, H.B. and Schroeder, W.C. (1953). Fishes of the Western North
Atlantic, part II. Sears, New Haven, Conn.

Bone, Q. (1989). The locomotion of rays. Mar. Bio. Assoc, of the U.K.
Annual Report 1988-1989, pp 24-25.

Buckholz, R.H. (1981). Measurements of unsteady periodic forces


generated by the blowfly flying in a wind tunnel. J. exp. Biol.
90,163-173.

113

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Calvert, R.A. (1983). Comparative anatomy and functional morphology of
the pectoral fin of stingrays. Masters Thesis, Duke University.

Chopra, M.G. (1974). Hydrodynamics of lunate-tail swimming


propulsion. J. Fluid Mech. 64, 375-391.

Chopra, M.G. and Kambe, T. (1977). Hydrodynamics of lunate-tail


swimming propulsion. Part 2. J. Fluid Mech. 79, 49-69.

Clark, E. (1963). Massive aggregations of large rays and sharks in and


near Sarasota, Florida. Zoologica 48(6), 61-65.

Cloupeau, M., Devillers, J.F. and Devezeaux, D. (1979). Direct


measurements of instantaneous lift in desert locust; comparison
with Jensen's experiments on detached wings. J. exp. Biol.
80,1-15.

Compagno, L.J.V. (1977). Phyletic relationships of living sharks and


rays. Amer. Zool. 17, 303-322.

Daniel, T.L. (1984). Unsteady aspects of aquatic locomotion. Amer. Zool.


24,121-134.

Daniel, T.L. (1988). Forward flapping flight from flexible fins. Can. J.
Zool. 66,630-638.

Ellington, C. (1984). The aerodynamics of hovering insect flight (I-VI).


Phil. Trans. R. Soc. London B. 305,1-181.

114

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Garrick, LE. (1936). Propulsion of a flapping and oscillating airfoil.
NACA tech. Rep. no 567, 419-427.

Gorlin, S.M. and Slezinger, I.I. (1966). Wind Tunnels and Their
Instrumentation. Israel Program for Scientific Translations,
Jerusalem, Israel.

Hui, C.A. (1988). Penguin swimming. I. Hydrodynamics. Physiol. Zool.


61(4), 333-343.

Hunter, J.R. and Zweifel, J.R. (1971). Swimming speed, tail beat
frequency, tail beat amplitude and size in jack mackerel,
Trachurus symmetricus, and other fishes. Fish. Bull.
69(2), 253-266.

Jensen, M. (1956). Biology and physics of locust flight. IH. The


aerodynamics of locust flight. Phil. Trans. R. Soc. B 239, 511-552.

Klausewitz, W. (1964). Der lokomotionsmodus der Flugelrochen


(Myliobatoidei) Zool. Anz. 173, 111-120.

Lan, C.E. (1979). The quasi-vortex-lattice method with applications to


animal propulsion. J. Fluid Mech. 93, 747-765.

Lighthill, M.J. (1970). Aquatic animal propulsion of high hydrodynamic


efficiency. J. Fluid Mech. 44, 265-301.

115

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Lighthill, M.J. (1973). On the Weis-Fogh mechanism of lift generation.
J. Fluid Mech. 60,1-17.

Lighthill, M.J. (1975). Mathematical Biofluiddynamics. Society for


Industrial Applied Mathematics, Philadelphia, PA.

Marey, M. (1893). Des mouvements de natation de la Raie. Comptes


Rendus Acad. Sci. Paris 116, 77-81.

McCroskey, W.J. (1982). Unsteady airfoils. Ann. Rev. Fluid Mech.


14, 285-311.

Mises, R. von (1945). Theory of Flight, Dover, New York, N.Y.

Nelson, J.S. (1984). Fishes of the World. 2nd ed. Wiley and Sons, New
York, N.Y.

Oehmichen, E. (1958). Locomotion des poissons. Grassse, Traite de


Zoologie 13(1), 819-853.

Osborne, M.F.M. (1951). Aerodynamics of flapping flight with


applications to insects. J.exp. Biol. 28, 221-245.

Pope, A. and Harper, J.J. (1966). Low Speed Wind Tunnel Testing. Wiley
& Sons, New York, N.Y.

Rayner, J.M. (1979). A new approach to animal flight mechanics. J. exp.


Biol. 80,17-54.

116

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Schwartz, F.J. (1965). Inter-American migrations and systematics of the
western Atlantic cownose ray, Rhinoptera bonasus. Assoc. Island
Mar. Lab. Meet. 6.

Schwartz, F.J. (1967). Embryology and feeding behavior of the Atlantic


cownose ray, Rhinoptera bonasus. Assoc. Island Mar.
Lab. Meet. 7.

Smith, J.W. and Merriner, J.V. (1985). Food habits and feeding behavior
of the cownose ray, Rhinoptera bonasus, in lower Chesapeake Bay.
Estuaries 8(3), 305-310.

Smith, J.W. and Merriner, J.V. (1986). Observations on the reproductive


biology of the cownose ray, Rhinoptera bonasus, in Chesapeake
Bay. Fish. Bull. U.S. 84, 871-877.

Smith, J.W. and Merriner, J.V. (1987). Age and growth, movements and
distribution of the cownose ray, Rhinoptera bonasus, in
Chesapeake Bay. Estuaries 10(2), 153-164.

Spedding, G.R. (1986). The wake of a jackdaw (Corvus monedula) in slow


flight. J. exp. Biol. 125,287-307.

Spedding, G.R., Rayner, J.M., and Pennycuick, C.J. (1984). Momentum


and energy in the wake of a pigeon (Columba livia) in slow flight.
J. exp. Biol. I l l , 81-102.

117

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Theodorsen, T. (1935). General theory of aerodynamic instability and the
mechanism of flutter. NACA tech. Rep. no 496 413-433.

Tucker, V.A. (1990a). Body drag, feather drag and interference drag of
the mounting strut in a peregrine falcon, Falco peregrinus.
J. exp. Biol. 149,449-468.

Tucker, V.A. (1990b). Measuring aerodynamic interference drag between


a bird body and the mounting strut of a drag balance. J. exp. Biol.
154,439-461.

Tucker, V.A. and Heine, C. (1990). Aerodynamics of gliding flight in the


Harris' hawk, Parabuteo unicinctus. J. exp. Biol. 149, 469-489.

Tucker, V.A. and Parrot, C.G. (1970). Aerodynamics of gliding flight in a


falcon and other birds. J. exp. Biol. 52, 345-367.

Van Dam, C.P. (1987). Efficiency characteristics of cresent-shaped wings


and caudal fins. Nature 325, 435-437.

Vanderplank, F.L. (1950). Air-speed / wingtip-speed ratios of insect


flight. Nature 165, 806-807.

Vogel, S. (1981). Life in Moving Fluids, Princeton University Press,


Princeton, N.J.

Vogel, S. (1966). Flight in Drosophila. I. Flight performance of tethered


flies. J. exp. Biol. 44, 567-578.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Von Karxnan, T. and Burgers, J.M. (1935). General aerodynamic theory -
Perfect fluids, in Aerodynamic Theory vol II. W.F. Durand ed.
Springer, Berlin.

Von Karman, T. and Sears, W.R. (1938). Airfoil theory for non-uniform
motion. J. aeronaut. Sci. 5, 379-390.

Wagner, H. (1925). Uber die entstschung des dynamaischen auftriebes


von tragflugeln. Z. angew. Math. Mech. 5, 17-35.

Walker, G.T. (1925). The flapping flight of birds. J. R. aeronaut. Soc.


29,590-594.

Weihs, D. (1972). Semi-infinite vortex trails, and their relation to


oscillating airfoils. J. Fluid Mech. 54, 679-690.

Weis-Fogh, T. (1973). Quick estimates of flight fitness in hovering


animals, including novel mechanisms for lift production. J. exp.
Biol. 59,169-230.

Weis-Fogh, T. (1956). Biology and physics of locust flight. II. Flight


performance of the desert locust Schistocerca gregaria. Phil.
Trans. R. Soc. B 239,459-510.

Wu, T.Y. (1961). Swimming of a waving plate. J. Fluid Mech. 10, 321-344.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Wu, T.Y. (1971a). Hydromechanics of swimming propulsion. Part 1.
Swimming of a two-dimensional flexible plate at variable forward
speeds in an inviscid fluid. J. Fluid Mech. 46, 337-355.

Wu, T.Y. (1971b). Hydromechanics of swimming propulsion. Part 2.


Some optimum shape problems. J. Fluid Mech. 46, 521-544.

Wu, T.Y. (1971c). Hydromechanics of swimming propulsion. Part 3.


S w im m in g and optim um movements of slender fish with side fins.
J. Fluid Mech. 46,545-568.

Yates, G.T. (1986). Optimum pitching axes in flapping wing propulsion.


J. theor. Biol. 120, 255-276.

120

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Biography

Carlton E Heine

Born: August 15, 1961, Philadelphia, Pennsylvania.

Education:

Undergraduate: B.Sc. 1985, McGill University. Montreal, Quebec.


Graduate: Major - Zoology. Minor - Mechanical Engineering.
1986 - present, Duke University. Durham, N.C.

Awards:

Zoology Department Graduate Teaching Assistantship, (1986-1990)


Duke University.
Cocos Foundation Morphology Fellowship (1990-1992).
Graduate School Conference Travel Fellowship, (1990) and (1991)
Duke University.
NSF Japan Society for the Promotion of Science Postdoctoral
Fellowship, (1992).

Professional Societies:

American Society of Zoologists


Sigma Xi
Society for Industrial and Applied Mathematics

121

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
Publications:

Aerodynamics of gliding flight in a Harris Hawk, Parabuteo


unicinctus. Tucker, V.A. and Heine, C. (1990),
J. Exp. Biol. 149,469-489.

The profile drag of a hawk's wing, measured by wake sampling in


a wind tunnel. Pennycuick, C. J., Heine, C. and
Kirkpatrick, S.J. (1992), J. Exp. Biol, (in press).

Abstracts o f Presented Papers:

The kinematics of flapping ray locomotion. Heine, C. (1990), Am.


Zool. 30(4) 132A. Presented at American Society of
Zoologists, San Antonio, Texas, December 1990.

Mechanics of flapping ray locomotion. Heine, C. (1991), J. Mar.


Biol. Assoc. 71(3), p715. Presented at Mechanics and
Physiology of Animal Swimming, Plymouth, England,
March 1991.

Flapping ray locomotion. Heine, C. Presented at the American


Society of Ichthyologists and Herpetologists (American
Elasmobranch Society), New York, N.Y., June 1991.

The forces generated by a flapping ray. Heine, C. (1991), Am. Zool.


30(5) 474A. Presented at American Society of Zoologists,
Atlanta, Georgia, December 1991.

122

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

You might also like