03.01.2015 Views

Biomimetics and Marine Technology - Marine Technology Society

Biomimetics and Marine Technology - Marine Technology Society

Biomimetics and Marine Technology - Marine Technology Society

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Journal<br />

The International, Interdisciplinary <strong>Society</strong> Devoted to Ocean <strong>and</strong> <strong>Marine</strong> Engineering, Science, <strong>and</strong> Policy<br />

Volume 45 Number 4 July/August 2011<br />

<strong>Biomimetics</strong> <strong>and</strong> <strong>Marine</strong> <strong>Technology</strong>


INTRODUCING<br />

The Next Generation<br />

in Underwater Acoustic Modems<br />

The ATM-900 Series Modems<br />

from Teledyne Benthos<br />

• Advanced Data Logger Capabilities<br />

– Increased from 704K to 6MB (flash file format)<br />

– Searchable by record number <strong>and</strong> time<br />

• Dual Serial Port – Interface with more than one<br />

sensor at a time (optional)<br />

• Integrated real-time clock for data time<br />

stamping<br />

ATM-900 Series<br />

• New Comm<strong>and</strong> Structure & Modem Parameters<br />

• Pinger Mode<br />

• Arbitrary Waveform Generator (optional)<br />

Available April 2011<br />

See us at Ocean Business - Booth #S1.<br />

A Member of Teledyne <strong>Marine</strong><br />

49 Edgerton Drive, North Falmouth, MA 02556 USA<br />

T: +1 508.563.1000 • F: +1 508.563.6444<br />

E: benthos@teledyne.com<br />

www.benthos.com


Volume 45, Number 4, July/August 2011<br />

<strong>Biomimetics</strong> <strong>and</strong> <strong>Marine</strong> <strong>Technology</strong><br />

Guest Editors: Frank E. Fish <strong>and</strong> Donna M. Kocak<br />

Turn to page 7 for a key to the cover images.<br />

Text: SPi<br />

Cover <strong>and</strong> Graphics:<br />

Michele A. Danoff, Graphics By Design<br />

The <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal<br />

(ISSN 0025-3324) is published by the <strong>Marine</strong> <strong>Technology</strong><br />

<strong>Society</strong>, Inc., 5565 Sterrett Place, Suite 108, Columbia,<br />

MD 21044.<br />

MTS members can purchase the printed Journal for $27<br />

domestic <strong>and</strong> $50 (plus $50 S&H) international.<br />

Non-members <strong>and</strong> library subscriptions are $420 online only,<br />

$124 print—domestic, $140 (plus $50 S&H) print—<br />

international, $435 print <strong>and</strong> online (worldwide);<br />

Single-issue (hardcopy) is $20 plus $7.50 S&H (domestic),<br />

$24.50 S&H (international); Pay-per-view (worldwide):<br />

$15/article. Postage for periodicals is paid at Columbia, MD,<br />

<strong>and</strong> additional mailing offices.<br />

POSTMASTER:<br />

Please send address changes to:<br />

<strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal<br />

5565 Sterrett Place<br />

Suite 108<br />

Columbia, Maryl<strong>and</strong> 21044<br />

Copyright © 2011 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong>, Inc.<br />

In This Issue<br />

8<br />

<strong>Biomimetics</strong> <strong>and</strong> <strong>Marine</strong> <strong>Technology</strong>:<br />

An Introduction<br />

Frank E. Fish, Donna M. Kocak<br />

14<br />

Biomimicking <strong>Marine</strong> Mechanisms <strong>and</strong><br />

Organizational Principles<br />

Commentary by Yoseph Bar-Cohen<br />

16<br />

Sink <strong>and</strong> Swim: Clues From Nature for<br />

Aquatic Robotics<br />

Commentary by Jeannette Yen<br />

19<br />

Developing Bioinspired Autonomous<br />

Systems<br />

Commentary by Thomas M. McKenna<br />

24<br />

GhostSwimmer AUV: Applying<br />

<strong>Biomimetics</strong> to Underwater Robotics for<br />

Achievement of Tactical Relevance<br />

Commentary by Michael Rufo,<br />

Mark Smithers<br />

31<br />

Autonomous Robotic Fish as Mobile<br />

Sensor Platforms: Challenges <strong>and</strong><br />

Potential Solutions<br />

Xiaobo Tan<br />

41<br />

Robotic Models for Studying Undulatory<br />

Locomotion in Fishes<br />

George V. Lauder, Jeanette Lim,<br />

Ryan Shelton, Chuck Witt, Erik Anderson,<br />

James L. Tangorra<br />

56<br />

Thrust Production in Highly Flexible<br />

Pectoral Fins: A Computational Dissection<br />

Srinivas Ramakrishnan, Meliha Bozkurttas,<br />

Rajat Mittal, George V. Lauder<br />

65<br />

Learning From the Fins of Ray-Finned<br />

Fish for the Propulsors of Unmanned<br />

Undersea Vehicles<br />

James L. Tangorra, Timo Gericke,<br />

George V. Lauder<br />

74<br />

Bioinspired Design Process for an<br />

Underwater Flying <strong>and</strong> Hovering Vehicle<br />

Jason D. Geder, John S. Palmisano,<br />

Ravi Ramamurti, Marius Pruessner,<br />

Banahalli Ratna, William C. S<strong>and</strong>berg<br />

83<br />

A Twistable Ionic Polymer-Metal<br />

Composite Artificial Muscle for<br />

<strong>Marine</strong> Applications<br />

Kwang J. Kim, David Pugal,<br />

Kam K. Leang<br />

99<br />

Batoid Fishes: Inspiration for the Next<br />

Generation of Underwater Robots<br />

Keith W. Moored, Frank E. Fish,<br />

Trevor H. Kemp, Hilary Bart-Smith<br />

110<br />

Bioinspired Propulsion Mechanisms<br />

Based on Manta Ray Locomotion<br />

Keith W. Moored, Peter A. Dewey,<br />

Megan C. Leftwich, Hilary Bart-Smith,<br />

Alex<strong>and</strong>er J. Smits<br />

119<br />

Inspired by Sharks: A Biomimetic<br />

Skeleton for the Flapping, Propulsive<br />

Tail of an Aquatic Robot<br />

John H. Long, Jr., Tom Koob,<br />

Justin Schaefer, Adam Summers,<br />

Kurt Bantilan, Sindre Grotmol,<br />

Marianne Porter


In This Issue<br />

130<br />

Lateral-Line-Inspired Sensor Arrays for<br />

Navigation <strong>and</strong> Object Identification<br />

Vicente I. Fern<strong>and</strong>ez, Audrey Maertens,<br />

Frank M. Yaul, Jason Dahl,<br />

Jeffrey H. Lang, Michael S. Triantafyllou<br />

147<br />

A Conserved Neural Circuit-Based<br />

Architecture for Ambulatory <strong>and</strong><br />

Undulatory Biomimetic Robots<br />

Joseph Ayers, Anthony Westphal,<br />

Daniel Blustein<br />

153<br />

A Hybrid Class Underwater Vehicle:<br />

Bioinspired Propulsion, Embedded<br />

System, <strong>and</strong> Acoustic Communication<br />

<strong>and</strong> Localization System<br />

Michael Krieg, Peter Klein,<br />

Robert Hodgkinson, Kamran Mohseni<br />

165<br />

Modeling of Artificial Aurelia aurita<br />

Bell Deformation<br />

Keyur B. Joshi, Alex Villanueva,<br />

Colin F. Smith, Shashank Priya<br />

181<br />

Swimming <strong>and</strong> Walking of an<br />

Amphibious Robot With Fin Actuators<br />

Naomi Kato<br />

198<br />

<strong>Marine</strong> Applications of the Biomimetic<br />

Humpback Whale Flipper<br />

Frank E. Fish, Paul W. Weber,<br />

Mark M. Murray, Laurens E. Howle<br />

208<br />

Shark Skin Separation Control<br />

Mechanisms<br />

Amy Lang, Philip Motta,<br />

Maria Laura Habegger, Robert Hueter,<br />

Farhana Afroz<br />

216<br />

Can Biomimicry <strong>and</strong> Bioinspiration<br />

Provide Solutions for Fouling Control<br />

Emily Ralston, Geoffrey Swain<br />

228<br />

BOOK REVIEW: Sex, Drugs, <strong>and</strong> Sea<br />

Slime: The Oceans’ Oddest Creatures<br />

<strong>and</strong> Why They Matter<br />

by Ellen Prager<br />

Reviewed by Jason Goldberg


Deep Tracks!<br />

Teledyne RD Instruments’ new PAVS150 DVL keeps your<br />

vehicle on course up to 500 meters above the seafloor.<br />

The 150 kHz Phased Array Velocity Sensor (PAVS150)<br />

Doppler Velocity Log delivers over 2x the range of our<br />

industry-st<strong>and</strong>ard Workhorse Navigator. Now that’s<br />

deep tracks!<br />

The PAVS150’s 1000M depth rating, combined with its<br />

increased range allows you to submerge your manned<br />

or unmanned vehicle sooner <strong>and</strong> transit longer, saving<br />

you time, money, <strong>and</strong> peace of mind.<br />

Stay the course with PAVS!<br />

Full details at:<br />

http://www.rdinstruments.com/PAVS.aspx<br />

Poway, CA USA • Tel. +1-858-842-2600<br />

www.rdinstruments.com


Learn how MTS can help your company<br />

reach the technology leaders <strong>and</strong> innovators<br />

in the marine technology industry.<br />

Opportunities include advertising in Currents, the MTS Journal,<br />

sponsorships, online <strong>and</strong> custom-tailored programs.<br />

MTS —<br />

here for the marine technology community,<br />

here for you!<br />

Email Mary Beth Loutinsky at<br />

mbloutinsky@gmail.com or<br />

call 703-629-3810


<strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Officers<br />

BOARD OF DIRECTORS<br />

President<br />

Jerry Boatman<br />

QinetiQ North America – <strong>Technology</strong> Solutions<br />

Group<br />

President-elect<br />

Drew Michel<br />

ROV Technologies, Inc.<br />

Immediate Past President<br />

Elizabeth Corbin<br />

VP—Section Affairs<br />

Lisa Medeiros<br />

OceanWorks International<br />

VP—Education <strong>and</strong> Research<br />

Jill Z<strong>and</strong>e<br />

MATE Center<br />

VP—Industry <strong>and</strong> <strong>Technology</strong><br />

Jerry C. Wilson<br />

Fugro Pelagos, Inc.<br />

VP—Publications<br />

Karin Lynn<br />

Treasurer <strong>and</strong> VP—Budget <strong>and</strong> Finance<br />

Debra Kill<br />

International Submarine Engineering<br />

VP—Government <strong>and</strong> Public Affairs<br />

Justin Manley<br />

Liquid Robotics<br />

SECTIONS<br />

Canadian Maritime<br />

Vacant<br />

Florida<br />

Vacant<br />

Gulf Coast<br />

Laurie Jugan<br />

Consultant<br />

Hampton Roads<br />

Raymond Toll<br />

SAIC<br />

Hawaii<br />

Stewart Burley<br />

Strategic Theories Unlimited<br />

Houston<br />

Robert Keith<br />

Phoenix International Holdings, Inc.<br />

Japan<br />

Prof. Toshitsugu Sakou<br />

Tokai University<br />

Monterey<br />

Jill Z<strong>and</strong>e<br />

MATE Center<br />

New Engl<strong>and</strong><br />

Chris Jakubiak<br />

UMASS Dartmouth-SMAST<br />

Newfoundl<strong>and</strong> <strong>and</strong> Labrador<br />

Bill O’Keefe<br />

Surmount Technologies, Inc.<br />

Oregon<br />

TBD<br />

Puget Sound<br />

Fritz Stahr<br />

University of Washington<br />

San Diego<br />

Scott Mau<br />

Southwest Fisheries Sciences Center<br />

South Korea<br />

Dr. Seok Won Hong<br />

Maritime & Ocean Engineering Research Inst.<br />

(MOERI/KORDI)<br />

Washington, D.C.<br />

Brent Evers<br />

Hadal Technologies, Inc.<br />

PROFESSIONAL COMMITTEES<br />

Industry <strong>and</strong> <strong>Technology</strong><br />

Buoy <strong>Technology</strong><br />

Dr. Walter Paul<br />

Woods Hole Oceanographic Institution<br />

Cables <strong>and</strong> Connectors<br />

Helmut H. Portmann<br />

National Data Buoy Center<br />

Deepwater Field Development <strong>Technology</strong><br />

Dr. Benton Baugh<br />

Radoil, Inc.<br />

Diving<br />

David C. Berry<br />

Subsea Construction <strong>and</strong> Diving Consultant<br />

Dynamic Positioning<br />

Howard Shatto<br />

Shatto Engineering<br />

Manned Underwater Vehicles<br />

William Kohnen<br />

SEAmagine Hydrospace Corporation<br />

Moorings<br />

Jack Rowley<br />

SAIC<br />

Oceanographic Instrumentation<br />

Dr. Jim Irish<br />

University of New Hampshire<br />

Offshore Structures<br />

Dr. Peter W. Marshall<br />

MHP Systems Engineering<br />

Remotely Operated Vehicles<br />

Drew Michel<br />

ROV Technologies, Inc.<br />

Renewable Energy<br />

Rich Chwaszczewski<br />

SAIC<br />

Ropes <strong>and</strong> Tension Members<br />

Evan Zimmerman<br />

Delmar Systems, Inc.<br />

Seafloor Engineering<br />

Herb Herrmann<br />

Naval Seafloor Cable Protection Office<br />

Underwater Imaging<br />

Dr. Fraser Dalgleish<br />

Harbor Branch Oceanographic Institute<br />

Unmanned Maritime Vehicles<br />

Rafael M<strong>and</strong>ujano<br />

Vehicle Control Technologies, Inc.<br />

Education <strong>and</strong> Research<br />

<strong>Marine</strong> Archaeology<br />

Dan Warren<br />

C & C Technologies<br />

<strong>Marine</strong> Education<br />

Erica Moulton<br />

MATE Center<br />

<strong>Marine</strong> Geodetic Information Systems<br />

Dave Zilkoski<br />

NOAA<br />

<strong>Marine</strong> Materials<br />

Vacant<br />

Ocean Exploration<br />

Guillermo Söhnlein<br />

OceanGate<br />

Physical Oceanography/Meteorology<br />

Dr. Richard L. Crout<br />

National Data Buoy Center<br />

Remote Sensing<br />

Herb Ripley<br />

Hyperspectral Imaging Limited<br />

Government <strong>and</strong> Public Affairs<br />

<strong>Marine</strong> Law <strong>and</strong> Policy<br />

Montserrat Gorina-Ysern<br />

Healthy Children–Healthy Oceans Foundation<br />

<strong>Marine</strong> Mineral Resources<br />

Dr. John C. Wiltshire<br />

University of Hawaii<br />

<strong>Marine</strong> Security<br />

Dallas Meggitt<br />

Sound & Sea <strong>Technology</strong><br />

Ocean Economic Potential<br />

James Marsh<br />

University of Hawaii<br />

Ocean Observing Systems<br />

Donna Kocak<br />

HARRIS CapRock Communications<br />

Ocean Pollution<br />

Jacob Sobin<br />

NOAA Coastal Services Center<br />

STUDENT SECTIONS<br />

Duke University<br />

Counselor: Douglas Nowacek, Ph.D.<br />

Florida Atlantic University<br />

Counselor: Douglas A. Briggs, Ph.D.<br />

Florida Institute of <strong>Technology</strong><br />

Counselor: Stephen Wood, Ph.D., P.E.<br />

Long Beach City College<br />

Counselor: Scott Fraser<br />

Massachusetts Institute of <strong>Technology</strong><br />

Counselor: Alex<strong>and</strong>ra Techet, Ph.D.<br />

Monterey Peninsula College/Hartnell College<br />

Counselor: Jeremy R. Hertzberg<br />

Texas A&M University—College Station<br />

Counselor: Patrick Lynett<br />

Texas A&M—Corpus Christi<br />

Counselor: Lea-Der Chen, Ph.D.<br />

Texas A&M University—Galveston<br />

Counselor: Frank Warnakula, Ph.D.<br />

United States Naval Academy<br />

Counselors: Capt. Joseph T. Arcano<br />

(USN Ret), Ph.D.<br />

Cmdr. David J. Robillard<br />

University of Hawaii<br />

Counselor: R. Cengiz Ertekin, Ph.D.<br />

University of Houston<br />

Counselors: Raresh Pascali, P.E.,<br />

Chuck Richards<br />

University of North Carolina—Charlotte<br />

Counselor: James Conrad, Ph.D.<br />

University of Southern Mississippi<br />

Counselor: Stephen Howden, Ph.D.<br />

Webb Institute<br />

Counselor: Matthew Werner<br />

HONORARY MEMBERS<br />

†Robert B. Abel<br />

†Charles H. Bussmann<br />

John C. Calhoun, Jr.<br />

John P. Craven<br />

†Paul M. Fye<br />

David S. Potter<br />

†Athelstan Spilhaus<br />

†E. C. Stephan<br />

†Allyn C. Vine<br />

†James H. Wakelin, Jr.<br />

†deceased


Editorial Board<br />

Brian Bingham, Ph.D.<br />

Editor<br />

University of Hawaii at Manoa<br />

Corey Jaskolski<br />

Hydro Technologies<br />

Donna Kocak<br />

HARRIS CapRock Communications<br />

Scott Kraus, Ph.D.<br />

New Engl<strong>and</strong> Aquarium<br />

Dhugal Lindsay, Ph.D.<br />

Japan Agency for <strong>Marine</strong>-Earth Science<br />

& <strong>Technology</strong><br />

Justin Manley<br />

Liquid Robotics<br />

Stephanie Showalter<br />

National Sea Grant Law Center<br />

Jason Stanley<br />

Schilling Robotics<br />

Edith Widder, Ph.D.<br />

Ocean Research <strong>and</strong> Conservation<br />

Association<br />

Jill Z<strong>and</strong>e<br />

MATE Center<br />

Editorial<br />

Karin Lynn<br />

VP of Publications<br />

Brian Bingham, Ph.D.<br />

Editor<br />

Amy Morgante<br />

Managing Editor<br />

Administration<br />

Jerry Boatman<br />

President<br />

Richard Lawson<br />

Executive Director<br />

Jeanne Glover<br />

Membership <strong>and</strong> Marketing Manager<br />

Michael Hall<br />

Member Groups Manager<br />

Chris Barrett<br />

Director of Professional Development<br />

<strong>and</strong> Meetings<br />

Suzanne Voelker<br />

Subscription Manager<br />

The <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> is<br />

a not-for-profit, international professional<br />

society. Established in 1963, the <strong>Society</strong>’s<br />

mission is to promote the exchange of<br />

information in ocean <strong>and</strong> marine engineering,<br />

technology, science, <strong>and</strong> policy.<br />

Please send all correspondence to:<br />

The <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong><br />

5565 Sterrett Place, Suite 108<br />

Columbia, MD 21044<br />

(410) 884-5330 Tel.<br />

(410) 884-9060 FAX<br />

MTS Journal: morganteeditorial@verizon.net<br />

Publications: publications@mtsociety.org<br />

Membership: Jeanne.Glover@mtsociety.org<br />

Programs: Michael.Hall@mtsociety.org<br />

Director: Rich.Lawson@mtsociety.org<br />

Online: www.mtsociety.org<br />

MEMBERSHIP INFORMATION<br />

may be obtained by contacting the <strong>Marine</strong><br />

<strong>Technology</strong> <strong>Society</strong>. Benefits include:<br />

■ Free subscription to the online <strong>Marine</strong><br />

<strong>Technology</strong> <strong>Society</strong> Journal, with highly<br />

reduced rates for the paper version<br />

■ Free subscription to the bimonthly news -<br />

letter, Currents, covering events, business<br />

news, science <strong>and</strong> technology, <strong>and</strong><br />

people in marine technology<br />

■ Member discounts on all MTS publications<br />

■ Reduced registration rates to all MTS <strong>and</strong><br />

MTS-sponsored conferences <strong>and</strong> workshops<br />

■ Member-only access to an expansive Job<br />

Bank <strong>and</strong> Member Directory<br />

■ Reduced advertising rates in MTS<br />

publications<br />

■ National recognition through our Awards<br />

Program<br />

Individual dues are $75 per year. Life<br />

membership is available for a one-time<br />

fee of $1,000. Patron, Student, Emeritus,<br />

Institutional, Business, <strong>and</strong> Corporate<br />

memberships are also available.<br />

ADVERTISING<br />

Advertising is accepted by the <strong>Marine</strong><br />

Tech nology <strong>Society</strong> Journal. For more information<br />

on MTS advertising <strong>and</strong> policy,<br />

please contact Mary Beth Loutinsky,<br />

mbloutinsky@att.net.<br />

COPYRIGHT<br />

Copyright © 2011 by the <strong>Marine</strong> <strong>Technology</strong><br />

<strong>Society</strong>, Inc. Authorization to photocopy<br />

items for internal or personal use, or the<br />

internal or personal use of specific clients, is<br />

granted by the <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong>,<br />

provided that the base fee of $1.00 per copy,<br />

plus .20 per page is paid directly to Copyright<br />

Clearance Center, 222 Rosewood Dr.,<br />

Danvers, MA 01923.<br />

For those organizations that have been<br />

granted a photocopy license by CCC, a separate<br />

sys tem of payment has been arranged.<br />

The fee code for users of the Transactional<br />

Reporting Service is 0025-3324/89 $1.00 +<br />

.20. Papers by U.S Government employees<br />

are declared works of the U.S. Government<br />

<strong>and</strong> are therefore in the public domain.<br />

The <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> cannot be<br />

held responsible for the opinions given <strong>and</strong><br />

the statements made in any of the articles<br />

published.<br />

ABSTRACTS<br />

MTS Journal article abstracts, if available,<br />

can be accessed for free at http://www.<br />

ingentaconnect.com/content/mts/mtsj.<br />

Print <strong>and</strong> electronic abstracts of MTS Journal<br />

articles are also available through GeoRef , Aquatic Sciences <strong>and</strong><br />

Fisheries Abstracts, published by Cambridge<br />

Scientific Abstracts , <strong>and</strong> Geobase’s<br />

Oceanbase published by Elsevier Science.<br />

CONTRIBUTORS<br />

Contributors can obtain an information <strong>and</strong><br />

style sheet by contacting the managing editor.<br />

Sub missions that are relevant to the concerns<br />

of the <strong>Society</strong> are welcome. All papers are subjected<br />

to a stringent review procedure directed<br />

by the editor <strong>and</strong> the editorial board. The<br />

Journal focuses on technical material that may<br />

not otherwise be available, <strong>and</strong> thus technical<br />

papers <strong>and</strong> notes that have not been published<br />

previously are given priority. General commentaries<br />

are also accepted <strong>and</strong> are subject to review<br />

<strong>and</strong> approval by the editorial board.<br />

6 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Key to Cover Images<br />

Front cover: GhostSwimmer AUVs from Boston<br />

Engineering Advanced Systems Group <strong>and</strong> Olin College<br />

Intelligent Vehicle Lab. Image courtesy of HARRIS CapRock<br />

Communication; underwater photo courtesy of Dr. Tamara Frank.<br />

Back cover: (Image courtesy of HARRIS CapRock Communications)<br />

1. Propeller with tubercles (propeller courtesy of Laurens Howls<br />

<strong>and</strong> morphing courtesy of John A. Lever)<br />

2. CephaloBot prototype hybrid vehicle<br />

3. Robotic Turtle, “RT-I”<br />

4. Mantabot<br />

5. Artificial Aurelia aurita bell deformation (photo courtesy of<br />

Alex Villanueva, CIMSS, Virginia Tech)<br />

6. Autonomous robotic fish (photo courtesy of Xiaobo Tan)<br />

7. Bluegill Sunfish (Lepomis macrochirus) with robotic pectoral<br />

fins (photo courtesy of George Lauder <strong>and</strong> James Tangorra)<br />

8. Four-fin vehicle based on Bird wrasse (Gomphosus varius)<br />

9. Tadro4 modeled after living electric ray Narcine<br />

(photo courtesy of Dr. Steve Kajiura)<br />

10. Lobster-based robot (robot photo courtesy of Brian Tucker,<br />

Bresnahan Photography; lobster photo courtesy of<br />

Dr. Tamara Frank)


INTRODUCTION<br />

<strong>Biomimetics</strong> <strong>and</strong> <strong>Marine</strong> <strong>Technology</strong>:<br />

An Introduction<br />

Frank E. Fish<br />

West Chester University<br />

Donna M. Kocak<br />

HARRIS CapRock Communications<br />

I<br />

nspiration for the development of new technologies is at the heart of the biomimetic<br />

approach. As there is a wide diversity of biological forms, particular attributes can be targeted<br />

that provide innovative solutions to engineering problems. Biology can provide new<br />

technological possibilities <strong>and</strong> enhance performance of existing technologies. Biomimicry is<br />

a tool for solving problems in the conceptual <strong>and</strong> embodiment phases of design (Reap<br />

et al., 2005). The goal of biomimetics is to use biological inspiration to engineer machines<br />

that emulate the performance of animals (Kumph & Triantafyllou, 1998; Taubes, 2000),<br />

particularly in instances where the animal’s performance exceeds current technology. The<br />

natural experimentation that has occurred through the evolutionary process has produced<br />

the plethora of organisms, both living <strong>and</strong> extinct. Within the phylogenetic lineages of<br />

these organisms, there has been essentially a “cost-benefit analysis” where particular designs<br />

for specific functions have been optimized to perform with respect to the rigors of their<br />

environment (Fish, 2006).<br />

Biologists are well acquainted with the specific adaptations present in animals, which may<br />

be of interest to engineers. For biologists, an adaptationist program has allowed for the identification<br />

of novel features of organisms based on engineering principles, whereas for engineers,<br />

identification of such novel features is necessary to exploit them for biomimetic<br />

development. This new synergy between biologists <strong>and</strong> engineers can be beneficial in advancing<br />

technology by looking to nature to provide solutions to current problems.<br />

For marine technologies, the biomimetic approach particularly holds the promise of enhanced<br />

performance <strong>and</strong> increased efficiency for operation in the aquatic realm. It was in the<br />

oceans that life first evolved <strong>and</strong> where complex animals have thrived for over 600 million<br />

years. There are marine representatives from every major animal phylum. <strong>Marine</strong> animals<br />

survive in environments as diverse as tropical coral reefs, polar ice-capped oceans, <strong>and</strong> the<br />

lightless abyssal depths. The diversity of habitats available in marine systems has led to a<br />

vast array of body designs <strong>and</strong> physiological <strong>and</strong> behavioral mechanisms. These adaptations<br />

that evolved in animals are used to overcome the biotic <strong>and</strong> abiotic challenges in the ocean.<br />

To deal with the rigors of the marine environment, animals have developed specialized<br />

sensory systems (e.g., echolocation, electroreception), mechanisms to deal with pressure<br />

8 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


(e.g., buoyancy control), strategies to economize on energy (e.g., fusiform body design,<br />

schooling, burst-<strong>and</strong>-glide swimming), armor (e.g., bony scales, mollusk shells), stability<br />

mechanisms (e.g., paired <strong>and</strong> median fins), maneuverability (e.g., flexible bodies, vectored<br />

thrust), speed (e.g., high-aspect-ratio oscillatory propulsors, jet propulsion), stealth (e.g.,<br />

camouflage, low acoustic signature), <strong>and</strong> use of compliant materials (e.g., collagen, protein<br />

rubbers, mucous). In using such specializations to enhance their own survival, animals attempt<br />

to function in a manner to minimize their total energy budget while maximizing the<br />

performance of the specialization. Animals are doing the type of optimization that engineers<br />

seek to incorporate into designs (Vincent, 1990), <strong>and</strong> specifically, marine animals<br />

are dealing with the very problems that are of concern for marine engineers.<br />

<strong>Marine</strong> technology is well suited for the application of bioinspired design, as there is a need<br />

for exploitation of the oceans for new sources of food, energy, <strong>and</strong> minerals. Exploration of<br />

the world’s oceans is expensive. Ship time, support personnel, maintenance facilities, <strong>and</strong><br />

energy costs can be prohibitively costly. Added to these costs are the expanse of the ocean<br />

surface (3.6 × 10 8 km 2 ) to be explored <strong>and</strong> the dangers associated with working in the<br />

deep-water environment (average depth = 3,650 m). The Challenger Deep (depth =<br />

10,902 m; pressure = 16,500 psi or 113,764 kPa) in the Mariana Trench was only visited<br />

once by a manned vehicle, Trieste, in 1960. Since that time, only two unmanned expeditions,<br />

robotic deep-sea probe Kaikō <strong>and</strong> HROV Nereus, have returned to the deepest surveyed point<br />

in the ocean.<br />

Historically, marine animals have served as the inspiration for technological design. During<br />

the Renaissance, animals were identified as streamlined bodies for drag reduction that<br />

could be applied to manufactured devices. Between 1505 <strong>and</strong> 1508, Leonardo da Vinci<br />

was particularly interested in flow in water, as revealed in his notebooks, Codex Leicester<br />

(Ball, 2009). Da Vinci wrote on the function of streamlined bodies in reducing drag <strong>and</strong><br />

noted the streamlined shape of a fish (Anderson, 1998). He argued that the fish could<br />

move through the water with little resistance because its shape allowed the water to flow<br />

smoothly over the afterbody without prematurely separating. Da Vinci recognized <strong>and</strong> demonstrated<br />

a similar design with the hull shape of ships.<br />

Giovanni Borelli in 1680 made an examination of the swimming motions of animals with<br />

their application to submarine technology (Borelli, 1680). In his book De Motu Animalium<br />

(The Movement of Animals), Borelli likened swimming to flying in that both were accomplished<br />

by the displacement of fluids, although he noted the differences in density of air <strong>and</strong><br />

water <strong>and</strong> their effects on stability <strong>and</strong> buoyancy. Borelli described the design of an early submarine<br />

that incorporated ideas based in part on animals for buoyancy regulation <strong>and</strong> propulsion.<br />

The submarine would submerge using a hydrostatic mechanism based on the swim<br />

bladder of a fish by filling goatskin bags, located inside the submarine, through holes in<br />

July/August 2011 Volume 45 Number 4 9


the sides of the boat. Propulsion would be accomplished by oars projecting through the hull<br />

<strong>and</strong> fitted with watertight seals. When the submarine was on the bottom, it was envisioned<br />

that the oars would push off the s<strong>and</strong>y substrate to move the boat along. In mid-water, the<br />

oars would paddle like the feet of frogs or geese. During the rearward power stroke, a flexible<br />

paddle at the end of the oar would exp<strong>and</strong> to work on a large mass of fluid. During the forward<br />

recovery stroke, the paddle would fold passively to reduce the frontal area <strong>and</strong> drag on the oar.<br />

However, Borelli considered that propulsion of the boat would be easier if a flexible oar were<br />

positioned at the stern, emulating the motion of a fish tail. Despite the elaborate design for its<br />

time, it is doubtful if this early biomimetic experiment was successfully used.<br />

Cayley examined the streamlined body shapes of a trout <strong>and</strong> a dolphin in 1809 as solids of<br />

least resistance design (Gibbs-Smith, 1962). Cayley unsuccessfully attempted to apply these<br />

natural designs to the hull of a boat for moving on the water surface (Vogel, 1998). The<br />

rounded configuration of the hull was unstable with respect to roll, <strong>and</strong> low drag did not<br />

occur. The design of fish <strong>and</strong> dolphins is similar to the optimal shape for drag reduction<br />

of submerged bodies, such as modern submarines. These natural swimmers <strong>and</strong> submarines<br />

have fusiform body shapes with a rounded leading edge <strong>and</strong> slowly tapering tail. The forerunner<br />

for hulls used by modern nuclear submarines, the USS Albacore, was built in 1953<br />

with a fusiform shape.<br />

Both Cayley’s hull <strong>and</strong> the USS Albacore demonstrate limitations due to a misuse of the<br />

biomimetic approach. For Cayley, strict adherence to copying biological designs without<br />

proper insight into the function <strong>and</strong> limitations of those designs proved disastrous (Vogel,<br />

1998; Fish, 2006). While the shapes of fish <strong>and</strong> dolphins are appropriate for movement<br />

underwater, their shapes are not effective at the water surface. Hulls with broad beams <strong>and</strong><br />

greater buoyancy, like those displayed by waterfowl, provide enhanced stability <strong>and</strong> opportunity<br />

for greater speed at the water surface (Aigeldinger & Fish, 1995).<br />

InthecaseoftheUSSAlbacore, the design is only analogous with fish <strong>and</strong> dolphins.<br />

Although the designs are convergent, there was no information exchange to determine design.<br />

The Albacore was likened to the shape of a fish, but the submarine’s design was not biologically<br />

inspired (Harris 1997; Largess & M<strong>and</strong>elblatt, 1999). This was similar to the description<br />

of the fictional submarine, the Nautilus, in Jules Verne’s Twenty Thous<strong>and</strong> Leagues under<br />

the Sea:<br />

We were lying upon the back of a sort of submarine boat, which appeared (as far as I could<br />

judge) like a huge fish of steel.<br />

The hull shape of the Albacore was based on the “Lyon form” of airship models (Largess &<br />

M<strong>and</strong>elblatt, 1999). Originally, submarine hulls were designed more as surface ships due to<br />

10 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


the limited amount of time that they could operate submerged. The streamlined hull of the<br />

Albacore made it the fastest <strong>and</strong> most maneuverable submarine of its time.<br />

The convergent designs of the Albacore <strong>and</strong> marine animals reflect selection, both artificial<br />

<strong>and</strong> natural, respectively, for optimizing similar performance parameters (i.e., speed, drag reduction,<br />

maneuverability). However, similarity of design does not necessarily always translate<br />

into identical performance <strong>and</strong> assumptions of performance expectations should be approached<br />

with caution. Similar shapes can have different functions or have limitations to a<br />

function when viewed in isolation without consideration of the whole system. Despite its<br />

enhanced maneuverability compared to other submarines, the Albacore’s rateofturnat<br />

2° s −1 is poor when compared with similarly shaped dolphins, which can turn at 453° s −1<br />

<strong>and</strong> in a confined space of 20% of body length (Fish, 2002). The rigid hull of the submarine<br />

in concert with yaw control mainly from an aft-positioned rudder limits agility <strong>and</strong> maneuverability<br />

compared to flexible-bodied dolphins with multiple fore <strong>and</strong> aft mobile control<br />

surfaces (Fish, 2002; Fish & Nicastro, 2003).<br />

The biomimetic approach dem<strong>and</strong>s first careful observation of the whole biological system<br />

to identify the principles <strong>and</strong> attributes of the system. Thus, major limitations <strong>and</strong> constraints<br />

of any biological design can be defined before translation to an engineered system.<br />

The association between biologists <strong>and</strong> engineers becomes paramount as biomimetic technologies<br />

are developed.<br />

This special issue of the <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal brings together both biologists<br />

<strong>and</strong> engineers who are currently involved with the development of biomimetic devices for<br />

applications in the marine environment. The intent is to assess the current state of technology<br />

<strong>and</strong> incite new applications that apply these <strong>and</strong> other innovative concepts as technology advances.<br />

The research presented in this issue mimics specific aspects of fish, skates, rays, sharks,<br />

lobsters, jellyfish, squids, <strong>and</strong> sea turtles. In many of these papers, the goal is focused on locomotion<br />

or propulsion in robotic counterparts so they can maneuver more efficiently in the<br />

animal’s designed-for environment. This may be acted out either individually or in multiples<br />

as schools (or swarms) of fish. Swimming <strong>and</strong> walking using amphibious designs are also considered.<br />

A recent student competition sponsored by the Office of Naval Research (ONR) was<br />

held for the first time (involving several of the authors in this issue) to evaluate the performance<br />

of “mantabots,” which aspire to perform as gracefully as the sleek <strong>and</strong> elegant manta<br />

rays they attempt to copy (Pennisi, 2011). A commentary by McKenna provides an overview<br />

of this <strong>and</strong> other biomimetic work sponsored by ONR. Other papers in this issue identify a<br />

single unique aspect <strong>and</strong> strive to achieve the same benefit in an engineered device. Simulating<br />

the lateral line system found in most aquatic vertebrates using pressure sensors is one example<br />

that can be used to supplement vision <strong>and</strong> sonar in turbid waters. Deriving a control<br />

mechanism based on shark skin properties to increase vehicle swimming speeds, imitating the<br />

July/August 2011 Volume 45 Number 4 11


tubercles of whale fins to enhance hydrodynamics, <strong>and</strong> engineering a bioinspired solution<br />

for natural antifouling mechanisms are other examples. Whether your interest is in underwater<br />

vehicles, imaging, dynamic positioning, marine materials, ocean pollution, remote<br />

sensing, oceanographic instrumentation, ocean observing, marine science, or education,<br />

this special issue should provide valuable content.<br />

References<br />

Aigeldinger, T.L., & Fish, F.E. 1995. Hydroplaning by ducklings: Overcoming limitations to swimming at the<br />

water surface. J Exp Biol. 198:1567-4.<br />

Anderson, J.D. 1998. A History of Aerodynamics. Cambridge: Cambridge University Press.<br />

Ball, P. 2009. Flow. Oxford: Oxford University Press.<br />

Borelli, G.A. 1680. De Motu Animalium Pars (The Movement of Animals, translated by P. Maquet (1989)).<br />

Berlin: Springer-Verlag.<br />

Fish, F.E. 2002. Balancing requirements for stability <strong>and</strong> maneuverability in cetaceans. Integr Comp Biol.<br />

42:85-93. doi: 10.1093/icb/42.1.85.<br />

Fish, F.E. 2006. Limits of nature <strong>and</strong> advances of technology in marine systems: What does biomimetics have<br />

to offer to aquatic robots Appl Bionics Biomech. 3:49-60. doi: 10.1533/abbi.2004.0028.<br />

Fish, F.E., & Nicastro, A.J. 2003. Aquatic turning performance by the whirligig beetle: constraints on<br />

maneuverability by a rigid biological system. J Exp Biol. 206:1649-56. doi: 10.1242/jeb.00305.<br />

Gibbs-Smith, C.H. 1962. Sir George Cayley’s Aeronautics 1796-1855. London: Her Majesty’s Stationery<br />

Office.<br />

Harris, B. 1997. The Navy Times Book of Submarines: A Political, Social, <strong>and</strong> Military History. New York:<br />

Berkley.<br />

Kumph, J.M., & Triantafyllou, M.S. 1998. A fast-starting <strong>and</strong> maneuvering vehicle, the ROBOPIKE.<br />

In: Proceedings of the International Symposium on Seawater Drag Reduction, ed. Meng, J.C.S.,<br />

pp. 485-90. Newport, Rhode Isl<strong>and</strong>.<br />

Largess, R.P., & M<strong>and</strong>elblatt, J.L. 1999. U.S.S. Albacore: Forerunner of the Future. Portsmouth, NH:<br />

The Portsmouth <strong>Marine</strong> <strong>Society</strong>.<br />

Pennisi, E. 2011. Manta Machines, 332:28-9, Science, 27 May 2001, www.sciencemag.org. Retrieved on<br />

June 14, 2011.<br />

Reap, J., Baumeister, D., & Bras, B. 2005. Holism, biomimicry <strong>and</strong> sustainable engineering. In: Proceedings<br />

of IMECE2005. November 5-11, 2005. Orl<strong>and</strong>o, FL.<br />

12 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Taubes, G. 2000. Biologists <strong>and</strong> engineers create a new generation of robots that imitate life. Science. 288:80-3.<br />

doi: 10.1126/science.288.5463.80.<br />

Vincent, J. 1990. Structural Biomaterials. Princeton: Princeton Univ. Press.<br />

Vogel, S. 1998. Cat’s Paws <strong>and</strong> Catapults. New York: W. W. Norton.<br />

July/August 2011 Volume 45 Number 4 13


COMMENTARY<br />

Biomimicking <strong>Marine</strong> Mechanisms<br />

<strong>and</strong> Organizational Principles<br />

AUTHOR<br />

Yoseph Bar-Cohen<br />

Jet Propulsion Laboratory,<br />

California Institute of <strong>Technology</strong><br />

Nature is effectively a giant laboratory<br />

where trial-<strong>and</strong>-error evolutionary<br />

experiments are taking place. As<br />

nature performs its experiments, all<br />

the fields of science <strong>and</strong> engineering<br />

are employed, including physics,<br />

chemistry, mechanical engineering,<br />

<strong>and</strong> materials science. The processes<br />

range in scale from nano <strong>and</strong> micro<br />

(e.g., viruses <strong>and</strong> bacteria) to macro<br />

<strong>and</strong> mega (e.g., our life scale, elephants,<br />

<strong>and</strong> whales). To address the<br />

many survival challenges, biological<br />

systems came up with superb solutions.<br />

The constraints in addressing<br />

these challenges are similar to those<br />

that human engineers are facing, including<br />

the need to maximize the<br />

functionality of their design <strong>and</strong> produce<br />

systems that use minimal resources<br />

(e.g., materials, energy, cost,<br />

etc.). Humans have always made efforts<br />

to use nature as a model for<br />

inspiring innovation <strong>and</strong> problem<br />

solving. However, biological <strong>and</strong><br />

botanical systems have superior<br />

capabilities, including producing<br />

materials—they use their body temperature,<br />

the materials are recyclable,<br />

<strong>and</strong> the process does not involve<br />

pollution.<br />

To take advantage of the capabilities<br />

of nature, the field of biomimetics<br />

involves seeking to underst<strong>and</strong> <strong>and</strong><br />

use of the capabilities as a model for<br />

copying, adapting, <strong>and</strong> inspiring concepts<br />

<strong>and</strong> designs (Bar-Cohen, 2005;<br />

Bar-Cohen, 2011; Benyus, 1998;<br />

Vincent, 2001). For this purpose,<br />

scientists are seeking rules, concepts,<br />

mechanisms, <strong>and</strong> principles to inspire<br />

new possibilities. Some of the benefits<br />

that resulted from biomimetic approaches<br />

have led to improved structures,<br />

actuators, sensors, interfaces,<br />

control algorithms, software, drugs,<br />

defense, <strong>and</strong> intelligence, <strong>and</strong> may<br />

help to improve our ability to recycle<br />

materials <strong>and</strong> protect the environment.<br />

Some of the biomimetic characteristics<br />

that are being developed include shape<br />

morphing, self-repair, self-replication,<br />

<strong>and</strong> self-reconfiguration (Bar-Cohen<br />

& Breazeal, 2003; Bar-Cohen, 2011).<br />

Increasingly, researchers are working<br />

towards adapting the capabilities of<br />

many creatures to perform tasks in<br />

hard-to-reach areas <strong>and</strong> in conditions<br />

that are too harsh or dangerous for<br />

humans.<br />

<strong>Marine</strong> biosystems are quite rich<br />

in capabilities that have <strong>and</strong> can further<br />

benefit humans from biomimicking.<br />

There are many examples that<br />

one can list, including one as simple<br />

as the fins that are used by swimmers<br />

<strong>and</strong> divers that significantly enhance<br />

their performance. While it may be<br />

arguable that the fins were a biologically<br />

inspired invention, one can state<br />

that it is common knowledge that<br />

swimming creatures (e.g., geese,<br />

swans, seagulls, seals, <strong>and</strong> frogs) have<br />

feet with membranes that help them<br />

swim. The stability, maneuvering,<br />

<strong>and</strong> swimming performance of underwater<br />

animals are determined by the<br />

morphology, position, <strong>and</strong> mobility<br />

of their control surfaces. For cetaceans<br />

(i.e., whales, dolphins, <strong>and</strong> porpoises)<br />

the pectoral flippers are mobile hydrofoils<br />

that generate lift similar to engineered<br />

hydrofoils. The flippers have<br />

various shapes <strong>and</strong> are used to perform<br />

lateral turning, dive, surface,<br />

brake, <strong>and</strong> other mobility-related<br />

functions. Studies of the 3-D geometry<br />

<strong>and</strong> hydrodynamic performance<br />

of cetacean flippers with various<br />

morphologies help provide insight<br />

into the maneuverability, drag <strong>and</strong><br />

lift performance at high Reynolds<br />

numbers (Fish et al., 2011). Wagging<br />

the body <strong>and</strong> tail is the main propulsion<br />

method of marine swimmers<br />

(e.g., billfish <strong>and</strong> sailfish), allowing<br />

them to reach significant speeds of<br />

over 75 km/h. Submarines that<br />

could perform efficiently as marine<br />

swimmers using a flexible body<br />

would be an important exploration<br />

tool for scientists <strong>and</strong> would potentially<br />

have many military applications.<br />

While significant advances were<br />

reached via mimicking <strong>and</strong> the inspiration<br />

of marine biology, there are many<br />

capabilities in nature that are still far<br />

superior to engineered capabilities;<br />

examples include the following:<br />

■ The sonar of marine animals is far<br />

superior to any existing marine<br />

sonar (Muller & Hallam, 2004).<br />

■ Sea shells <strong>and</strong> skeletons of marine<br />

invertebratesarefarstronger<strong>and</strong><br />

lighter than human-made materials,<br />

<strong>and</strong> their fabrication does not create<br />

pollution concerns.<br />

14 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


■ Muscles stick to rocks even though<br />

the adhesion is done in water, <strong>and</strong><br />

they sustain sticksion in spite of<br />

the strong impacts of ocean waves.<br />

On the other h<strong>and</strong>, most humanmade<br />

adhesives fail when the adhesion<br />

is done in on a wet surface.<br />

■ The chiton, which is a diminutive<br />

mollusk, has very strong teeth that<br />

it uses to munch on rocks <strong>and</strong> extract<br />

food (Weaver et al., 2010).<br />

This capability may be used to inspire<br />

the development of effective<br />

lightweight bits for in situ planetary<br />

exploration sampling drills<br />

(Figure 1).<br />

One hopes that engineers will be<br />

able to rapidly prototype biological<br />

capabilities as fast as it is now possible<br />

to graphically edit photos of biological<br />

systems, as illustrated in Figure 2. In<br />

this figure, a photo of a shark was edited<br />

to create an imaginary image of a<br />

U.S. Navy vessel that is shark-like.<br />

While we are somewhat far from this<br />

capability, significant advances have<br />

been made by scientists <strong>and</strong> engineers<br />

whoseektomimicmarinebiology.<br />

This special issue is dedicated to describing<br />

<strong>and</strong> discussing the latest advances<br />

that were inspired by marine<br />

mechanisms <strong>and</strong> their organizational<br />

principles.<br />

FIGURE 2<br />

An example is shown where software was<br />

used to turn a photo of a natural shark into<br />

a vessel-like naval system. The photograph<br />

(a) was taken by the author; the image<br />

below (b) shows the modification to a Navy<br />

marine vehicle <strong>and</strong> is courtesy of David<br />

Hanson, Hanson Robotics LLC, TX.<br />

Acknowledgment<br />

Some of the research reported in<br />

this article was conducted at the Jet<br />

Propulsion Laboratory, California Institute<br />

of <strong>Technology</strong>, under a contract<br />

with the National Aeronautics<br />

<strong>and</strong> Space Administration.<br />

Author:<br />

Yoseph Bar-Cohen<br />

Jet Propulsion Laboratory,<br />

California Institute of <strong>Technology</strong><br />

4800 Oak Grove Drive,<br />

Pasadena, CA 91109-8099<br />

Email: yosi@jpl.nasa.gov<br />

Benyus, J.M. 1998. Biomimicry: Innovation<br />

Inspired by Nature. New York, NY: Harper-<br />

Collins Publishers. pp. 1-302.<br />

Fish, F.E., Smits, A.J., Haj-Hariri, H.,<br />

Iwasaki, T., & Bart-Smith, H. 2011.<br />

Biomimetic swimmer inspired by the manta<br />

ray, Chapter 17. In <strong>Biomimetics</strong>: Nature-<br />

Based Innovation, ed. Bar-Cohen, Y. Boca<br />

Raton, FL: CRC Press, Taylor & Francis<br />

Group.<br />

Muller, R., & Hallam, J.C.T. 2004. From<br />

bat pinnae to sonar antennae: augmented<br />

obliquely truncated horns as a novel parametric<br />

shape model. In: Proceeding of the 8th<br />

International Conference on the Simulation<br />

of Adaptive Behavior, SAB’04. Massachusetts,<br />

USA: The International <strong>Society</strong> for Adaptive<br />

Behavior.<br />

Vincent, J.F.V. 2001. Stealing ideas from<br />

nature, Chapter 3. In: Deployable Structures,<br />

ed. Pellegrino, S. Vienna, Austria: Springer-<br />

Verlag. pp. 51-58.<br />

Weaver, J.C., Wang, Q., Miserez, A.,<br />

Tantuccio, A., Stromberg, R., Bozhilov,<br />

K.N.P., … Kisailus, D. 2010. Analysis of an<br />

ultra hard magnetic biomineral in chiton<br />

radular teeth. Mater Today, 13(1-2):<br />

pp. 42-52.<br />

FIGURE 1<br />

A front view of the Eudoxochiton nobilis<br />

(“Noble” chiton). Courtesy of Iain Anderson.<br />

References<br />

Bar-Cohen, Y. (Ed.). 2005. <strong>Biomimetics</strong>—<br />

Biologically Inspired Technologies. Boca<br />

Raton, FL: CRC Press. pp. 1-527.<br />

Bar-Cohen, Y. (Ed.). 2011. <strong>Biomimetics</strong>:<br />

Nature-Based Innovation. Boca Raton, FL:<br />

CRC Press, Taylor & Francis Group.<br />

pp. 1-788.<br />

Bar-Cohen, Y., & Breazeal, C. (Eds.). 2003.<br />

Biologically Inspired Intelligent Robots.<br />

Bellingham, WA: SPIE Press. pp. 1-393.<br />

Vol. PM122.<br />

July/August 2011 Volume 45 Number 4 15


COMMENTARY<br />

Sink <strong>and</strong> Swim: Clues From Nature<br />

for Aquatic Robotics<br />

AUTHOR<br />

Jeannette Yen<br />

Georgia Institute of <strong>Technology</strong><br />

There are many ways to travel by<br />

water. Watercraft can use propellers<br />

<strong>and</strong> sails while living organisms exhibit<br />

quite a variety of modes of transport.<br />

They sink <strong>and</strong> swim, flap <strong>and</strong><br />

glide, stroke <strong>and</strong> jet. From this, we<br />

see a distinction between the limited<br />

human solutions <strong>and</strong> the diverse natural<br />

solutions. This natural diversity<br />

continues to inspire inventions in robotics.<br />

Leonardo da Vinci envisioned<br />

walking on water (Figure 1), something<br />

water striders could do millions<br />

of years ago (Hu et al., 2007). Bathtub<br />

toys such as the TwiddleFish,<br />

designed by Chuck Pell of Duke<br />

University (Figure 2), taught us the<br />

importance of stiffness in how fish<br />

FIGURE 1<br />

Leonardo da Vinci: Shoes for walking on water<br />

(image from http://en.wikipedia.org/wiki/<br />

Walking_on_water).<br />

swim so well. Pell commented, “In<br />

the h<strong>and</strong>s of kids…they can really<br />

feel what’s goingon.” To the delight<br />

of pre-K–12 children <strong>and</strong> their parents,<br />

there are propulsive toys on the<br />

market with remarkably functional<br />

body parts, serving as inspiration <strong>and</strong><br />

outreach to youngsters.<br />

Recent efforts have focused on<br />

flapping by fish <strong>and</strong> mantas (e.g.,<br />

Tangorra et al., 2011; Fish et al.,<br />

2011a) or jetting by jellyfish <strong>and</strong><br />

squid (e.g., Moslemi & Krueger,<br />

2011). By replicating nature via robotics,<br />

we underst<strong>and</strong> the significance<br />

of the number of joints (Dean et al.,<br />

2009), the shape of the fin (Curet<br />

et al., 2011), <strong>and</strong> the direction of a<br />

tail swish (Long et al., 2006) for controlling<br />

movement. Using these robots<br />

to repeatedly vary the timing of<br />

undulations of a fish or jellyfish, we<br />

discover that the frequency of the<br />

real organism is optimized to capture<br />

FIGURE 2<br />

TwiddleFish by Charles Pell, Bio-Design Studio,<br />

Duke University (cited by Guterl, 1996). Photo<br />

courtesy of F. Fish.<br />

the energy shed in the vortices left by<br />

the previous pulse (Triantafyllou &<br />

Triantafyllou, 1995; Fish, 2006;<br />

Ruiz et al., 2010). This is not something<br />

you could ask a fish or a jellyfish<br />

to do over <strong>and</strong> over again (though as<br />

biologists, we have patiently waited<br />

<strong>and</strong> recorded our aquatic creatures<br />

performing these behaviors over <strong>and</strong><br />

over <strong>and</strong> over again; see Catton et al.,<br />

2011). We identify the body parts that<br />

are important for propulsion, <strong>and</strong> we<br />

figure out how many propulsors are<br />

needed, along with their placement.<br />

Control systems coordinate the multiple<br />

fins. Sensors integrating different<br />

sensory modalities are sought to develop<br />

navigation systems that reliably<br />

guide the robot to reach its destination.<br />

We test out new actuators that<br />

accelerate unsteadily to increase thrust.<br />

We use new materials for joints or<br />

bodies with a compliance to achieve<br />

the flexibility needed for realistic undulations<br />

<strong>and</strong> squeezes (Lauder et al.,<br />

2007; Cutkosky & Kim, 2009; Ruiz<br />

et al., 2010). By matching nature as<br />

closely as possible, we gain a greater<br />

underst<strong>and</strong>ing of how nature works.<br />

But is matching reality necessary<br />

What if instead of underst<strong>and</strong>ing<br />

how shape <strong>and</strong> structure are optimized<br />

for best propulsion, we determine<br />

how to achieve stealth by<br />

designing robots that blend in with<br />

the background turbulence or match<br />

the disturbance made by the other<br />

members of the school A robot like<br />

that would enable us to spy on<br />

schools from the inside out, by becoming<br />

a schoolmate. Can we study<br />

16 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


how the sensing system is integrated<br />

with the propulsors to enhance the<br />

acuity of information capture What<br />

principles can we abstract from natural<br />

aquatic propulsion to improve<br />

how we save energy or save materials<br />

or improve performance, as all surviving<br />

species do in a variety of ways to<br />

suit the constraints of the environment<br />

in which they have evolved<br />

Perhaps we can take advantage of the<br />

environment <strong>and</strong> glide <strong>and</strong> surf where<br />

possible, using the free energy of the<br />

sea. Oceanographic gliders like the<br />

Slocum glider Scarlet 27 (Figure 3;<br />

Schofield et al., 2010) are beautifully<br />

designed, wherein one version varies its<br />

ballast by a phase change in response to<br />

ambient ocean temperature (Webb et al.,<br />

2001). By using whale- or copepodlike<br />

buoyancy control (Clarke, 1978;<br />

Pond & Tarling, 2011) <strong>and</strong> harvesting<br />

environmental energy in the ocean<br />

temperature gradient, this glider traveled<br />

great distances with much less<br />

energy than other forms of propulsion.<br />

Simple modifications of shape on<br />

key control surfaces can lead to large<br />

variations in the balance of lift <strong>and</strong><br />

drag essential for tight maneuvering,<br />

as exemplified by the tubercles of<br />

whale fins that now enable wind turbines<br />

to capture energy at lower wind<br />

speeds (Fish et al., 2011b). Can we further<br />

achieve an economy of materials<br />

by adopting the streamlined form of<br />

the shark <strong>and</strong> applying the hierarchical<br />

structure of its denticles to fine<br />

tune that drag reduction (Dean &<br />

Bhushan, 2010)<br />

Indeed, looking at the familiar<br />

players in the sea points out many<br />

cases of convergent evolution in<br />

terms of undulatory or jet-like propulsion.<br />

A closer look reveals other<br />

unusual adaptations for aquatic mobility<br />

such as the parachutes of pteropods<br />

or the floats of siphonophores or<br />

FIGURE 3<br />

(A) Autonomous robots that follow the routes of swimming penguins are collecting information<br />

that could help scientists underst<strong>and</strong> why the birds’ populations are dropping rapidly. The underwater<br />

robots, called gliders, are programmed to record ocean conditions as they follow the<br />

tracks of Adelie penguins swimming in the Southern Ocean surrounding Antarctica. (B) Diagram<br />

of Teledyne-Webb Corporation’s Slocum Glider (coastal model). The Front Main Housing<br />

Section glider’s ballast, <strong>and</strong> consequently its flight, is controlled by moving water into or out<br />

of the Fore Wet Section. From Kahl et al. (2010). (C) The 221-day path taken by Scarlet Knight<br />

(adapted from Schofield et al., 2010; Kahl et al., 2010; images courtesy of O. Schofield: http://<br />

rucool.marine.rutgers.edu).<br />

the multiple oars of copepods, krill,<br />

<strong>and</strong> polychaetes. Analyses of their<br />

transport mechanisms may again<br />

open our eyes to novel designs of future<br />

underwater vehicles that can<br />

steadily hover, smoothly cruise, rapidly<br />

escape, quietly sink or perform any<br />

gait as needed in tight quarters. With<br />

July/August 2011 Volume 45 Number 4 17


the rapid evolution occurring in materials<br />

science research <strong>and</strong> in control<br />

systems, we may find ourselves traveling<br />

through fluids in unexpected<br />

ways.<br />

Author:<br />

Jeannette Yen<br />

School of Biology,<br />

Center for Biologically Inspired Design<br />

Georgia Institute of <strong>Technology</strong>,<br />

Atlanta, GA. 30332-0230<br />

Email: jeannette.yen@biology.gatech.<br />

edu<br />

References<br />

Catton, K.B., Webster, D.R., Kawaguchi, S.,<br />

& Yen, J. 2011. The hydrodynamic disturbances<br />

of two species of krill: Implications for aggregation<br />

structure. J Exp Biol. 214:1845-56.<br />

Clarke, M.R. 1978. Buoyancy control as a<br />

function of the spermaceti organ in the sperm<br />

whale. J Mar Biol Assoc UK. 58:27-71.<br />

doi: 10.1017/S0025315400024395.<br />

Curet, O.M., Patankar, N.A., Lauder, G.V.,<br />

& MacIver, M.A. 2011. Aquatic manoeuvering<br />

with counter-propagating waves: A novel<br />

locomotive strategy. J Roy Soc Interface.<br />

8(60):1041-50. doi: 10.1098/rsif.2010.0493.<br />

Cutkosky, M.R., & Kim, S. 2009. Design<br />

<strong>and</strong> fabrication of multi-material structures for<br />

bioinspired robots. Philos T R Soc A.<br />

367:1799-813. doi: 10.1098/rsta.2009.0013.<br />

Dean, B., & Bhushan, B. 2010. Shark-skin<br />

surfaces for fluid-drag reduction in turbulent<br />

flow: A review. Philos T R Soc A. 368(1929):<br />

4775-806. doi: 10.1098/rsta.2010.0201.<br />

Dean, M.N., Swanson, B.O., & Summers,<br />

A.P. 2009. Biomaterials: Properties, variation<br />

<strong>and</strong> evolution. Integr Comp Biol. 49(1):<br />

15-20. doi: 10.1093/icb/icp012.<br />

Fish, F.E. 2006. Limits of nature <strong>and</strong> advances<br />

of technology: What does biomimetics have to<br />

offer to aquatic robots Appl Bionics Biomech.<br />

3(1):49-60. doi: 10.1533/abbi.2004.0028.<br />

Fish, F.E., Nichols, R.H., Dudas, M.A.,<br />

Moored, K.W., & Bart-Smith, H. 2011a.<br />

Kinematics of swimming in the manta ray<br />

(Manta birostris): 3D analysis of open<br />

water maneuverability. Integr Comp Biol.<br />

51(Suppl. 1):E42.<br />

Fish, F.E., Weber, P.W., Murray, M.M., &<br />

Howle, L.E. 2011b. The tubercles on<br />

humpback whale’s flipper: Application of<br />

bio-inspired technology. Integr Comp Biol.<br />

51(1):203-13. doi: 10.1093/icb/icr016.<br />

Guterl, F. 1996. Playthings of Science.<br />

http://discovermagazine.com/1996/dec/<br />

playthingsofscie946. (accessed 15 July 2011).<br />

Hu, D., Chan, B., & Bush, J.W.M. 2007.<br />

Water-walking devices. Exp Fluids. 43:<br />

769-78. doi: 10.1007/s00348-007-0339-6.<br />

Kahl, L., Oscar Schofield, A., & Fraser, W.R.<br />

2010. Autonomous gliders reveal features<br />

of the water column associated with foraging<br />

by adelie penguins. Integr Comp Biol.<br />

50(6):1041-50. doi: 10.1093/icb/icq098.<br />

Lauder, G.V., Anderson, E.J., Tangorra, J., &<br />

Madden, P.G. 2007. Fish biorobotics: Kinematics<br />

<strong>and</strong> hydrodynamics of self-propulsion.<br />

J Exp Biol. 210:2767-80. doi: 10.1242/<br />

jeb.000265.<br />

Long, J.H.J., Koob, T.J., Irving, K., Combie,<br />

K., Engel, V., Livingston, N., … Schumacher,<br />

J. 2006. Biomimetic evolutionary analysis:<br />

Testing the adaptive value of vertebrate tail<br />

stiffness in autonomous swimming robots.<br />

J Exp Biol. 209:4732-46. doi: 10.1242/<br />

jeb.02559.<br />

Moslemi, A.A., & Krueger, P.S. 2011. The<br />

effect of Reynolds number on the propulsive<br />

efficiency of a biomorphic pulsed-jet<br />

underwater vehicle. Bioinspir Biomim.<br />

6(2):026001. doi: 10.1088/1748-<br />

3182/6/2/026001.<br />

Pond, D.W., & Tarling, G.A. 2011. Phase<br />

transitions of wax esters adjust buoyancy<br />

in diapausing Calanoides acutus. Limnol<br />

Oceanogr. 56(4):1310-8.<br />

Ruiz, L.A., Whittlesey, R.W., & Dabiri, J.O.<br />

2010. Vortex-enhanced propulsion.<br />

J Fluid Mech. 668:5-32. doi: 10.1017/<br />

S0022112010004908.<br />

Schofield, O., Kohut, J., Glenn, S., Morell,<br />

J., Capella, J., Corredor, J., … Boicourt, W.<br />

2010. A regional Slocum glider network in the<br />

Mid-Atlantic coastal waters leverages broad<br />

community engagement. Mar Technol Soc J.<br />

44(6):64-74.<br />

Tangorra, J., Phelan, C., Esposito, C., &<br />

Lauder, G.V. 2011. Use of biorobotic models<br />

of highly deformable fins for studying the<br />

mechanics <strong>and</strong> control of fin forces in fishes.<br />

Integr Comp Biol. 51(1):176-89. doi:<br />

10.1093/icb/icr036.<br />

Triantafyllou, M.S., & Triantafyllou, G.S.<br />

1995. An efficient swimming machine.<br />

Sci Am. 272:64-70. doi: 10.1038/<br />

scientificamerican0395-64.<br />

Webb, D.C., Simonetti, P.J., & Jones, C.P.<br />

2001. SLOCUM: An underwater glider<br />

propelled by environmental energy.<br />

IEEE J Oceanic Eng. 26(4):447-52.<br />

doi: 10.1109/48.972077.<br />

18 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


COMMENTARY<br />

Developing Bioinspired Autonomous Systems<br />

AUTHOR<br />

Thomas M. McKenna<br />

Office of Naval Research<br />

Research that seeks to identify the<br />

principles, strategies, <strong>and</strong> mechanisms<br />

used by motile aquatic animals offers<br />

opportunities to develop undersea<br />

vehicles that exceed current capabilities<br />

<strong>and</strong> enable the Navy to exp<strong>and</strong><br />

the operational envelope of autonomous<br />

undersea vehicles. Autonomous<br />

undersea vehicles serve in a number of<br />

important current <strong>and</strong> emerging roles<br />

in Navy missions, including surveillance<br />

for anti-submarine warfare,<br />

mine countermeasures, Improvised<br />

Explosive Device (IED) detection<br />

<strong>and</strong> localization, force protection (e.g.,<br />

counter-diver missions) in harbors, <strong>and</strong><br />

riverine exploration <strong>and</strong> characterization.<br />

However, current unmanned undersea<br />

vehicles (UUVs) have technical<br />

gaps in areas such as mission duration<br />

(due to power constraints), excessive<br />

noise generation, speed, limited maneuverability,<br />

lack of tight integration<br />

of sensing <strong>and</strong> maneuver, propulsion<br />

<strong>and</strong> maneuver dead zones at low<br />

speeds, <strong>and</strong> inability to operate in the<br />

surf zone. Considering the capabilities<br />

of sea creatures for efficient propulsion<br />

over a large range of speeds, their ability<br />

to operate in or even exploit energetic<br />

ocean environments like the surf<br />

zone, extraordinary maneuverability,<br />

<strong>and</strong> the special sensing evolved for predation,<br />

schooling <strong>and</strong> navigation,<br />

there are many lessons for technologists.<br />

The basic research programs in<br />

Bio-Inspired Autonomous Systems at<br />

the Office of Naval Research (ONR)<br />

have supported research using four<br />

approaches: (1) the identification,<br />

modeling, <strong>and</strong> emulation of the biomechanics<br />

<strong>and</strong> fluid mechanics of<br />

underwater propulsion <strong>and</strong> control<br />

in swimming organisms, (2) identification<br />

<strong>and</strong> exploration of the<br />

sensorimotor control of animals <strong>and</strong><br />

integrated closed loop control using<br />

biosensing (e.g., biosonar, electrosense,<br />

lateral line sensors, optic flow,<br />

magnetic sense), (3) the development<br />

of muscle-like actuators <strong>and</strong> fin designs<br />

that exploit these materials,<br />

<strong>and</strong> (4) design <strong>and</strong> development of<br />

swimming prototype Autonomous<br />

Underwater Vehicle (AUVs) as proof<br />

of principle <strong>and</strong> for performance<br />

evaluation. These studies take their<br />

inspiration from diverse sea creatures<br />

<strong>and</strong> include high-performance swimmers<br />

like bluefin tuna, squid, rays<br />

<strong>and</strong> seals; animals that thrive in the<br />

surf zone like lobsters <strong>and</strong> crabs; <strong>and</strong><br />

animals with special senses like<br />

biosonar (i.e., dolphins) <strong>and</strong> electrosense<br />

(i.e., sharks <strong>and</strong> ribbonfish).<br />

More recently, opportunities have<br />

emerged for microrobotic underwater<br />

systems capable of sensing, reporting,<br />

<strong>and</strong> remediation of environmental<br />

chemicals that exploit the convergence<br />

of synthetic biology, nanotechnology,<br />

<strong>and</strong> electro-optic systems.<br />

This has prompted renewed interest<br />

in the propulsion biology of organisms<br />

that use cilia <strong>and</strong> flagella for<br />

locomotion.<br />

Key Science <strong>and</strong><br />

<strong>Technology</strong> Issues<br />

There are a number of key issues<br />

for the development of bioinspired<br />

autonomous systems.<br />

1. Developing high-efficiency<br />

propulsion that exceeds the capability<br />

of propellers. This is particularly<br />

important for achieving long-duration<br />

missions. Although there is a sizable<br />

Navy effort to develop new energy<br />

sources for underwater vehicles, introduction<br />

of more efficient propulsion<br />

systems would reduce the power requirements<br />

<strong>and</strong> thereby lengthen mission<br />

durations. Early efforts to mimic<br />

the caudal fins of high-performance<br />

swimmers like tuna did not produce<br />

performance that exceeds propellers<br />

(but recent work on the Ghostswimmer<br />

TM , discussed in Rufo <strong>and</strong> Smithers’<br />

commentary in this issue, looks promising),<br />

<strong>and</strong> B<strong>and</strong>yopadhyay (2005) has<br />

produced a meta-analysis showing that<br />

animals do not have an advantage over<br />

man-made systems in cruise, but animals<br />

do show greater maneuverability<br />

relative to man-made underwater<br />

vehicles. One promising new form of<br />

propulsion is the use of bioinspired<br />

high-lift foil propulsors. High-lift propulsion<br />

was introduced in the biological<br />

context in the analysis of fly wing<br />

lift (B<strong>and</strong>yopadhyay, 2009; Ellington,<br />

1984; Dickinson et al., 1999). Fly<br />

wings, which pitch <strong>and</strong> heave with a<br />

90° phase difference, can achieve lift<br />

coefficients substantially greater than<br />

rigid foils that have a constant angle<br />

of attack, <strong>and</strong> high-lift foil propulsors<br />

should be capable of an order of magnitude<br />

greater lift than traditional<br />

naval propellers. High-lift foils also<br />

produce substantially less noise than<br />

traditional propellers. B<strong>and</strong>yopadhyay<br />

et al. (2008) have produced a series of<br />

rigid high-lift foils (roughly analogous<br />

to penguin pectoral fins), characterized<br />

July/August 2011 Volume 45 Number 4 19


their propulsion properties, <strong>and</strong><br />

mounted them on an undersea vehicle.<br />

This first version was called bioinspired<br />

autonomous undersea vehicle (BAUV)<br />

<strong>and</strong> a second, lower-diameter version<br />

was called self-propelled line array<br />

(SPLINE) (Figures 1 <strong>and</strong> 2). High-lift<br />

flapping foils are efficient, <strong>and</strong> combining<br />

six multiple foils on a vehicle can<br />

achieve extraordinary maneuverability<br />

<strong>and</strong> hover <strong>and</strong> exhibit low-noise emission,<br />

but the flapping foil configuration<br />

is not consistent with producing high<br />

speeds. Recently, B<strong>and</strong>yopadhay has<br />

designed a new propulsor called<br />

“Slosher” that has multiple foils with<br />

variable angles of attack arranged radially<br />

that can operate as a high-lift propulsor<br />

at low speeds or, with the foils<br />

locked, as a traditional propeller at<br />

high speeds. This avoids the deadb<strong>and</strong><br />

of controllability at low speeds.<br />

Moreover, hybrid vehicles, such as the<br />

RAZOR (Figure 3), have been developed<br />

that combine four high-lift<br />

foils <strong>and</strong> two props. At low speeds,<br />

the high-lift foils provide high maneuverability<br />

<strong>and</strong> hover, but when the<br />

props provide higher speeds for transit<br />

of the vehicle, the foils become control<br />

surfaces.<br />

FIGURE 1<br />

BAUV developed by Dr. Promode B<strong>and</strong>yopadyay<br />

at the Naval Undersea Warfare Center, Newport,<br />

RI (NUWC-NPT). The BAUV has six high-lift<br />

foils <strong>and</strong> a controller based on a model of the<br />

olivo-cerebellar circuits of mammals.<br />

FIGURE 2<br />

SPLINE. This is a refinement of the BAUV, designed<br />

for long-duration testing of its ability<br />

to pull a load, maintain accurate position, <strong>and</strong><br />

keep a line taut that has been fixed at the distal<br />

end while maneuvering (NUWC-NPT).<br />

FIGURE 3<br />

RAZOR vehicle. This vehicle was developed by<br />

Richard Berube <strong>and</strong> Promode B<strong>and</strong>yopadyay<br />

at NUWC-NPT. It has four high-left foils <strong>and</strong><br />

two rotating propulsors. At low speeds, it<br />

uses the foils for high maneuverability.<br />

2. Developing adaptive controllers<br />

for high-degree-of-freedom<br />

bioinspired propulsors. Many of<br />

the bio-inspired propulsion systems<br />

exhibit high degrees of freedom. For<br />

the case of the high-lift flapping<br />

foils,roll,pitch,<strong>and</strong>frequencyare<br />

motion parameters to be specified to<br />

achieve efficient fin kinematics, <strong>and</strong><br />

for multiple fin vehicles, the phasing<br />

of the fins is critical. In addition to<br />

comm<strong>and</strong>ing these parameters, for a<br />

given vehicle maneuver, the fin controller<br />

must respond to perturbations.<br />

Fortunately, there are bioinspired solutions<br />

to such control issues. One of<br />

the key components of the mammalian<br />

motor control system is the olivocerebellar<br />

system. Coupled neurons<br />

in the inferior olive generate a 10-Hz<br />

rhythm, <strong>and</strong> animal muscle contractions<br />

are initiated from particular<br />

phases of this rhythm. The inferior<br />

olive has multiple oscillating domains,<br />

with phase shifts between these domains,<br />

<strong>and</strong> these domains are under<br />

the control of the cerebellar cortex,<br />

which is a recipient of sensory inputs<br />

from many modalities. Llinas et al.<br />

(2004) developed a model whereby<br />

sequences of motor comm<strong>and</strong>s can<br />

be generated by the olivo-cerebellar<br />

system. The biophysics of coupling<br />

of these neurons also promotes a<br />

phase reset property for rapid synchronization<br />

(Kazantsev et al.,<br />

2004). B<strong>and</strong>yopadhyay (2008) has<br />

implemented this model in analog circuits<strong>and</strong>hasshownittobeaneffective<br />

controller of power <strong>and</strong> thrust in<br />

thehigh-liftfoils<strong>and</strong>theabilityto<br />

rapidly return to normal following<br />

perturbations. This model is able to<br />

synchronize across multiple fins in<br />

order to achieve optimum gaits that<br />

minimize pitching <strong>and</strong> rolling of the<br />

BAUV.Thiscontrollerenabledthe<br />

BAUV to precisely hold a line load<br />

for 20 days. Based on these laboratory<br />

results, the BAUV vehicle is predicted<br />

to support a mission duration of<br />

3 weeks with current battery technology,<br />

although this assumes most of<br />

that time is spent in hover.<br />

For fish <strong>and</strong> eels that use whole<br />

body or caudal fin propulsion,the<br />

oscillations are generated by central<br />

pattern generators. However, the modulation<br />

of these neural generators<br />

during locomotion to achieve desired<br />

thrust <strong>and</strong> vectors has not been fully<br />

characterized for animals, <strong>and</strong> using<br />

this approach in artificial systems requires<br />

learning or genetic algorithms<br />

to tune them to particular maneuvers.<br />

20 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


3. Exploitation of fish swimming<br />

modes for underwater vehicle propulsion<br />

<strong>and</strong> maneuvers. Fish have a<br />

number of swimming modes that are<br />

worthy of consideration for emulation.<br />

Fish swimming types can be<br />

classified as either body <strong>and</strong>/or caudal<br />

fin movements vs. those that use median<br />

<strong>and</strong>/or paired fin propulsion.<br />

One review of fish swimming modes<br />

(Sfakiotakis et al., 1999) singled out<br />

lunate tail propulsion (e.g., tuna), undulating<br />

fins (e.g., some rays), <strong>and</strong><br />

labriform (oscillatory pectoral fin)<br />

swimming mechanisms as having the<br />

greatest potential for exploitation in<br />

artificial systems. ONR is currently<br />

supporting research to exploit all<br />

three of these mechanisms in addition<br />

to the gymnotiform mode that involves<br />

undulations of a long ventral<br />

median fin. Batoid fishes utilize one<br />

of two modes of locomotion, employing<br />

either undulatory (passing multiple<br />

waves down the fin orbody)<br />

or oscillatory (flapping) kinematics<br />

(Rosenberger, 2001). An ambitious<br />

effort to build batoid vehicles, using<br />

predominately oscillating kinematics,<br />

is being undertaken by a team led<br />

by Hillary Bart-Smith (described in<br />

Moored, Fish, Kemp, & Bart-Smith,<br />

this issue) supported by an ONR program<br />

managed by Robert Brizzolara.<br />

Two prototype manta vehicles were<br />

recently demonstrated in a student<br />

competition (Pennisi, 2011).<br />

Exploitation of fish locomotion<br />

using lunate tail fins began with<br />

the seminal studies of Triantafyllou<br />

(Barrett et al., 1996; Anderson et al.,<br />

1998) analyzing how bio-inspired<br />

foils create <strong>and</strong> exploit vortex structures.<br />

His laboratory also developed<br />

the interesting Robotuna (based on<br />

the bluefin tuna) <strong>and</strong> Robopike prototypes,<br />

which were propelled using<br />

caudal fins. Triantafyllou’s group<br />

demonstrated very high-propulsion<br />

efficiencies (up to 91%) for the<br />

Robotuna; however, the Robotuna<br />

did not achieve very high speeds.<br />

One of the creators of Robotuna,<br />

David Barrett at Olin College, is now<br />

teamed with Boston Engineering to<br />

develop the Ghostswimmer vehicle<br />

(described in Rufo’s commentary<br />

in this issue). The objective of the<br />

Ghostswimmer project is to exceed<br />

the performance of current similar<br />

size UUVs on speed, maneuver, mission<br />

duration, noise, rapid response<br />

<strong>and</strong> cost. The goal is to demonstrate<br />

the capability to conduct fully autonomous<br />

missions.<br />

Another active research area seeks<br />

to exploit the principles of fish pectoral<br />

fins for propulsion <strong>and</strong> maneuver.<br />

Fish pectoral fins are highly flexible<br />

<strong>and</strong> enable a high degree of precise<br />

maneuver <strong>and</strong> station keeping in<br />

currents. Sunfish fins have flexible<br />

rays, <strong>and</strong> attached membranes exhibit<br />

a cupping motion on the forward<br />

stroke that produces upper <strong>and</strong> lower<br />

leading edge vortices. These fins generate<br />

positive thrust throughout the fin<br />

beat, <strong>and</strong> turning involves asymmetric<br />

use of these fins. Small prototype fins<br />

with this ray <strong>and</strong> membrane structure<br />

have been shown in the laboratory,<br />

but no free swimming vehicles have<br />

yet been developed (Gottlieb et al.,<br />

2010; Tangorra & Lauder, 2011, this<br />

issue). Rigid pectoral fins have been<br />

implemented on a number of swimming<br />

vehicles, but these are mainly<br />

used as control surfaces.<br />

Aquatic animals that are propelled<br />

by jetting can also provide inspiration<br />

for novel propulsion mechanisms.<br />

Mohseni (Krieg & Mohseni, 2010;<br />

Krieg et al., this issue) has characterized<br />

the biomechanics <strong>and</strong> hydrodynamics<br />

of squid propulsion <strong>and</strong><br />

developed new bioinspired thrusters,<br />

<strong>and</strong> Priya <strong>and</strong> his colleagues (see<br />

Joshi et al., this issue) have built prototype<br />

jellyfish that closely mimic<br />

their swimming modes.<br />

4. Development of muscle-like<br />

actuators. To fully exploit animallike<br />

locomotion <strong>and</strong> sensorimotor<br />

control mechanisms, it is essential to<br />

develop muscle-like actuators, linear<br />

actuators, adaptive compliant structural<br />

materials, <strong>and</strong> elastic skins with<br />

properties much closer to biological<br />

systems than current technology.<br />

There are substantial inefficiencies<br />

with implementing bio-inspired locomotion<br />

using rotary motors <strong>and</strong> complex<br />

power transmission systems. Two<br />

recent efforts to develop fins based on<br />

ionic polymer-metal composites are<br />

described in the papers by Kim et al.<br />

<strong>and</strong> Tan in this issue.<br />

5. Closed-loop control of bioinspired<br />

underwater vehicles. In<br />

crustaceans, there are detailed accounts<br />

of sensorimotor reflexes for control of<br />

tail <strong>and</strong> legs, <strong>and</strong> robotic lobsters have<br />

been built with many levels of bioinspiration<br />

(Ayers & Crisman, 1992;<br />

Ayers et al., this issue). However, for<br />

fish, the neural mechanisms by which<br />

the sensory afferent information on<br />

motion, flows, <strong>and</strong> hydroacoustic<br />

pressure is processed, integrated <strong>and</strong><br />

relayed to motor neurons is not<br />

known. One reason for this is that<br />

electrophysiological recording in<br />

swimming fish is technically challenging.<br />

One open question is whether fish<br />

can sense vortices <strong>and</strong> maneuver or<br />

time fin movements to exploit them<br />

or avoid them. The ONR is currently<br />

supporting a number of efforts in<br />

closed-loop control <strong>and</strong> bionavigation.<br />

These projects include identifying the<br />

role of hydroacoustic receptors in the<br />

lateral line <strong>and</strong> fin <strong>and</strong>bodymechano-reception<br />

in flow <strong>and</strong> vortex<br />

sensing <strong>and</strong> tracing the connectivity<br />

July/August 2011 Volume 45 Number 4 21


of these receptors into spinal sensorimotor<br />

circuits involved in locomotion<br />

<strong>and</strong> fin control (Green et al., 2011;<br />

Green & Hale, 2011). This project is<br />

also developing robotic fish prototypes<br />

(see Tangorra et al., this issue; Lauder<br />

et al., this issue) with detailed pectoral<br />

fins. This work entails some basic research<br />

on fish neurophysiology, since<br />

circuit-level neurophysiology is much<br />

more challenging in fish than in terrestrial<br />

vertebrates, <strong>and</strong> many of the<br />

spinal motor reflexes <strong>and</strong> circuits involved<br />

in limb control that were well<br />

established for mammals by the<br />

1980s are still in nascent stage for the<br />

pectoral fins of fish.<br />

6. Exploitation of special senses<br />

of aquatic animals. Anumberof<br />

fish species <strong>and</strong> sharks are able to navigate<br />

<strong>and</strong> search for prey using electroreception;<br />

some animals also exploit<br />

geomagnetic signals for navigation,<br />

<strong>and</strong> the biosonar of dolphins supports<br />

a range of behaviors. The ribbon fish<br />

can sense prey in murky waters using<br />

electrosense <strong>and</strong> then use its ventral<br />

ribbon fin to maneuver to this prey.<br />

Ongoing ONR projects are performing<br />

system identification of visual<br />

<strong>and</strong> eletroreceptive feedback control of<br />

locomotion in the ribbon fish (Roth<br />

et al., 2011; Mitchell et al., 2011) to<br />

build a working electrosense <strong>and</strong> electromagetosense<br />

module for navigation<br />

<strong>and</strong> target localization <strong>and</strong> to build a<br />

prototype ribbon fish (“Ghostbot”)<br />

(Curet et al., 2011). A vehicle designed<br />

to sense electric <strong>and</strong> magnetic anomalies<br />

could enable new mine countermeasures<br />

systems or enable navigation<br />

that does not depend on GPS or<br />

expensive Inertial Measurement Units<br />

(IMUs).<br />

Sharks exhibit exquisite sensitivity<br />

to perturbations of electric fields <strong>and</strong><br />

one current project seeks to develop<br />

new highly sensitive electromagnetic<br />

<strong>and</strong> hydroacoustic sensors based on<br />

shark sensor biophysics (Kalmijn,<br />

Scripps).<br />

Dolphins have extraordinary abilities<br />

to recognize objects using biosonar.<br />

ONR has supported research characterizing<br />

this ability in terms of psychoacoustics,<br />

<strong>and</strong> several dolphin-inspired<br />

sonars have been developed <strong>and</strong> demonstrated.<br />

Recently, Forsythe et al.<br />

(2008) demonstrated closed-loop<br />

control of his bio-inspired autonomous<br />

underwater vehicle using a simple<br />

dolphin-inspired biosonar in the exploration<br />

of acoustic targets. Dolphins,<br />

however, have the disadvantage that<br />

one cannot analyze directly the neural<br />

circuits involved in their biosonar.<br />

Hence, from the earliest days of its existence,<br />

ONR has supported the study<br />

of bat biosonar. The neural circuits of<br />

bat biosonar have been well characterized.<br />

Recently research has focused on<br />

how bats use multistatic biosonar for<br />

obstacle avoidance. Bats not only have<br />

an extraordinary ability to navigate<br />

around obstacles <strong>and</strong> strike prey<br />

using their own sonar, but they can<br />

also perform these feats using the returns<br />

from calls emitted by other bats<br />

in a swarm. This has implications for<br />

the cooperative behavior of multiple<br />

AUVs.<br />

7. Group behaviors. The ONR<br />

Science of Autonomy program is supporting<br />

efforts to identify the principles<br />

of fish schooling <strong>and</strong> applying them to<br />

multiple AUVs conducting ocean surveys.<br />

Such studies have addressed the<br />

minimal sensing requirements of fish<br />

used in schooling, including vision.<br />

Additionally, for AUVs, one could conceivably<br />

use man-made communications<br />

like acoustic modems or lasers to<br />

enhance coordinated actions, but successful<br />

emulation of sense modalities<br />

used in nature (i.e., fish hydroacoustic<br />

sensing) might greatly simplify coordination<br />

of multiple vehicles—in particular<br />

when a large number of vehicles<br />

is operating at close quarters. There is<br />

also an effort to emulate the group<br />

hunting behaviors of dolphins in intelligent<br />

control systems for cooperative<br />

autonomous vehicles.<br />

8. The Navy is interested in distributed,<br />

persistent sensing systems<br />

for tasks such as anti-submarine warfare.<br />

One bioinspired approach to<br />

this is to develop small sensor platforms<br />

that float in the water column<br />

with limited mobility, similar to jellyfish.<br />

An ambitious effort to exploit<br />

multifunctional materials <strong>and</strong> build<br />

such jellyfish colonies (Priya) is underway.<br />

Additionally, the emerging convergence<br />

of nanotechnology, synthetic<br />

biology <strong>and</strong> micro-optoelectronics has<br />

fostered the nascent development of<br />

very small autonomous systems. Some<br />

of the new microrobots being designed<br />

have components consisting of colonies<br />

of micro-organisms <strong>and</strong> hybrid<br />

biomolecules, with motility being<br />

provided by a multitude of cilia or<br />

flagella. These new systems could<br />

function as distributed sensors of hazardous<br />

materials, but with integrated<br />

mobility, mitigation <strong>and</strong> reporting<br />

capabilities.<br />

In summary, there are a number of<br />

promising opportunities for the Navy<br />

to extend the capabilities of autonomous<br />

undersea platforms using bioinspired<br />

technologies in propulsion,<br />

control <strong>and</strong> sensing.<br />

Author:<br />

Dr. Thomas McKenna<br />

Office of Naval Research<br />

Division of Human <strong>and</strong><br />

Bioengineered Systems<br />

875 N. R<strong>and</strong>olph St., Suite 1425<br />

Arlington, VA 22203-1995<br />

Email: tom.mckenna@navy.mil<br />

22 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


References<br />

Anderson, J.M., Streitlien, K., Barrett, D.S.,<br />

& Triantafyllou, M.S. 1998. Oscillating<br />

foils of high propulsive efficiency. J Fluid<br />

Mech. 360:41-72. doi: 10.1017/<br />

S0022112097008392.<br />

Ayers, J., & Crisman, J. 1992. The lobster as<br />

a model for an omnidirectional robotic<br />

ambulation control architecture. In: Biological<br />

Neural Networks in Invertebrate Neuroethology<br />

<strong>and</strong> Robotics, eds. Beer, R., Ritzmann, R.,<br />

& McKenna, T., 287-316. San Diego:<br />

Academic Press.<br />

B<strong>and</strong>yopadhyay, P.R. 2005. Trends in biorobotic<br />

autonomous undersea vehicles. IEEE J<br />

Oceanic Eng. 30(1):109-39. doi: 10.1109/<br />

JOE.2005.843748.<br />

B<strong>and</strong>yopadhyay, P.R. 2009. Swimming <strong>and</strong><br />

flying in nature—The route toward applications:<br />

The Freeman Scholar lecture. ASME J Fluid<br />

Eng. 131:031801-29. doi: 10.1115/1.3063687.<br />

B<strong>and</strong>yopadhyay, P., Beal, D.N., & Menozzi,<br />

A. 2008. Biorobotic insights into how animals<br />

swim. J Exp Biol. 211:206-14. doi: 10.1242/<br />

jeb.012161.<br />

B<strong>and</strong>yopadhyay, P.R. 2008. Synchronization<br />

of animal-inspired multiple high-lift fins<br />

in an underwater vehicle using olivo-cerebellar<br />

dynamics. IEEE J Oceanic Eng. 33(4):<br />

563-78. doi: 10.1109/JOE.2008.2005356.<br />

Barrett, D., Grosenbaugh, M., & Triantafyllou,<br />

M. 1996. The optimal control of a flexible<br />

hull robotic undersea vehicle propelled by an<br />

oscillating foil. In: Proc. 1996 IEEE AUV<br />

Symp. 1-9.<br />

Curet, O.M., Patankar, N.A., Lauder, G.V.,<br />

& MacIver, M.A. 2011. Mechanical properties<br />

of a bio-inspired robotic knifefish with an<br />

undulatory propulsor. Bioinspir Biomim.<br />

6(2). doi: 10.1088/1748-3182/6/2/026004.<br />

Dickinson, M.H., Lehman, F., & Sane, S.P.<br />

1999. Wing Rotation <strong>and</strong> the aerodynamic<br />

basis of insect flight. Science. 284:1954-60.<br />

doi: 10.1126/science.284.5422.1954.<br />

Ellington, C.P. 1984. The aerodynamics<br />

of hovering insect flight. IV. Aerodynamic<br />

mechanisms. Philos T R Soc Lond B.<br />

305:79-113.<br />

Forsythe, S.E., Leinhos, H.A., &<br />

B<strong>and</strong>yopadhyay, P.R. 2008. Dolphininspired<br />

combined maneuvering <strong>and</strong> pinging<br />

for short-distance echolocation. J Acoust<br />

Soc Am. 124(4):255-61.<br />

Gottlieb, J.R., Tangorra, J.L., Esposito,<br />

C.J., & Lauder, G.V. 2010. A biologically<br />

derived pectoral fin for yaw turn maneuvers.<br />

Appl Bionics Biomech. 7:41-55.<br />

doi: 10.1080/11762320903537782.<br />

Green, M.H., Ho, R.K., & Hale, M.E. 2011.<br />

Movement <strong>and</strong> function of the pectoral fins<br />

of the zebrafish (Danio rerio) during slow<br />

swimming. J Exp Biol. in review.<br />

Green, M.H., & Hale, M.E. 2011. Motoneurons,<br />

muscles <strong>and</strong> movement: Neural<br />

control of the pectoral fins during slow <strong>and</strong><br />

fast swimming. Integr Comp Biol. E51.<br />

Kazantsev, V.B., Nekorkin, V.I., Makarenko,<br />

V.I., & Llinas, R. 2004. Self-referential phase<br />

reset based on inferior olive oscillator dynamics.<br />

P Natl Acad Sci. 101(52):18183-8.<br />

doi: 10.1073/pnas.0407900101.<br />

Krieg, M., & Mohseni, K. 2010. Dynamic<br />

modeling <strong>and</strong> control of biologically inspired<br />

vortex ring thrusters for underwater robot<br />

locomotion. IEEE T Robot. 26(3):542-54.<br />

doi: 10.1109/TRO.2010.2046069.<br />

Llinas, R., Leznick, E., & Makarenko, V.I.<br />

2004. The olivo-cerebellar circuit as a universal<br />

motor control system. IEEE J Ocean Eng.<br />

29(3):631-9. doi: 10.1109/JOE.2004.833212.<br />

Mitchell, T., Fortune, E.S., & Cowan, N.J.<br />

2011. Reweighting of vision <strong>and</strong> electrosensation<br />

during locomotion in the Glass Knifefish.<br />

PLoS, to be submitted.<br />

Pennisi, E. 2011. Manta machines. Science.<br />

332:28-9. doi: 10.1126/science.332.6033.1028.<br />

Rosenberger, L.J. 2001. Pectoral fin locomotion<br />

in batoid fishes: Undulation versus<br />

oscillation. J Exp Biol. 204:379-94.<br />

Roth, E., Zhuang, K., Stamper, S., Fortune,<br />

E.S., & Cowan, N.J. 2011. Stimulus predictability<br />

mediates a switch in locomotor<br />

smooth pursuit performance for Eigenmannia<br />

virescens. J Exp Biol. 214:1170-80.<br />

doi: 10.1242/jeb.048124.<br />

Sfakiotakis, M., Lane, D.M., & Davies,<br />

J.B.C. 1999. Review of fish swimming modes<br />

for aquatic locomotion. IEEE J Oceanic Eng.<br />

24(2):237-52. doi: 10.1109/48.757275.<br />

Tangorra, J.L., & Lauder, G.V. 2011. Biorobotic<br />

models of highly deformable fins for<br />

Studies of the mechanics <strong>and</strong> control of<br />

fin forces. Integr Comp Biol. in press.<br />

doi: 10.1093/icb/icr036.<br />

July/August 2011 Volume 45 Number 4 23


COMMENTARY<br />

GhostSwimmer AUV: Applying <strong>Biomimetics</strong><br />

to Underwater Robotics for Achievement<br />

of Tactical Relevance<br />

AUTHORS<br />

Michael Rufo<br />

Mark Smithers<br />

Boston Engineering Corporation<br />

Introduction<br />

I<br />

t is clear that unmanned underwater<br />

vehicles (UUVs), autonomous<br />

underwater vehicles (AUVs), <strong>and</strong><br />

other waterborne robots are successfully<br />

addressing critical capability<br />

requirements for many customers, including<br />

those in defense <strong>and</strong> oil <strong>and</strong><br />

gas. There remains, however, despite<br />

the best efforts of many entities, capability<br />

gaps for these systems.<br />

While UUVs share many of the<br />

same navigation, power, <strong>and</strong> logistical<br />

challenges as their unmanned ground<br />

vehicle (UGV) <strong>and</strong> unmanned aerial<br />

vehicle (UAV) counterparts, they are<br />

subject to additional challenges including<br />

limited communications <strong>and</strong> harsh<br />

environments.<br />

Developing biologically-inspired<br />

(or biomimetic) technologies can provide<br />

some guidance for next generation<br />

systems. The Advanced Systems<br />

Group (ASG) at Boston Engineering<br />

(Waltham, MA) is developing the<br />

GhostSwimmer<br />

<br />

AUV, for example,<br />

which endeavors to attack many of the<br />

problems facing current UUVs. The<br />

increasing interest in the use of longrange/long-duration<br />

UUVs for littoral<br />

observation, military surveillance, <strong>and</strong><br />

other missions dem<strong>and</strong>s alternative or<br />

new technology development. As per<br />

the U.S. Navy UUV Master Plan<br />

(2004), the areas of autonomy, sensors,<br />

<strong>and</strong> communications are<br />

among the leading areas of interest.<br />

Energy <strong>and</strong> propulsion are also main<br />

areas of interest as per this document,<br />

where it states that “advanced energy<br />

<strong>and</strong> propulsion, in combination with<br />

other UUV technologies, will enable<br />

the use of smaller vehicles (reducing<br />

cost) in the long term, <strong>and</strong> will provide<br />

greater performance.” (The<br />

Navy UUV Master Plan, 2004).<br />

GhostSwimmer is a tactical, biomimetic<br />

autonomous “artificial fish”<br />

UUV that employs the mechanics<br />

<strong>and</strong> dynamics of biological systems to<br />

create efficient swimming <strong>and</strong> high<br />

maneuverability while remaining responsive<br />

to the needs of current riverine<br />

<strong>and</strong> littoral missions (among other<br />

possibilities). Its importance lies in its<br />

ability to provide advanced mobility<br />

in a system that employs payloads.<br />

Considerable work has been completed<br />

on underst<strong>and</strong>ing the propulsive<br />

characteristics of individual fish<br />

fins (extensive information on oscillating<br />

foil propulsion exists). The<br />

GhostSwimmer<br />

<br />

builds upon this<br />

<strong>and</strong> is modeled after a tuna. It is propelled<br />

by a composite construction lunate<br />

caudal tail fin (takingitsname<br />

from its crescent moon-like profile).<br />

In the biological tuna, almost all of<br />

theusablethrustisgeneratedbythe<br />

oscillating tail fin. Magnuson describes<br />

it as “a tapered hydrofoil with high<br />

aspect ratio, curved leading edge <strong>and</strong><br />

moderate sweep back. In cross section,<br />

the fin is shaped like a thin symmetrical<br />

airfoil with a rounded anterior, or<br />

leading edge, <strong>and</strong> a sharp posterior or<br />

trailing edge” (Magnuson, 1978).<br />

This oscillating foil has the capability<br />

of producing thrust at high efficiencies.<br />

Triantafyllou <strong>and</strong> Triantafyllou<br />

(1993) stated “Oscillating foils are<br />

known under certain conditions to<br />

produce substantial thrust at high propulsive<br />

efficiency. Fish <strong>and</strong> cetaceans<br />

have developed through evolution a<br />

presumed optimal manner of propulsion<br />

primarily through flapping of<br />

the posterior part of their body. It is<br />

shown that thrust production depends<br />

significantly on the dynamics of the<br />

unstable wake formed behind the foil<br />

or fish tail: thrust develops through<br />

the formation of a reverse Karman<br />

street, whose preferred Strouhal numbers<br />

are between 0.25 <strong>and</strong> 0.35. Experimental<br />

data from flapping airfoils<br />

<strong>and</strong> data from fish observations confirm<br />

the theoretical predictions.”<br />

Triantafyllou <strong>and</strong> Barrett were able to<br />

construct a precision oscillating foil<br />

test apparatus with which propulsive<br />

efficiencies in the range of 80%<br />

were repeatedly measured (Barrett,<br />

1996).<br />

While based on this concept,<br />

GhostSwimmer<br />

<br />

strives to further<br />

the underst<strong>and</strong>ing of how fish use all<br />

their fins to enhance propulsive performance,<br />

maneuverability, <strong>and</strong> stealth<br />

characteristics (all the while providing<br />

a capability for the Warfighter).<br />

24 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Overcoming the Challenges<br />

for the Future AUVs<br />

At this time, UUVs <strong>and</strong> AUVs<br />

struggle with complex underwater<br />

environments such as shallow <strong>and</strong><br />

obstacle-filled waters. These vehicles<br />

are subject to an intrinsic paradox;<br />

they require stable dynamic behavior<br />

for maintaining intended path/heading<br />

without changing control surfaces or<br />

propulsion force. However, in complex<br />

environments, the vehicle must also be<br />

able to quickly change both path <strong>and</strong><br />

speed safely. Dynamic stability <strong>and</strong> response<br />

aspects such as overshoot, rise<br />

time, <strong>and</strong> peak time become critical<br />

(Azarsinal et al., 2007). Based on their<br />

prowess in these areas, it makes sense to<br />

investigate biological sources of inspiration<br />

in the development of new<br />

man-made systems hoping to excel in<br />

these environments.<br />

The Boston Engineering Advanced<br />

Systems Group’s GhostSwimmer<br />

<br />

AUV (Figure 1) is intended to leverage<br />

this biological inspiration. Existing<br />

AUVs are generally propelled by rotary<br />

propellers driven by electric motors<br />

with energy stored in batteries. Small<br />

diameter propellers typically operate<br />

at low efficiencies <strong>and</strong> can suffer serious<br />

lag times in transient response<br />

(Barrett, 1996). Cost effective propeller<br />

improvements are often limited by<br />

a maximum practical diameter that can<br />

be mounted on a UUV. In complex<br />

underwater areas, rotating propellers<br />

FIGURE 1<br />

Boston Engineering’s GhostSwimmer PH I<br />

AUV (sponsor: ONR Code 341).<br />

also present a significant snagging<br />

risk. Energy technology, despite recent<br />

progress, is awaiting a breakthrough to<br />

provide significantly larger (<strong>and</strong> safer)<br />

power densities; therefore, producing<br />

a technology that can address both<br />

efficiency <strong>and</strong> control, independent<br />

of battery technology, has benefits.<br />

Additional challenges that UUVs<br />

are subject to range from being as simple<br />

as packaging electronics to be water<br />

tight at depth (versus splash-proof ),<br />

to using materials that are resistant to<br />

aggressive corrosion, to being able to<br />

localize underwater without the aid of<br />

conventional techniques (such as the<br />

GPS or radio frequency [RF] communications).<br />

As discussed later in this<br />

commentary, biomimetics can offer<br />

guidance in these areas as well.<br />

<strong>Biomimetics</strong> <strong>and</strong><br />

Pragmatism<br />

<strong>Biomimetics</strong> (from the Greek bios<br />

[life] + mimesis [to imitate]) is not<br />

new. Mankind has been mimicking<br />

nature in products for centuries. Simple<br />

everyday items such as salad tongs<br />

(based on bird beaks) through complex<br />

sensors such as sonar (based on<br />

dolphins <strong>and</strong> bats) have each been<br />

generated by mimicking biology.<br />

Even human terminology mimics nature,<br />

consider saw “teeth”, computer<br />

“viruses” <strong>and</strong> “worms”, orlaptops<br />

that “hibernate.” Engineers <strong>and</strong> researchers<br />

have even mimicked plants<br />

in robotic systems (turning to face<br />

the sun for charging, for example)<br />

(Bar-Cohen, 2006). It should be<br />

noted that this discussion is not<br />

about synthetic life (involving using<br />

biological components to build biological<br />

systems).<br />

Nature evolves by responding to<br />

needs for surviving to the next generation.<br />

It benefits from millions of<br />

variations <strong>and</strong> trial <strong>and</strong> error experiments<br />

over the millennia. Engineers<br />

trying to develop tactically relevant<br />

technologies obviously must “evolve”<br />

much faster <strong>and</strong> in a pragmatic manner.<br />

As such, developing biomimetic<br />

technologies is as much about deciding<br />

what not to imitate as it is deciding<br />

what to imitate (Figure 2). This is because<br />

direct <strong>and</strong> absolute mimicry is<br />

often not appropriate or necessary.<br />

Consider that mankind flies without<br />

flapping wings as birds do. While underst<strong>and</strong>ing<br />

<strong>and</strong> mimicking the control<br />

surface design has direct benefits<br />

to engineer fast <strong>and</strong> high flying aircraft,<br />

engineers diverted from nature’s design<br />

with great success. However, the<br />

definition of the mission must always<br />

be at the core of any engineering effort.<br />

If the goal were to perch on a power<br />

line, even mankind’s best aircraft<br />

would fail.<br />

Evolutionary improvements in nature<br />

also do not always exactly coincide<br />

with the reasons for mimicking them.<br />

This should be considered when deciding<br />

whether the model is appropriate<br />

or ideal for mimicking in an engineered<br />

system designated for use in<br />

critical applications. Clearly, biologists<br />

<strong>and</strong> others trying to learn about “how<br />

biology works” are concerned with<br />

mimicking exactly how a biological<br />

FIGURE 2<br />

Early attempts at biomimetics often missed<br />

the mark (Fuller).<br />

July/August 2011 Volume 45 Number 4 25


model operates, its exact structure, its<br />

mechanical parameters, <strong>and</strong> more.<br />

The focus of this commentary is on<br />

the development of field-capable technologies<br />

for performing unmanned<br />

tasks in relevant mission spaces. In<br />

this sense, it is important to recognize<br />

that animals were evolved for survival<br />

in a particular environment with specific<br />

predators <strong>and</strong> other influences.<br />

In the GhostSwimmer<br />

<br />

case, extant<br />

tuna appeared roughly 60 million<br />

years ago (Dickson & Graham, 2004).<br />

These early tunas lived in a large circumtropical<br />

waterway that encircled<br />

theentireplanet(theTethysSea).<br />

This waterway existed for roughly<br />

50 million years during which the development<br />

of tunas was influenced by<br />

changes in the oceanography, induced<br />

primarily by tectonic activity. These<br />

changes increased productivity, exp<strong>and</strong>ed<br />

food webs, <strong>and</strong> opened potential niches<br />

(Dickson & Graham, 2004).<br />

Graham <strong>and</strong> Dickson suggest that<br />

the appearance of these more extensive<br />

ocean areas with high productivity <strong>and</strong><br />

diversified food webs provided the<br />

catalyst for tuna evolution: enhanced<br />

locomotor performance <strong>and</strong> favored<br />

migratory behavior. Tuna’s specializations,<br />

thunniform swimming, capacity<br />

for regional endothermy, <strong>and</strong> an elevated<br />

aerobic capacity, are thought to<br />

be based on the need for extended<br />

swimming (Graham & Dickson,<br />

2004). Despite the fact that this is different<br />

than the pressing missions of a<br />

UUV (such as mine counter measures),<br />

this extended swimming implies<br />

endurance <strong>and</strong> propulsion that<br />

isofvaluetoapragmaticmission.<br />

This leads to the question the roboticist<br />

must ask: What aspects of a<br />

fish’s “design” or mechanics is applicable<br />

to the desired mission<br />

Another consideration is the different<br />

resources that are available. Evolution<br />

<strong>and</strong> robotics engineers have<br />

distinctly different tool sets, materials,<br />

actuators, <strong>and</strong> power <strong>and</strong> control systems<br />

at their disposal. In certain<br />

areas, engineers have an advantage; exotic<br />

materials such as titanium may<br />

allow advances that nature could not<br />

provide. However, nature still has the<br />

advantage (at least currently) in actuator<br />

power density (when considering<br />

b<strong>and</strong>width <strong>and</strong> other factors), regenerative<br />

capability, sensing, <strong>and</strong> control<br />

prowess. For example, many biomechanics<br />

sources have noted that muscle<br />

tendon series elasticity is critical for<br />

energetic <strong>and</strong> efficiency purposes<br />

(Paluska & Herr, 2006). Many advances<br />

are in the works in the areas of<br />

artificial muscles, neural networks, <strong>and</strong><br />

similar technologies, but to date they<br />

lag behind their biological counterparts<br />

in key areas. Some areas in need<br />

of advancement for actuation include<br />

power to weight ratio, efficiency, fatigue<br />

life, <strong>and</strong> controllability.<br />

Where the Challenges<br />

of Underwater Operation<br />

<strong>and</strong> <strong>Biomimetics</strong> Intersect<br />

The challenges of achieving autonomy<br />

with unmanned systems are<br />

detailed in many articles <strong>and</strong> publications.<br />

From one perspective, autonomy<br />

intrinsically dem<strong>and</strong>s that<br />

engineered systems act like biological<br />

systems. For instance, these systems<br />

need obstacle detection <strong>and</strong> avoidance<br />

or situational awareness (“what<br />

is happening around me”), decisionmaking<br />

capacity (“should I respond<br />

aggressively or passively”), <strong>and</strong> health<br />

monitoring (“am I OK”). Interesting<br />

work in developing biologically<br />

inspired autonomy solutions based<br />

ontheseprinciplesiscurrentlyoccurring<br />

at Jet Propulsion Lab (JPL)<br />

(space mission systems) (Huntsberger,<br />

2001), Office of Naval Research<br />

(ONR) (sensory control, biomechanics)<br />

(ONR: Bio-Inspired Autonomous<br />

Systems), <strong>and</strong> Cornell (making increasingly<br />

complex machines) (Lipson).<br />

The modeling <strong>and</strong> underst<strong>and</strong>ing<br />

of unsteady hydrodynamics is also a<br />

challenge for underwater vehicle<br />

developers. While analogous to aerodynamic<br />

forces in some ways, hydrodynamics<br />

in unsteady, obstacle-laden<br />

environments is often more complex<br />

<strong>and</strong>, to date, difficult to model. In<br />

particular, controlled 6 degrees of freedom<br />

movement, particularly in unsteady<br />

conditions, is challenging for<br />

control, power, <strong>and</strong> propulsion systems.<br />

Underwater vehicles have a<br />

more challenging environment for<br />

controlled mobility than most UGVs<br />

due to this dynamic nature of their<br />

surroundings, <strong>and</strong> unless operating in<br />

open seas (blue water), they can face<br />

far more obstacles within tighter spaces<br />

than UAVs.<br />

Sensing <strong>and</strong> communication is another<br />

area where underwater systems<br />

have very different challenges. Biological<br />

systems have remarkable <strong>and</strong> innate<br />

abilities to detect obstacles or<br />

prey using a variety of sensory capabilities<br />

from echolocation (bottlenose<br />

dolphin biosonar is “probably the<br />

most sophisticated target location <strong>and</strong><br />

analysis system in existence”) (Fulton,<br />

2010) to lateral lines (in fish such as<br />

the tuna, water flow parallel to the motion<br />

deflects cilium in the lateral line<br />

which in turn defines water velocity)<br />

(Martiny et al., 2009), to electric fields<br />

(weakly electric black ghost knife fish<br />

generate an electric field that causes<br />

voltage perturbations due to the differences<br />

in electrical conductivity between<br />

an object <strong>and</strong> the water, thereby<br />

sensing its surroundings) (Martiny<br />

et al., 2009).<br />

26 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


However, unmanned systems can<br />

leverage only those sensory capabilities<br />

developed by the technology community<br />

at large. Herein lies the challenge,<br />

UAVs may have similar collision<br />

avoidance sensory needs, but the technologies<br />

for in-air sensing are currently<br />

more advanced <strong>and</strong>, due to relatively<br />

low signal attenuation in air, can effectively<br />

sense objects hundreds or<br />

thous<strong>and</strong>s of feet away (ignoring<br />

cloud cover <strong>and</strong> other aspects for sake<br />

of argument). The attenuation of<br />

various sensing signals in salt water is<br />

drastically more than in air, making<br />

technologies that many take for<br />

granted (WiFi <strong>and</strong> Bluetooth for example)<br />

impractical underwater. RF<br />

communications can work underwater<br />

but attenuation dem<strong>and</strong>s that low frequencies<br />

be used <strong>and</strong> the resulting<br />

trade-off is b<strong>and</strong>width (a reasonably<br />

sized acoustic modem can provide<br />

140 bps-15 kbps [www.benthos.com]<br />

where even limited WiFi [802.11 b for<br />

example] can provide 2 Mbps “in air”<br />

[Mitchell]). To achieve real-time<br />

or near real-time control, a UGV or<br />

UAV could use high-frequency,<br />

high-b<strong>and</strong>width RF communications.<br />

To date, this is a major challenge in underwater<br />

communications. Additionally,<br />

the change in medium<br />

(from air into <strong>and</strong> through water or<br />

vice versa) makes communication difficult<br />

due to refraction losses at the<br />

interface (losses are large for electromagnetic<br />

waves going from air into<br />

seawater <strong>and</strong> intrinsic wavelength differences<br />

make an underwater antenna<br />

for the same radio different than its<br />

in-air counterpart) (Butler, 2011).<br />

St<strong>and</strong>ard acoustic <strong>and</strong> light-based<br />

sensors are greatly affected by the attenuation<br />

as well. Additionally, their<br />

operation often dem<strong>and</strong>s smooth,<br />

controlled motion at relatively slow<br />

speeds. When coupling the hydrodynamic<br />

maneuverability challenges<br />

for AUVs with these sensing challenges,<br />

one can only marvel at the ability of<br />

fast-moving fish <strong>and</strong> marine mammals<br />

to detect <strong>and</strong> maneuver. For these animals,<br />

this detection <strong>and</strong> maneuver is a<br />

system level process where sensing<br />

provides exterioception <strong>and</strong> proprioception<br />

<strong>and</strong> enables the vorticity <strong>and</strong><br />

optimized flow control that is essential<br />

to maneuverability <strong>and</strong> fast swimming<br />

(Triantafyllou et al., 2002).<br />

The GPS is another technology<br />

taken for granted in the 21st century.<br />

However, current AUVs must surface<br />

to get GPS fixes because GPS will not<br />

work underwater; the signals from the<br />

satellites cannot sufficiently penetrate<br />

the water. This presents the AUV<br />

with a significant challenge in localization<br />

<strong>and</strong> navigation. Until these technologies<br />

advance in a cost effective<br />

manner for underwater applications,<br />

one of the most effective means for<br />

overcoming the challenge includes<br />

having the ability to surface quickly<br />

(<strong>and</strong> often covertly) where the manmade<br />

underwater system can then<br />

leverage in-air wireless <strong>and</strong> geopositioning<br />

technologies. This rapid<br />

surfacing can also benefit frombiologically<br />

inspired techniques, as the<br />

GhostSwimmer has implemented.<br />

Beyond Propulsion—<br />

Leveraging Biology<br />

to Advance the State<br />

of the Art<br />

Designing systems with bioinspired<br />

control systems can open avenues<br />

for efficiently performing, <strong>and</strong><br />

switching between, necessary behavior<br />

states. Advances in computing technology<br />

(increased memory <strong>and</strong> computational<br />

power in smaller, less<br />

expensive packages) now enables incorporation<br />

of newer <strong>and</strong> more advanced<br />

biologically inspired control<br />

systems. This includes the development<br />

of intelligent control through<br />

distributed control in systems containing<br />

their own “nervous systems,” structured<br />

like their biological analog that<br />

can then “learn” through adaptive<br />

techniques (Thomopoulos & Braught,<br />

1995).<br />

The study of biology itself can provide<br />

advances in the underst<strong>and</strong>ing of<br />

unsteady hydrodynamics as mentioned<br />

above. Biological systems<br />

achieve high efficiencies <strong>and</strong> maneuverability<br />

through the sensing, manipulation,<br />

<strong>and</strong> creation of optimized flow<br />

around them. This includes the manipulation<br />

of tip vortices at control<br />

<strong>and</strong> propulsive surfaces as well as the<br />

optimization of output motion to generate<br />

smooth <strong>and</strong> effective jets in the<br />

flow behind the body. The underst<strong>and</strong>ing<br />

of these phenomena is directly<br />

applicable to the creation of<br />

higher performance AUVs <strong>and</strong> even<br />

manned systems (Barrett, 1996).<br />

Boston Engineering’s ASG is currently<br />

applying many of these principles to<br />

develop improved control surfaces for<br />

the U.S. Navy Sea Systems Comm<strong>and</strong><br />

as well as advanced AUVs for ONR.<br />

Navigation is also an area that can<br />

benefit from biomimetics. Specific<br />

areas of interest in autonomy to the<br />

Navy include path planning, behavior<br />

development, localization, on-board<br />

mapping of environmental variability,<br />

<strong>and</strong> effective man-machine interfaces<br />

with a limited communication capability(Wernli,2001).Additionally,increasing<br />

uncertainty regarding Global<br />

Navigation Satellite Systems reliability<br />

has led unmanned systems developers<br />

to seek valuable alternatives. Integrating<br />

inertial systems with reference maps of<br />

Geophysical Fields of the Earth (GFE)<br />

is an area being explored by aerospace<br />

July/August 2011 Volume 45 Number 4 27


entities that offers promise through use<br />

of the recent advancements in embedded<br />

micro-processing, including<br />

memory devices’ capability <strong>and</strong> miniature<br />

size. GFEs, properties of the earth<br />

itself, are already well mapped in geographical<br />

system coordinates <strong>and</strong> can<br />

be considered a reliable navigation<br />

data source. Earth’s MagneticField<br />

(EMF) maps, models, <strong>and</strong> charts are<br />

currently in use for military <strong>and</strong> commercial<br />

entities (for directional information)<br />

<strong>and</strong> are available for at least<br />

98% of the earth’s surface (including<br />

water-covered areas) (Goldenberg,<br />

2006).<br />

Research <strong>and</strong> behavioral experiments<br />

have shown that various animals<br />

sense <strong>and</strong> use the earth’s magnetic field<br />

for navigation over both long <strong>and</strong> short<br />

distances (Johnsen & Lohmann,<br />

2005). Migratory animals, capable of<br />

sensing variations in geomagnetic<br />

fields, reference the earth’s mean field<br />

<strong>and</strong> its inclination at many points.<br />

Measurements of the field, including<br />

spatial variations <strong>and</strong> temporal evolutions,<br />

tabulated by the U.S. Geological<br />

Survey, allow the potential for geomagnetic<br />

field sensors to mimic animal<br />

behavior related to navigation (Zhai<br />

et al., 2007).<br />

Some animals, including certain<br />

birds, sea turtles, salam<strong>and</strong>ers, <strong>and</strong> lobsters<br />

can discriminate small differences<br />

in some of the earth’s magnetic features.<br />

They use positional information<br />

in the earth’s field in several different<br />

ways <strong>and</strong> some actually learn the magnetic<br />

topography of the areas they call<br />

home. Some have postulated that animals<br />

have two separate magnetosensory<br />

systems (Goldenberg, 2006). A<br />

compass alone is rarely sufficient to<br />

guide animals to specific destinations<br />

or along a long <strong>and</strong> complex migratory<br />

route due to currents <strong>and</strong> other errorinducing<br />

phenomena. Navigation<br />

must be enhanced by the ability to<br />

determine position relative to a destination<br />

(human travelers use a GPS),<br />

i.e., positional information inherent<br />

in the earth’s magneticfield provides<br />

a similar, although less precise, assessment<br />

of location (Goldenberg, 2006).<br />

Animals also use other cues; research<br />

has shown that bees navigate<br />

relative to the sun by using the sun as<br />

a fixed point <strong>and</strong> orienting themselves<br />

by maintaining a fixed angle between<br />

its line of flight <strong>and</strong> the line to the<br />

sun (www.physics.ohio-state.edu).<br />

Ocean waves combine with the earth’s<br />

magnetic field to serve as orientation<br />

cues for newly hatched turtles; while<br />

older turtles are following a map they<br />

learned that enables them to establish<br />

their position relative to some distant<br />

target (Lohmann et al., 2004). Regardless<br />

of surface currents or other<br />

oceanographic features, sockeye salmon<br />

use an internal “map sense” to navigate<br />

home after several migrations <strong>and</strong> induced<br />

errors. An olfactory “imprint”<br />

is made on smelts as they leave their<br />

home stream, but approaching their<br />

stream from open sea dem<strong>and</strong>s one<br />

other imprint. Fish are perceptive of<br />

the azimuth <strong>and</strong> altitude of the sun,<br />

but during overcast days, ferromagnetic<br />

mineral magnetite in their brain<br />

may function as a biological compass.<br />

Another means for a fish to sense the<br />

magnetic field is by merely moving<br />

through the water (like a wire moved<br />

across a magnetic field, electrical<br />

current occurs in the wire) (Gedney,<br />

1984). It is possible that AUVs<br />

could mimic these approaches.<br />

The electric field of ocean currents<br />

indicate to sharks their drift relative to<br />

the bottom or to deeper water layers.<br />

Electric current of an ocean stream invading<br />

a quiet bay may provide a shark<br />

with directional cues in familiar territory.<br />

They may also explore the fields<br />

by occasionally diving deeper or to the<br />

bottom <strong>and</strong> their orientating in uniform<br />

DC electric fields has been proven<br />

behaviorally. The electric sense operates<br />

in a passive mode, whereas magnetic<br />

field detection is an active mode, allowing<br />

them to simultaneously sense drift<br />

with ocean streams <strong>and</strong> magnetic headings.<br />

Based on this research, measured<br />

potentials, sensitivity <strong>and</strong> noise effects,<br />

<strong>and</strong> amplification are relatively well<br />

known (Kalmijn, 2000).<br />

The sun’s movement across the sky<br />

gives orientation signals that vary with<br />

time of day. Migratory birds gain information<br />

from the sun better than<br />

humans, detecting changing patterns<br />

of polarization. Using magnetism <strong>and</strong><br />

celestial rotation together can increase<br />

reliability (birds constantly cross check<br />

them <strong>and</strong> adjust) (www.teara.govt.nz).<br />

The sun’s position can be analyzed by<br />

looking at the polarization pattern of<br />

the sky arising from sunlight scattering<br />

(useful when the sun is obscured).<br />

Many insects use celestial polarization<br />

for compass orientation by using<br />

polarization-sensitive photoreceptors.<br />

Biological research has shown that<br />

some underwater animals use polarization<br />

of light for navigation, communication,<br />

<strong>and</strong> hunting. Studies have<br />

been performed using polarization of<br />

scattered light underwater to improve<br />

visibility. Polarization of light in water<br />

is caused by refraction through the surface,<br />

scattering light in water, refraction<br />

by polarizing objects, <strong>and</strong> emission<br />

from polarized light sources (Karpel &<br />

Schecher, 2011).<br />

The Present <strong>and</strong><br />

Future of Biomimetic<br />

Underwater Robotics<br />

Rather than relying on the future<br />

development of a novel <strong>and</strong><br />

28 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 3<br />

Boston Engineering’s GhostSwimmer PH I<br />

AUV in field testing.<br />

ground-breaking power source, Boston<br />

Engineering’sASG,alongwithitsteammate,<br />

Olin College Intelligent Vehicles<br />

Laboratory (Needham, MA), is taking<br />

inspiration from a comparably sized<br />

biological system to develop an AUV<br />

optimized for complex underwater<br />

environments.<br />

Thanks to funding from the ONR<br />

<strong>and</strong> a large internal R&D effort by<br />

both Boston Engineering <strong>and</strong> Olin<br />

College, this new generation of AUVs<br />

is emerging. By leveraging past research,<br />

the latest technologies, <strong>and</strong> innovative<br />

rapid prototyping techniques,<br />

the team developed an AUV that could<br />

autonomously swim like a fish in only<br />

6 months (Figure 3). The team is currently<br />

in Phase II with ONR <strong>and</strong> expects<br />

to showcase its next generation AUV in<br />

late 2011 to demonstrate advancements<br />

in efficiency <strong>and</strong> maneuverability while<br />

addressing <strong>and</strong> investigating the other<br />

challenges facing underwater vehicles<br />

as described above.<br />

While there have been many attempts<br />

to duplicate the swimming actions<br />

of a fish (Xianzhong DAI, 2003;<br />

Valdivia et al., 2006), a breakthrough<br />

achieved by the Boston Engineering<br />

research team has been in producing<br />

these movements efficiently within a<br />

tactically relevant vehicle. After proving<br />

that fish-like oscillating foil propulsion<br />

could indeed have higher<br />

propulsive efficiency <strong>and</strong> could be<br />

achieved by man-made methods, the<br />

next challenge is to properly control<br />

<strong>and</strong> coordinate the movements. The<br />

intendedsolutionforthiscontrolis<br />

itself bioinspired. The result is an<br />

AUV platform that could outperform<br />

conventional technologies deployed<br />

today in both endurance <strong>and</strong> mobility.<br />

This work holds the potential to provide<br />

a paradigm shift in underwater<br />

AUV capability <strong>and</strong> usefulness.<br />

The trend of mimicking biology<br />

in engineering certainly is not new,<br />

but through the efforts of the various<br />

parties working in the field of biologically<br />

inspired engineering <strong>and</strong> biological<br />

mechanics, a reduction in the<br />

number of limitations in enabling<br />

technologies <strong>and</strong> knowledge is being<br />

seen. Recent advances in the underst<strong>and</strong>ing<br />

of biological hydrodynamics,<br />

sensing, navigation techniques, <strong>and</strong><br />

control coupled with low-power highthroughput<br />

computing (including<br />

Quad Core Processors, www.intel.<br />

com), advances in material science<br />

(ionomers <strong>and</strong> self healing materials,<br />

www.bimat.org), <strong>and</strong> improved prototyping<br />

techniques (such as direct metal<br />

laser sintering, www.morristech.com)<br />

has enabled robotics engineers as<br />

never before. As industry <strong>and</strong> academia<br />

continue to make progress in demonstrating<br />

how effective biomimetic designs<br />

can be, the more developers of<br />

next generation tactical systems such<br />

as Boston Engineering’s Advanced<br />

Systems Group will be able to rise to<br />

the technical challenges presented by<br />

current adopters of unmanned technology,<br />

especially in the underwater<br />

space.<br />

Authors:<br />

Michael Rufo <strong>and</strong> Mark Smithers<br />

Advanced Systems Group,<br />

Boston Engineering Corporation<br />

411 Waverley Oaks Road,<br />

Waltham, MA 02452<br />

Emails: mrufo@boston-engineering.<br />

com; msmithers@boston-engineering.<br />

com<br />

References<br />

Azarsinal, F., Bose, N., & Seif, M.S. 2007.<br />

An underwater vehicle maneuvering simulation:<br />

Focus on turning maneuvers. J Ocean<br />

Technol. 2(1):54-73.<br />

Bar-Cohen, Y. 2006. <strong>Biomimetics</strong>: Biologically<br />

Inspired Technologies. Boca Raton,<br />

FL: CRC Press. 469 pp.<br />

Barrett, D.S. 1996. Propulsive efficiency of<br />

a flexible hull underwater vehicle. Ph.D.<br />

thesis, Massachusetts Institute of <strong>Technology</strong>.<br />

Bee Navigation. http://www.physics.ohiostate.edu/~wilkins/writing/Samples/shortmed/<br />

fiskemedium. Accessed June 20, 2011.<br />

Bird Migration. Navigating by the stars<br />

<strong>and</strong> sun. http://www.teara.govt.nz/<br />

EarthSeaAndSky/BirdsOfSeaAndShore/<br />

BirdMigration/7/en. Accessed June 20, 2011.<br />

Butler, L. 1987. Underwater Radio Communication.<br />

http://www.qsl.net/vk5br/<br />

UwaterComms.htm. Accessed June 20, 2011.<br />

Dickson, K.A., & Graham, J.B. 2004.<br />

Evolution <strong>and</strong> consequences of endothermy<br />

in fishes. Physiol Biochem Zool. 77(6):<br />

998-1018. doi: 10.1086/423743.<br />

Fuller, J. Top 10 bungled attempts at oneperson<br />

flight. http://science.howstuffworks.com/<br />

transport/flight/classic/ten-bungled-flightattempt4.htm.<br />

Accessed March 1, 2011.<br />

Fulton, J.T. 2010. Dolphin biosonar echolocation<br />

A case study. http://neuronresearch.net/<br />

July/August 2011 Volume 45 Number 4 29


hearing/pdf/Dolphin_biosonar_echolocation.<br />

pdf. Accessed June 20, 2011.<br />

Gedney, L. 1984. Do Salmon navigate by<br />

the earth’s magnetic field Article #691.<br />

http://www.gi.alaska.edu/ScienceForum/<br />

ASF6/691.html. Accessed June 20, 2011.<br />

Goldenberg, F. 2006. Geomagnetic navigation<br />

beyond the magnetic compass. Goodrich<br />

Corporation, Advanced Sensors Technical<br />

Center, 14300 Judicial Road, Burnsville,<br />

MN 55306. http://ieeexplore.ieee.org/xpl/<br />

freeabs_all.jsparnumber=1650662. Accessed<br />

June 20, 2011.<br />

Graham, J.B., & Dickson, K.A. 2004.<br />

Tuna comparative physiology. J Exp Biol.<br />

207(iii):4015-24. doi: 10.1242/jeb.01267.<br />

Huntsberger, T. 2001. Biologically inspired<br />

autonomous rover control autonomous<br />

robots. Auton Robot. 11:341-6.<br />

doi: 10.1023/A:1012467829785.<br />

Intel. Intel core quad processors. http://www.<br />

intel.com/products/processor/core2quad/.<br />

Accessed June 20, 2011.<br />

Johnsen, S., & Lohmann, K. 2005. The<br />

physics <strong>and</strong> neurobiology of magnetoreception.<br />

Nature Reviews Neuroscience. 6:703-12.<br />

doi: 10.1038/nrn1745.<br />

Kalmijn, J. 2000. Detection <strong>and</strong> processing<br />

of electromagnetic <strong>and</strong> near-field acoustic<br />

signals in elasmobranch fishes. Philos T R<br />

Soc Lond B. 355(1401):1135-41.<br />

doi: 10.1098/rstb.2000.0654.<br />

Karpel, N., & Schechner, Y. 2011. Portable<br />

polarimetric underwater imaging system with<br />

a linear response. Dept. of Electrical Eng.,<br />

Technion - Israel Institute of <strong>Technology</strong>,<br />

Haifa 32000, Israel. http://webee.technion.<br />

ac.il/~yoav/publications/polarimetric.pdf.<br />

Accessed June 20, 2011.<br />

Lipson, H. Associate Professor. Cornell<br />

University Mechanical <strong>and</strong> Aerospace Engineering.<br />

http://www.mae.cornell.edu/lipson/.<br />

Accessed June 20, 2011.<br />

Lohmann, K.J., Lohmann, C.M.F., Llewellyn,<br />

M., Ehrhart, L.M., Bagley, D.A., & Swing, T.<br />

2004. Animal behavior: Geomagnetic map<br />

used in sea-turtle navigation. Nature.<br />

428:909-10. doi: 10.1038/428909a.<br />

Magnuson, J.J. 1978. Locomotion by scombrid<br />

fishes: hydromechanics, morphology,<br />

<strong>and</strong> behavior. Fish Physiol. 7:239-313.<br />

doi: 10.1016/S1546-5098(08)60166-1.<br />

Martiny, N. 2009. Design of a lateral-line<br />

sensor for an autonomous underwater vehicle.<br />

In: 8th IFAC International Conference on<br />

Maneuvering <strong>and</strong> Control of <strong>Marine</strong> Craft,<br />

MCMC 2009. pp. 292-7. Sao Paulo, Brazil:<br />

IFAC.<br />

Mitchell, B. Wireless St<strong>and</strong>ards-802.11b<br />

802.11a 802.11g <strong>and</strong> 802.11n: The 802.11<br />

family explained. http://compnetworking.<br />

about.com/cs/wireless80211/a/aa80211st<strong>and</strong>ard.<br />

htm. Accessed June 20, 2011.<br />

Morris Technologies. Direct metal laser sintering<br />

(DMLS). http://www.morristech.com/<br />

dmls.asw. Accessed June 20, 2011.<br />

Office of Naval Research. Bio-inspired<br />

autonomous systems. http://www.onr.<br />

navy.mil/en/Media-Center/Fact-Sheets/<br />

Bio-Inspired-Autonomous-Systems.aspx.<br />

Accessed June 20, 2011.<br />

Paluska, D., & Herr, H. 2006. Series<br />

elasticity <strong>and</strong> actuator power output. In:<br />

Proceedings of the 2006 IEEE International<br />

Conference on Robotics <strong>and</strong> Automation.<br />

Orl<strong>and</strong>o, FL: IEEE.<br />

Teledyne benthos acoustic modems.<br />

http://www.benthos.com/acoustic-telesonarmodem-subsea-ATM885.asp.<br />

Accessed<br />

June 20, 2011.<br />

The Navy Unmanned Undersea Vehicle<br />

(UUV) Master Plan. November 9, 2004.<br />

Thomopoulos, S., & Braught, G. 1995. A<br />

biologically inspired mechanism of machine<br />

perception <strong>and</strong> intelligent control for multirobot<br />

coordination. Penn State University.<br />

Triantafyllou, M.S., Techet, A.H., Zhu, Q.,<br />

Beal, D.N., Hover, F.S., & Yue, D.K.P.<br />

2002. Vorticity control in fish-like propulsion<br />

<strong>and</strong> maneuvering. Integr Comp Biol.<br />

42(5):1026-31. doi: 10.1093/icb/42.5.1026.<br />

Triantafyllou, M.S., & Triantafyllou, G.S.<br />

1993. Propulsion through oscillating foils in<br />

nature <strong>and</strong> technology. J Fluids Struct.<br />

7:205-24. doi: 10.1006/jfls.1993.1012.<br />

Valdivia y Alvarado, P., & Youcef-Toumi, K.<br />

2006. Design of machines with compliant<br />

bodies for biomimetic locomotion in liquid<br />

environments. J Dyn Syst-T ASME.<br />

128(1,3):11.<br />

Wernli, R. 2001. Recent U.S. Underwater<br />

Vehicle Projects. San Diego, CA: Space <strong>and</strong><br />

Naval Warfare Systems Center.<br />

Wernli, R. 2011. Recent US Navy Underwater<br />

Vehicle Projects. http://www.nmri.go.<br />

jp/main/cooperation/ujnr/24ujnr_paper_us/<br />

Ocean_Engineering_<strong>and</strong>_Resource/OERD_<br />

Wernli.pdf. Accessed June 20, 2011.<br />

Xianzhong, D. 2003. Some new progress of<br />

advanced robotics in China. In: Report to JCF<br />

2003 of IARP, Xianzhong DAI Robotics<br />

Expert Group of National High-Tech R&D<br />

Programme, China. Madrid, Spain.<br />

Zhai, J. 2007. Geomagnetic sensor based on<br />

giant magnetoelectric effect. Appl Phys Lett.<br />

91(12):1-3. doi: 10.1063/1.2789391.<br />

30 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Autonomous Robotic Fish as Mobile Sensor<br />

Platforms: Challenges <strong>and</strong> Potential Solutions<br />

AUTHOR<br />

Xiaobo Tan<br />

Smart Microsystems Laboratory,<br />

Department of Electrical <strong>and</strong><br />

Computer Engineering,<br />

Michigan State University<br />

Introduction<br />

ABSTRACT<br />

With advances in actuation <strong>and</strong> sensing materials <strong>and</strong> devices, there is a growing<br />

interest in developing underwater robots that propel <strong>and</strong> maneuver themselves<br />

as real fish do. Such robots, often known as robotic fish, could provide an engineering<br />

tool for underst<strong>and</strong>ing fish swimming. Equipped with communication<br />

capabilities <strong>and</strong> sensors, they could also serve as economical, dynamic samplers<br />

of aquatic environments. In this paper we discuss some of the major challenges in<br />

realizing adaptive, cost-effective, mobile sensor networks that are enabled by<br />

resource-constrained robotic fish. Such challenges include maneuvering in the<br />

presence of ambient disturbances, localization with adequate precision, sustained<br />

operation with minimal human interference, <strong>and</strong> cooperative control <strong>and</strong> sensing<br />

under communication constraints. We also present potential solutions <strong>and</strong> promising<br />

research directions for addressing these challenges, some of which are inspired<br />

by how fish solve similar problems.<br />

Keywords: robotic fish, adaptive sampling, mobile sensing platforms, aquatic<br />

sensor networks, water quality monitoring<br />

With 500 million years of evolution,<br />

fish <strong>and</strong> other aquatic animals<br />

areendowedwithavarietyofmorphological<br />

<strong>and</strong> structural features<br />

that enable them to move through<br />

water with speed, efficiency, <strong>and</strong> agility<br />

(Lauder & Drucker, 2004; Fish &<br />

Lauder, 2006). The remarkable feats<br />

in biological swimming have stimulated<br />

extensive theoretical, experimental,<br />

<strong>and</strong> computational research<br />

by biologists, mathematicians, <strong>and</strong><br />

engineers, in an effort to underst<strong>and</strong><br />

<strong>and</strong> mimic locomotion, maneuvering,<br />

<strong>and</strong> sensing mechanisms adopted by<br />

aquatic animals.<br />

Over the past two decades, there<br />

has also been significant interest in<br />

developing underwater robots that<br />

propel <strong>and</strong> maneuver themselves<br />

like real fish do (Triantafyllou &<br />

Triantafyllou, 1995; Kato, 2000;<br />

Anderson & Chhabra, 2002; Alvarado<br />

& Youcef-Toumi, 2006; Hu et al.,<br />

2006; Low, 2006; Epstein et al.,<br />

2006; Morgansen et al., 2007; Lauder<br />

et al., 2007; Chen et al., 2010; Aureli<br />

et al., 2010; Smithers, 2011). Often<br />

termed robotic fish, these robots provide<br />

an experimental platform for<br />

studying fish swimming <strong>and</strong> hold<br />

strong promise for a number of underwater<br />

applications. Instead of using<br />

propellers, robotic fish accomplish<br />

swimming by deforming the body<br />

<strong>and</strong>/or fin-like appendages, mostly<br />

functioning as caudal fins <strong>and</strong> sometimes<br />

as pectoral fins. Body deformation<br />

<strong>and</strong> fin movements are typically<br />

achieved with motors. On the other<br />

h<strong>and</strong>, advances in smart materials<br />

have been explored to actuate robotic<br />

fish in a noiseless <strong>and</strong> compact way<br />

(Paquette & Kim, 2004; Tangorra<br />

et al., 2007; Chen et al., 2010; Aureli<br />

et al., 2010). Robotic fish produce<br />

wake signatures similar to those of<br />

real fish <strong>and</strong> are thus less detectable<br />

than propeller-driven underwater vehicles,<br />

which is an important advantage<br />

in applications requiring stealth.<br />

Recent advances in computing,<br />

communication, electronics, <strong>and</strong><br />

materials have made it possible to<br />

create untethered robotic fish with<br />

onboard power, control, navigation,<br />

wireless communication, <strong>and</strong> sensing<br />

modules, which turns these robots<br />

into mobile sensing platforms in<br />

aquatic environments. Schools of robotic<br />

fish can form wireless sensor<br />

networks, which will have numerous<br />

promising applications, such as monitoring<br />

water quality, tracing oil spills,<br />

<strong>and</strong> patrolling harbors <strong>and</strong> coasts.<br />

Figure 1a shows a prototype of a robotic<br />

fish swimming in an inl<strong>and</strong><br />

lake. Figure 1b shows the close-up of<br />

another prototype, equipped with a<br />

dissolved oxygen (DO) sensor, global<br />

positioning system (GPS), <strong>and</strong> other<br />

electronic components, which has<br />

been developed for monitoring the<br />

DO level in aquafarms. Collected<br />

DO information will then be used to<br />

control the aerators to maintain a<br />

healthy environment for the aquatic<br />

animals on the farm.<br />

Autonomous robotic fish schools<br />

will provide a competitive alternative<br />

July/August 2011 Volume 45 Number 4 31


FIGURE 1<br />

Prototypes of autonomous robotic fish developed by the Smart Microsystems Laboratory at<br />

Michigan State University: (a) testing in a lake <strong>and</strong> (b) prototype for dynamically monitoring<br />

the DO level in aquafarms.<br />

to existing sensing technologies for<br />

aquatic <strong>and</strong> marine environments.<br />

Manual sampling, sometimes boat or<br />

ship-based, is still a common practice<br />

in environmental monitoring, which<br />

is labor-intensive with difficulty in<br />

capturing dynamic phenomena of<br />

interest. In-situ sensing with fixed or<br />

buoyedsensorsorverticalprofilers is<br />

another approach (Doherty et al.,<br />

1999; Reynolds-Fleming et al.,<br />

2002). However, these sensors have<br />

little freedom to move laterally, <strong>and</strong> it<br />

would require prohibitively many<br />

units for capturing distributed, spatially<br />

inhomogeneous information. The past<br />

decade has seen great progress in the use<br />

of robotic technology in aquatic sensing.<br />

Autonomous underwater vehicles<br />

(AUVs) (B<strong>and</strong>yopadhyay, 2005), for<br />

example, are being used for hydrographic<br />

survey, fishery operations, <strong>and</strong><br />

environmental monitoring (Hydroid,<br />

2009). Another highly successful technology<br />

is autonomous sea gliders,<br />

which has remarkable duration for<br />

continuous field operation because<br />

of highly energy-efficient design<br />

(Ericksen et al., 2001; Sherman et al.,<br />

2001; Webb et al., 2001; Rudnick<br />

et al., 2004). The downside for both<br />

AUVs <strong>and</strong> gliders is their cost, starting<br />

at US $50,000 per unit (not including<br />

the cost of sensors), prohibiting the<br />

deployment of many of them for<br />

observing with high spatial resolution,<br />

<strong>and</strong> excluding them from many applications<br />

(such as aquafarm monitoring)<br />

where cost is critical. The size (meters<br />

long) <strong>and</strong> weight (at the order of 50 kg)<br />

of these vehicles also make them cumbersome<br />

to h<strong>and</strong>le by a single person.<br />

Small autonomous robotic fish<br />

have the potential to address many of<br />

the aforementioned challenges. By a<br />

small robotic fish, we mean one that<br />

has length of 50 cm or less, displaces<br />

volume of up to 5 liters, <strong>and</strong> costs no<br />

more than US $5,000 (excluding that<br />

of aquatic sensors to be mounted). Its<br />

low cost, compact size, <strong>and</strong> light<br />

weight would make it affordable <strong>and</strong><br />

convenient to deploy these robots in<br />

groups for versatile applications <strong>and</strong><br />

various environments, such as ponds,<br />

lakes, rivers, <strong>and</strong> even oceans. Schools<br />

of robotic fish could form dynamic,<br />

adaptive sensor networks <strong>and</strong> provide<br />

distributed sensing coverage with desired<br />

spatiotemporal resolution.<br />

The realization of such a vision,<br />

however, is faced with a myriad of<br />

challenges. The size <strong>and</strong> cost considerations<br />

put stringent constraints on<br />

the robot’s locomotion, battery, computing,<br />

<strong>and</strong> communication capacities.<br />

The wide adoption of the<br />

robotic fish-based sensing technology<br />

will hinge on the robots’ ability to operate<br />

robustly in the unfriendly <strong>and</strong><br />

often unpredictable environment,<br />

with their limited onboard resources<br />

<strong>and</strong> with minimal human intervention.<br />

This poses challenges across a<br />

wide spectrum, ranging from locomotion<br />

<strong>and</strong> maneuvering mechanisms, to<br />

energy-efficient designs, localization<br />

<strong>and</strong> communication schemes, <strong>and</strong><br />

control <strong>and</strong> coordination strategies,<br />

to name a few. In this paper, we outline<br />

some of the most critical challenges<br />

<strong>and</strong> discuss potential approaches or<br />

opportunities in research <strong>and</strong> technology<br />

advancement for addressing the<br />

challenges.<br />

Maneuvering in<br />

Uncertain Environment<br />

As a sensor platform, the robotic<br />

fish often needs to survey a given<br />

path or hover over a particular region<br />

in the presence of ambient disturbances<br />

caused by wind, waves, currents,<br />

<strong>and</strong> turbulences. Regardless of<br />

its propulsion mechanism, however, a<br />

small robotic fish has limited actuation<br />

authority to counteract the disturbances.<br />

It is thus of great interest to<br />

be able to sense the flow <strong>and</strong> react in<br />

the most effective way under the actuation<br />

constraints. We can look to live<br />

fish for inspiration, because they deal<br />

with this very problem on a regular<br />

basis <strong>and</strong> have developed intricate<br />

sensing <strong>and</strong> actuation systems that<br />

offer us interesting insight.<br />

Artificial Lateral Line<br />

Most fish use the lateral line system<br />

as an important sensory organ to probe<br />

their environment (Coombs, 2001).<br />

A lateral line consists of arrays of so<br />

32 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


called neuromasts, each containing<br />

bundles of sensory hairs, encapsulated<br />

in a gelatinous structure called cupula.<br />

Under an impinging flow, the hairs are<br />

deflected, which elicits firing of the<br />

hair cell neurons <strong>and</strong> thus enables the<br />

animal to sense the flow field, perform<br />

hydrodynamic imaging, <strong>and</strong> identify<br />

new field objects of interest. The lateral<br />

line system plays an important role in<br />

various fish behaviors, including prey/<br />

predator detection, schooling, rheotaxis,<br />

courtship <strong>and</strong> communication.<br />

A lateral line-like sensory module<br />

or an artificial lateral line will be very<br />

useful for a robotic fish to improve its<br />

maneuverability. For example, with<br />

feedback from the lateral line, the<br />

robot could manipulate vortices in<br />

the flow with its actuated fins <strong>and</strong> exploit<br />

the ambient flow energy for locomotion<br />

(Beal et al., 2006) or perform<br />

station-keeping by responding appropriately<br />

to the sensed ambient flow.<br />

Artificial lateral line systems, where arrays<br />

of beam or hair-like structures are<br />

used to measure flow velocities, have<br />

been proposed based on various physical<br />

transduction principles, including<br />

hot wire anemometry (Yang et al.,<br />

2006), piezoresistivity (Yang et al.,<br />

2010), capacitive sensing (Dagamseh<br />

et al., 2010), <strong>and</strong> encapsulated interface<br />

bilayers (Sarles et al., 2011).<br />

Recently, we have exploited the intrinsic<br />

mechanosensory property of<br />

ionic polymer-metal composites<br />

(IPMCs) to construct artificial lateral<br />

lines (Abdulsadda & Tan, 2011;<br />

Abdulsadda et al., 2011). As illustrated<br />

in Figure 2, an IPMC consists of three<br />

layers, with an ion-exchange polymer<br />

membrane (e.g., Nafion) s<strong>and</strong>wiched<br />

by metal electrodes. Inside the polymer,<br />

(negatively charged) anions covalently<br />

fixed to polymer chains are balanced<br />

by mobile (positively charged) cations.<br />

An applied mechanical stimulus, such<br />

FIGURE 2<br />

Illustration of the IPMC sensing principle.<br />

as a flow impinging on the IPMC,<br />

redistributes the cations inside <strong>and</strong><br />

produces a detectable electrical signal<br />

(typically open-circuit voltage or<br />

short-circuit current) that is correlated<br />

with the mechanical or hydrodynamic<br />

stimulus (Chen et al., 2007). Conversely,<br />

an applied voltage across an<br />

IPMC leads to the transport of cations<br />

<strong>and</strong> accompanying solvent molecules,<br />

resulting in both differential swelling<br />

<strong>and</strong> electrostatic forces inside the<br />

FIGURE 3<br />

material, which cause the material to<br />

bend <strong>and</strong> hence the actuation effect<br />

(Shahinpoor & Kim, 2001). Figure 3a<br />

shows a prototype of an artificial lateral<br />

line consisting of four IPMC sensors.<br />

While the physical construction of<br />

robust <strong>and</strong> sensitive artificial lateral<br />

lines remains an active research area,<br />

it is of equal importance to make<br />

sense out of the data collected by the<br />

lateral line. Existing studies on biological<br />

<strong>and</strong> artificial lateral lines have<br />

Experimental results on localization of a dipole source with unknown location <strong>and</strong> vibration amplitude:<br />

(a) prototype of IPMC-based lateral line, consisting of four IPMC sensors, <strong>and</strong> (b) localization<br />

results along three different tracks, based on solving a model-based nonlinear estimation<br />

problem (Abdulsadda et al., 2011).<br />

July/August 2011 Volume 45 Number 4 33


mostly focused on the problem of<br />

localizing a vibrating sphere, known<br />

as a dipole, which is used to emulate periodic<br />

tail beating or other appendage<br />

movement of aquatic animals. Several<br />

approaches to signal processing have<br />

been reported, which include exploitation<br />

of the characteristic points (e.g.,<br />

zero-crossings, maxima, etc.) in the<br />

measured velocity profile (Dagamseh<br />

et al., 2010), matching of the measured<br />

data with preobtained templates<br />

(P<strong>and</strong>ya et al., 2006), beamforming<br />

techniques (Yang et al., 2010), <strong>and</strong><br />

artificial neural networks (Abdulsadda<br />

& Tan, 2011). We have further considered<br />

a source localization problem<br />

where both the source location <strong>and</strong><br />

its vibrating amplitude are unknown.<br />

The posed problem is interesting,<br />

since a source far away but with large<br />

vibration could produce a signal that<br />

has similar amplitude as a signal produced<br />

by a source nearby but with<br />

small vibration. By formulating <strong>and</strong><br />

solving a nonlinear estimation problem<br />

based on an analytical model for<br />

dipole-generated flow, we are able to<br />

resolve both the source location <strong>and</strong><br />

the vibration amplitude simultaneously<br />

(Abdulsadda et al., 2011). As shown in<br />

Figure 3b, experimental results on an<br />

IPMC-based lateral line prototype<br />

(Figure 3a) have confirmed the effectiveness<br />

of the model-based estimation<br />

approach.<br />

Other than the dipole source localization<br />

problem, there are a few<br />

interesting directions for the signal<br />

processing of artificial lateral lines.<br />

The first is the detection <strong>and</strong> localization<br />

of multiple, more sophisticated<br />

moving sources (including vortices).<br />

With the sources moving, the resultingflow<br />

is no longer at a steady state, <strong>and</strong><br />

the processing algorithm needs to<br />

localize the sources with minimal<br />

latency. Another major problem to<br />

consider is the information processing<br />

for a lateral line that is mounted on a<br />

robotic fish, where the motion of the<br />

robot itself <strong>and</strong> its fins adds significant<br />

“noise” to the lateral line signal. Biological<br />

fish deal with these problems<br />

effectively through biomechanical filtering<br />

for enhanced signal-to-noise<br />

ratio <strong>and</strong> through dynamic filtering<br />

in the central nervous system to remove<br />

the unwanted signal components<br />

(Coombs & Braun, 2003; Bodznick<br />

et al., 2003). For example, dynamic<br />

neural mechanisms have been identified<br />

for suppressing self-generated<br />

noise (Coombs & Braun, 2003). Such<br />

biological insight will prove valuable in<br />

devising the mechanical, electrical, <strong>and</strong><br />

digital filtering mechanisms for solving<br />

complex processing problems faced by<br />

artificial lateral lines.<br />

Bioinspired Fin<br />

Achieving high-maneuverability<br />

hinges on the ability to manipulate<br />

the fluid in a delicate manner. Fish<br />

often use their pectoral fins to perform<br />

sophisticated maneuvers (Drucker &<br />

Lauder, 2001, 2003). These maneuvers<br />

involve complex conformational<br />

changes of the fins, involving cupping,<br />

twisting, <strong>and</strong> bending motions.<br />

Robotic fish fins, on the other h<strong>and</strong>,<br />

often use rigid foils (Kato, 2000;<br />

Morgansen et al., 2007). Recently,<br />

advances in soft actuation materials,<br />

e.g., IPMCs, have led to the exploration<br />

of these materials as flexible propulsors<br />

(Paquette & Kim, 2004).<br />

However, the resulting robotic fins<br />

typically have simple deformation<br />

modes, e.g., bending only (Chen et al.,<br />

2010; Aureli et al., 2010), <strong>and</strong> fall<br />

short of emulating the complex deformation<br />

of biological fins.<br />

Underst<strong>and</strong>ing of the morphology<br />

<strong>and</strong> mechanics of fish fins has spurred<br />

effort on mimicking these features<br />

(typically at a higher level) in designing<br />

robotic fins (Lauder et al., 2007). In<br />

particular, the complex shape change<br />

of fish fins is enabled by multiple<br />

muscle-controlled, relatively rigid, bony<br />

fin rays that are connected via collagenous<br />

membrane (Lauder & Madden,<br />

2006). Coordinated movement of individual<br />

fin rays results in conformational<br />

changes of the fins desired in<br />

maneuvers. On the engineering side,<br />

by patterning electrodes of IPMC<br />

materials, one can expect to produce<br />

complex deformation by applying different<br />

voltage inputs to different electrode<br />

areas. The patterning can be<br />

achieved with masking during electroless<br />

plating or by selective removal of<br />

electrodes post-IPMC fabrication<br />

using laser or machining (Kim et al.,<br />

2011). Inspired by the pectoral fins<br />

of bluegill sunfish, we have developed<br />

a lithography-based monolithic fabrication<br />

process for creating IPMC actuators<br />

capable of sophisticated shape<br />

changes (Chen & Tan, 2010). As<br />

shown in Figure 4, the fabricated sampleconsistsofmultipleactiveIPMC<br />

regions, coupled through much thinner<br />

passive regions. By phasing the voltage<br />

inputs to different active regions, we<br />

can realize various deformation modes<br />

including bending, twisting, <strong>and</strong> cupping<br />

(Figure 5). For example, a peak-topeak<br />

twisting angle of 16° is achieved<br />

with actuation voltages of 3 V (Chen<br />

& Tan, 2010).<br />

While the progress made in biomimetic<br />

fins is encouraging, significant<br />

further advances in both material<br />

fabrication <strong>and</strong> fin control are needed,<br />

before robotic fish are capable of manipulating<br />

the flow in a manner close<br />

to what their biological counterparts<br />

do. In particular, the materials need<br />

to be improved so that they can produce<br />

much larger deformation with<br />

reasonable b<strong>and</strong>width (a few Hz). On<br />

34 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 4<br />

Monolithically fabricated IPMC sample inspired by fish fins: (a) top view <strong>and</strong> (b) SEM picture of the<br />

cross section, showing that the passive area is much thinner than the active area (Chen & Tan, 2010).<br />

the control side, we need to model <strong>and</strong><br />

underst<strong>and</strong> the deformation <strong>and</strong> its<br />

hydrodynamic consequences of a given<br />

input by combining observation of<br />

kinetic patterns of fish fin movement,<br />

nonlinear elasticity modeling, computational<br />

fluid dynamics modeling,<br />

<strong>and</strong> experimental flow measurements<br />

using digital particle image velocimetry.<br />

Energy-Efficient<br />

Sustained Operation<br />

For the robotic fish-based sensing<br />

technology to gain widespread adoption,<br />

these robots will have to be able<br />

FIGURE 5<br />

to work continuously in the field<br />

with minimal human intervention. In<br />

particular, they need to operate for at<br />

least weeks, if not for months, before<br />

returning for battery recharge <strong>and</strong><br />

other manual maintenance. Power is<br />

arguably the most crucial factor that<br />

limits the operational time. While<br />

fuel represents a potential energy<br />

source with high power <strong>and</strong> energy<br />

density, it is unclear when fuel-based<br />

propulsion will become feasible for<br />

small underwater robots. Therefore,<br />

battery is expected to be the primary<br />

power source for robotic fish, for at<br />

least the next 5–10 years.<br />

Examples of deformation modes demonstrated by the fabricated IPMC fin: (a) bending <strong>and</strong><br />

(b) twisting.<br />

Thereareanumberofwaysone<br />

can potentially extend the run time<br />

of battery-powered robotic fish. For<br />

example, many onboard devices can<br />

be put to the sleep mode to save energy,<br />

when they are not active. Photovoltaic<br />

films can be mounted on the<br />

robot to harvest solar energy <strong>and</strong> replenish<br />

the battery, when the robot is<br />

on the water surface. Wave energy<br />

could be another source to tap into,<br />

but how to harvest it on an untethered<br />

<strong>and</strong> often goal-oriented robotic fish remains<br />

a challenge.<br />

Whilealltheaforementionedapproaches<br />

could stretch the mileage<br />

per battery charge to some extent,<br />

they are not game-changers. Design<br />

of energy-efficient locomotion mechanisms<br />

will be critical in realizing<br />

long-duration field operation, since<br />

locomotion is the biggest source of<br />

energy expenditure for autonomous<br />

robotic fish. To this end, we are currently<br />

developing a novel class of underwater<br />

robots, called gliding robotic<br />

fish (Figure 6a). Such a robot will represent<br />

a hybrid of underwater glider<br />

<strong>and</strong> robotic fish; for example, it will<br />

have wings for gliding <strong>and</strong> fins for maneuvering<br />

<strong>and</strong> assistive propulsion.<br />

Consequently, a gliding robotic fish<br />

is expected to possess both high energy<br />

efficiency <strong>and</strong> great maneuverability.<br />

Figure 6b further illustrates the<br />

gliding principle <strong>and</strong> why a gliding robotic<br />

fish will be energy-efficient.<br />

Under the combined influence of gravity<br />

<strong>and</strong> buoyancy, the body will experience<br />

vertical (up or down) motion.<br />

When the glider is properly pitched,<br />

the lift generated during buoyancyinduced<br />

vertical motion will enable<br />

horizontal travel. Through the control<br />

of pitch direction <strong>and</strong> buoyancy, one<br />

can switch between the descent/ascent<br />

gliding motion, resulting in a sawtoothshaped<br />

trajectory. Since buoyancy<br />

July/August 2011 Volume 45 Number 4 35


FIGURE 6<br />

Energy-efficient gliding robotic fish: (a) the concept of a gliding robotic fish with a hydrodynamic<br />

gliding body <strong>and</strong> a caudal fin <strong>and</strong> (b) illustration of the gliding principle.<br />

control <strong>and</strong> pitch control are the major<br />

sources of energy expenditure <strong>and</strong> take<br />

placeonlyduringascent/descent<br />

switching, the motion is very energyefficient,<br />

especially if the dive depth<br />

is relatively large.<br />

Communication <strong>and</strong><br />

Localization<br />

Robotic fish need to communicate<br />

with a base station to receive comm<strong>and</strong>s<br />

<strong>and</strong> send back the collected environmental<br />

information. They also<br />

need to communicate with each<br />

other for information relay <strong>and</strong> motion<br />

coordination. Underwater communication,<br />

however, is particularly challenging<br />

for small robotic fish that<br />

have stringent power <strong>and</strong> size constraints.<br />

Radio frequency (RF) signals<br />

attenuate quickly in water, severely<br />

limiting the achievable communication<br />

range <strong>and</strong> data rate. Light communication<br />

is possible (Verzijlenberg<br />

& Jenkin, 2010), but again the range<br />

<strong>and</strong> data rate are very limited <strong>and</strong> it<br />

does not work in a turbid environment.<br />

Acoustic <strong>and</strong> sonar communication<br />

underwater has been studied for<br />

many decades <strong>and</strong> was recently explored<br />

for communication among robotic<br />

fish (Science Daily, 2008).<br />

However, the associated power <strong>and</strong><br />

hardware required to achieve reasonably<br />

large communication distance<br />

<strong>and</strong> data rate are typically not affordable<br />

by small robotic fish. For these<br />

reasons, the most viable solution<br />

wouldbetocommunicatewhenthe<br />

robot surfaces, in which case lowpower,<br />

low-cost RF communication<br />

protocols such as ZigBee can be readily<br />

used. Unlike the Bluetooth protocol,<br />

which is intended for eliminating cables<br />

between electronic devices, the ZigBee<br />

protocol is built on top of the IEEE<br />

802.15.4 st<strong>and</strong>ard <strong>and</strong> it targets specifically<br />

wireless sensor network applications.<br />

For wider range communication,<br />

cellular networks could also be employed<br />

if such networks are available.<br />

Limiting the communication to the<br />

water surface entails additional challenges<br />

in robotic fish coordination,<br />

control, <strong>and</strong> networking. For effective<br />

FIGURE 7<br />

networking, we need to have a sufficient<br />

number of nodes on the surface.<br />

This can be achieved through joint<br />

motion planning <strong>and</strong> control. For example,<br />

we can hold robotic fish on<br />

the surface until a network with adequate<br />

density <strong>and</strong> coverage is formed<br />

<strong>and</strong> completes data transmission.<br />

Localization is another challenge in<br />

robotic fish-based sensor networks.<br />

For small robotic fish, having onboard<br />

localization capability is essential for<br />

successful navigation of the robot <strong>and</strong><br />

for effective coordination of robotic<br />

fish networks. Accurate localization is<br />

also critical for tagging the sensed information<br />

so that the data collected<br />

by robotic fish are associated correctly<br />

to the physical location in water. While<br />

the GPS is readily available <strong>and</strong> does<br />

not take up much space, its typical precision<br />

of 5–10 m is inadequate for<br />

many applications of robotic fish due<br />

to their small size <strong>and</strong> relatively low<br />

speeds (50 cm/s or less). In addition,<br />

theGPSmaytakeafewminutesto<br />

lock satellites every time the robot<br />

emerges from underwater, which severely<br />

limits the networking <strong>and</strong> control<br />

performance. More agile <strong>and</strong> precise<br />

localization technology is needed.<br />

We have developed an efficient localization<br />

scheme for small robotic fish<br />

(Shatara & Tan, 2010). As illustrated<br />

in Figure 7a, the scheme is based on<br />

Underwater acoustic ranging-based localization: (a) schematic of the ranging protocol <strong>and</strong><br />

(b) localization performance in a pool test (Shatara & Tan, 2010).<br />

36 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


acoustic ranging, which measures the<br />

time it takes an acoustic signal to travel<br />

from one node to the other. For example,<br />

node 1 simultaneously sends an<br />

RF packet <strong>and</strong> an acoustic pulse to<br />

node 2. When node 2 starts its onboard<br />

timer when it receives the RF<br />

packet <strong>and</strong> then stops its timer when<br />

it detects the acoustic pulse. Since the<br />

RF signal travels much faster than<br />

the acoustic signal, we can estimate<br />

thedistancebetweenthetwonodes<br />

based on the timer reading. The<br />

scheme involves simple hardware, a<br />

buzzer <strong>and</strong> a microphone, for each<br />

node. A sliding discrete Fourier transform<br />

algorithm, implemented on a<br />

digital signal controller, is employed<br />

for the detection of arrival of the acoustic<br />

signal. Figure 7b shows the results<br />

from experiments in a swimming<br />

pool, where a small robotic fish was<br />

towed across the deep side of the<br />

pool (about 13 m long) while its distances<br />

to the two beacon nodes<br />

mounted on the pool wall were measured<br />

through acoustic ranging. The<br />

resulting localization error was less<br />

than 1 m for the entire tested range,<br />

which was a significant improvement<br />

over the precision of a commercial<br />

GPS.<br />

Note that the above localization<br />

scheme works only when the robot<br />

surfaces, since it involves RF communication.<br />

The location of a robot when<br />

it is underwater can be inferred using<br />

dead reckoning. The scheme in<br />

Figure 7 does require beacon nodes<br />

(whose locations are known) to obtain<br />

the absolute location of a node. In the<br />

absence of such beacon nodes, the<br />

scheme can be used to get relative locations<br />

among nodes, which is of interest<br />

in coordinating schools of robotic fish.<br />

There are many other in-air localization<br />

schemes for wireless sensor networks,<br />

e.g., ranging based on received<br />

signal strength. While these schemes<br />

can be adapted for robotic fish-based<br />

aquatic networks, care must be taken<br />

to address the challenges associated<br />

with noises, disturbances, <strong>and</strong> signal<br />

attenuation at the air/water interface.<br />

Autonomous Control<br />

<strong>and</strong> Coordination<br />

With onboard communication,<br />

navigation, control, <strong>and</strong> sensing devices,<br />

robotic fish are desired to operate<br />

autonomously, as individuals <strong>and</strong> as<br />

schools, to carry out envisioned monitoring<br />

tasks. A few challenges arise in<br />

the control <strong>and</strong> coordination of these<br />

robots. A robotic fish needs to h<strong>and</strong>le<br />

multiple functions subject to environmental<br />

uncertainties <strong>and</strong> resource constraints.<br />

In particular, the functions<br />

could include sampling the environment,<br />

processing <strong>and</strong> transmitting<br />

the measured data, maintaining network<br />

connectivity, <strong>and</strong> controlling<br />

its motion. There are various uncertainties<br />

that interfere with these functions,<br />

examples of which include<br />

motion perturbations due to waves<br />

<strong>and</strong> turbulences, imperfect sensor<br />

measurements, <strong>and</strong> localization error<br />

<strong>and</strong> communication packet drops.<br />

Furthermore, all of these functions<br />

compete for limited onboard computing<br />

<strong>and</strong> power resources. This is a classic<br />

multi-objective, multi-constraint<br />

optimization problem, <strong>and</strong> it dem<strong>and</strong>s<br />

asystematicapproachtothejoint<br />

consideration of control, networking,<br />

<strong>and</strong> sensor fusion. Evolutionary algorithms<br />

(Deb, 2001), which codify<br />

basic principles of genetic evolution,<br />

can offer a promising solution to this<br />

multi-objective optimization problem.<br />

Another challenge lies in coordinating<br />

a school of robotic fish. It is intriguing<br />

to deploy groups of robotic<br />

fish that cooperatively perform sensing<br />

tasks. In that case, it is often desirable<br />

not to use centralized control, because<br />

the centralized paradigm would entail<br />

prohibitive cost in communication,<br />

<strong>and</strong> it would paralyze the whole network<br />

if the comm<strong>and</strong> node fails.<br />

Therefore, individual robotic fish are<br />

expected to communicate only with<br />

their local neighbors <strong>and</strong> make decisions<br />

in a distributed manner. Animals<br />

including fish often exhibit coordinated<br />

collective movement facilitated<br />

by only local interactions, which has<br />

inspired great interest from the controls<br />

community in analyzing <strong>and</strong> synthesizing<br />

control laws for groups of<br />

unmanned vehicles. Significant progress<br />

has been made in this area, even<br />

with some demonstrated success in<br />

adaptive sampling using underwater<br />

gliders (Leonard et al., 2007).<br />

While these accomplishments can<br />

provide a sound starting point for the<br />

control <strong>and</strong> coordination of robotic<br />

fish schools, we need to recognize<br />

many new <strong>and</strong> subtle difficulties<br />

faced by the latter. For example, a robotic<br />

fish can only communicate with<br />

its peers when it surfaces, which renders<br />

communication <strong>and</strong> feedback<br />

intermittent <strong>and</strong> asynchronous. This<br />

again points to the need to jointly consider<br />

control, communication, <strong>and</strong><br />

networking issues.<br />

Other Challenges<br />

As sensor platforms, the potential<br />

of robotic fish in environmental sensing<br />

will be ultimately limited by the<br />

availability of versatile sensors that are<br />

compact <strong>and</strong> easy to interface with.<br />

Most commercial sensors available<br />

today are not amenable to integration<br />

into small robotic fish, since sensor<br />

manufacturers have mostly been targeting<br />

h<strong>and</strong>held, fixed, or buoyed<br />

July/August 2011 Volume 45 Number 4 37


sensors where miniaturization is not<br />

critical. It is expected that, with the development<br />

of robotic fish <strong>and</strong> wireless<br />

networking technologies, manufacturers<br />

will see the growth opportunities<br />

in robotic fish-enabled aquatic<br />

sensing <strong>and</strong> start investing in the development<br />

of compact, economical, <strong>and</strong><br />

robust aquatic sensors.<br />

There are other engineering challenge,<br />

one example of which is biofouling,<br />

where microorganisms <strong>and</strong><br />

other organisms accumulate on the<br />

surface of robotic fish <strong>and</strong> their sensors,<br />

degrading movement <strong>and</strong> sensing<br />

performance. Periodic cleaning is<br />

an option; thanks to the mobility of<br />

robotic fish, access to these robots is<br />

relatively easy. Another possibility is<br />

to apply anti-fouling coatings.<br />

Conclusion<br />

In this paper, we have explored the<br />

potential of small robotic fish as<br />

mobile sensor platforms for aquatic<br />

<strong>and</strong> marine environments. Realization<br />

of this vision poses a rich set of challenges<br />

across a wide spectrum of areas,<br />

such as actuation/sensing materials,<br />

mechanism design, communication,<br />

control, <strong>and</strong> packaging. We have reviewed<br />

some of the major challenges<br />

<strong>and</strong> discussed possible routes to overcome<br />

them. The list of challenges outlined<br />

in this paper is by no means<br />

exhaustive, but even partial success<br />

in addressing them could have farreaching<br />

impact on aquatic environmental<br />

monitoring <strong>and</strong> other<br />

engineering applications.<br />

Acknowledgment<br />

This work was supported in part by<br />

the Office of Naval Research under<br />

grant N000140810640 (program<br />

manager Dr. T. McKenna) <strong>and</strong> the<br />

National Science Foundation under<br />

grants ECCS 0547131, CCF<br />

0820220, EEC 0908810, IIS<br />

0916720, DBI 0939454, ECCS<br />

1050236, <strong>and</strong> ECCS 1029683. The<br />

author gratefully acknowledges the<br />

contributions of many former <strong>and</strong><br />

current members of the Smart Microsystems<br />

Laboratory at Michigan State<br />

University, for the results <strong>and</strong> ideas<br />

presented in this paper.<br />

Author:<br />

Xiaobo Tan<br />

Smart Microsystems Laboratory<br />

Department of Electrical <strong>and</strong><br />

Computer Engineering<br />

Michigan State University,<br />

East Lansing, MI 48824<br />

Email: xbtan@egr.msu.edu<br />

References<br />

Abdulsadda, A.T., & Tan, X. 2011.<br />

Underwater source localization using an<br />

IPMC-based artificial lateral line. In:<br />

Proceedings of the 2011 IEEE International<br />

Conference on Robotics <strong>and</strong> Automation,<br />

to appear. 2719-24. Shanghai, China: IEEE.<br />

Abdulsadda, A.T., Zhang, F., & Tan, X.<br />

2011. Localization of source with unknown<br />

amplitude using IPMC sensor arrays. In:<br />

Proceedings of Electroactive Polymer Actuators<br />

<strong>and</strong> Devices (EAPAD) XIII. Proc. SPIE, eds.<br />

Bar-Cohen, Y., & Carpi, F., 797627(1-11).<br />

San Diego, CA: SPIE or <strong>Society</strong> of Photo-<br />

Optical Instrumentation Engineers.<br />

Alvarado, P.V., & Youcef-Toumi, K. 2006.<br />

Design of machines with compliant bodies<br />

for biomimetic locomotion in liquid environments.<br />

Trans ASME J Dyn Syst Meas<br />

Control. 128:3-13. doi: 10.1115/1.2168476.<br />

Anderson, J.M., & Chhabra, N.K. 2002.<br />

Maneuvering <strong>and</strong> stability performance of a<br />

robotic tuna. Integr Comp Biol. 42:118-126.<br />

doi: 10.1093/icb/42.1.118.<br />

Aureli, M., Kopman, V., & Porfiri, M. 2010.<br />

Free-locomotion of underwater vehicles<br />

actuated by ionic polymer metal composites.<br />

IEEE/ASME Trans Mech. 15(4):603-14.<br />

doi: 10.1109/TMECH.2009.2030887.<br />

B<strong>and</strong>yopadhyay, P.R. 2005. Trends in<br />

biorobotic autonomous undersea vehicles.<br />

IEEE J Oceanic Eng. 30:109-39.<br />

doi: 10.1109/JOE.2005.843748.<br />

Beal, D.N., Hover, F.S., Triantafyllou,<br />

M.S., Liao, J.C., & Lauder, G.V. 2006.<br />

Passive propulsion in vortex wakes.<br />

J Fluid Mech. 549:385-402. doi: 10.1017/<br />

S0022112005007925.<br />

Bodznick, D., Montgomery, J., & Tricas,<br />

T.C. 2003. Electroreception: Extracting<br />

behaviorally important signals from noise.<br />

In: Sensory Processing in Aquatic Environments,<br />

eds. Collin, S.P., & Marshall, N.J.,<br />

389-403. New York: Springer-Verlag.<br />

Chen, Z., Shatara, S., & Tan, X. 2010.<br />

Modeling of biomimetic robotic fish propelled<br />

by an ionic polymer-metal composite<br />

caudal fin. IEEE/ASME Trans Mech. 15(3):<br />

448-459. doi: 10.1109/TMECH.2009.<br />

2027812.<br />

Chen, Z., & Tan, X. 2010. Monolithic<br />

fabrication of ionic polymer-metal composite<br />

actuators capable of complex deformation.<br />

Sensor Actuat A: Phys. 157:246-57.<br />

doi: 10.1016/j.sna.2009.11.024.<br />

Chen, Z, Tan, X., Will, A., & Ziel, C. 2007.<br />

A dynamic model for ionic polymer-metal<br />

composite sensors. Smart Mater Struct. 16:<br />

1477-88. doi: 10.1088/0964-1726/16/4/063.<br />

Coombs, S. 2001. Smart skins: Information<br />

processing by lateral line flow sensors.<br />

Auton Robot. 11:255-61. doi: 10.1023/<br />

A:1012491007495.<br />

Coombs, S., & Braun, C.B. 2003. Information<br />

processing by the lateral line system.<br />

In: Sensory Processing in Aquatic Environments,<br />

eds. Collin, S.P., & Marshall, N.J.,<br />

122-38. New York: Springer-Verlag.<br />

Dagamseh, A., Lammerink, T., Kolster, M.,<br />

Bruinink, C., Wiegerink, R., & Krijnen, G.<br />

38 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


2010. Dipole-source localization using biomimetic<br />

flowsensor arrays positioned as<br />

lateral-line system. Sensor Actuat A. 162(2):<br />

355-60. doi: 10.1016/j.sna.2010.02.016.<br />

Deb, K. 2001. Multi-objective Optimization<br />

using Evolutionary Algorithms. Chichester,<br />

UK: John Wiley & Sons.<br />

Doherty, K.W., Frye, D.E., Liberatore,<br />

S.P., & Toole, J.M. 1999. A moored profiling<br />

instrument. J Atmos Ocean Technol.<br />

16:1816-29. 2.0.CO;2"target=_blankdoi:-<br />

10.1175/1520-0426(1999)0162.0.CO;2.<br />

Drucker, E.G., & Lauder, G.V. 2001.<br />

Wake dynamics <strong>and</strong> fluid forces of turning<br />

maneuvers in sunfish. J Exp Biol. 204:431-42.<br />

Drucker, E.G., & Lauder, G.V. 2003.<br />

Function of pectoral fins in rainbow trout:<br />

behavioral repertoire <strong>and</strong> hydrodynamic<br />

forces. J Exp Biol. 206:813-26. doi: 10.1242/<br />

jeb.00139.<br />

Epstein, M., Colgate, J.E., & MacIver,<br />

M.A. 2006. Generating thrust with a<br />

biologically-inspired robotic ribbon fin.<br />

In: Proceedings of the 2006 IEEE/RSJ<br />

International Conference on Intelligent<br />

Robots <strong>and</strong> Systems. 2412-17. Beijing, China:<br />

IEEE. doi: 10.1109/IROS.2006.281681.<br />

Ericksen, C.C., Osse, T.J., Light, R.D.,<br />

Wen, T., Lehman, T.W., Sabin, P.L., …<br />

Chiodi, A.M. 2001. Seaglider: A long-range<br />

autonomous underwater vehicle for oceanographic<br />

research. IEEE J Oceanic Eng. 26:<br />

424-36. doi: 10.1109/48.972073.<br />

Fish, F.E., & Lauder, G.V. 2006. Passive<br />

<strong>and</strong> active flow control by swimming fishes<br />

<strong>and</strong> mammals. Annu Rev Fluid Mech.<br />

38:193-224. doi: 10.1146/annurev.<br />

fluid.38.050304.092201.<br />

Hu, H., Liu, J., Dukes, I., & Francis, G.<br />

2006. Design of 3D swim patterns for<br />

autonomous robotic fish. In: Proc. 2006<br />

IEEE/RSJ Int. Conf. Intell. Robots Syst.<br />

2406-11. Beijing, China: IEEE.<br />

Hydroid, LLC. 2009. REMUS: Autonomous<br />

underwater vehicles - REMUS AUVs.<br />

Available at http://www.km.kongsberg.<br />

com/ks/web/nokb0240.nsf/AllWeb/<br />

D5682F98CBFBC05AC1257497002976E4<br />

OpenDocument/ (accessed 5 July 2011).<br />

Kato, N. 2000. Control performance in<br />

the horizontal plane of a fish robot with<br />

mechanical pectoral fins. IEEE J Oceanic<br />

Eng. 25(1):121-9. doi: 10.1109/48.820744.<br />

Kim, K.J., Pugal, D., & Leang, K.K. 2011.<br />

A twistable ionic polymer-metal composite<br />

artificial muscle for marine applications.<br />

Mar Technol Soc J. This issue.<br />

Lauder, G.V., Anderson, E.J., Tangorra, J.,<br />

& Madden, P.G.A. 2007. Fish biorobotics:<br />

Kinematics <strong>and</strong> hydrodynamics of selfpropulsion.<br />

J Exp Biol. 210:2767-80.<br />

doi: 10.1242/jeb.000265.<br />

Lauder, G.V., & Drucker, E.G. 2004.<br />

Morphology <strong>and</strong> experimental hydrodynamics<br />

of fish fin control surfaces. IEEE J Oceanic<br />

Eng. 29(3):556-71. doi: 10.1109/JOE.<br />

2004.833219.<br />

Lauder, G.V., & Madden, P.G.A. 2006.<br />

Learning from fish: Kinematics <strong>and</strong> experimental<br />

hydrodynamics for roboticists. Int<br />

J Automat Comput. 4:325-35. doi: 10.1007/<br />

s11633-006-0325-0.<br />

Leonard, N.E., Paley, D.A., Lekien, F.,<br />

Sepulchre, R., Fratantoni, D.M., & Davis,<br />

R.E. 2007. Collective motion, sensor<br />

networks, <strong>and</strong> ocean sampling. Proc IEEE.<br />

95(1):48-74. doi: 10.1109/JPROC.2006.<br />

887295.<br />

Low, K.H. 2006. Locomotion <strong>and</strong> depth control<br />

of robotic fish with modular undulating<br />

fins. Int J Automat Comput. 4:348-57.<br />

doi: 10.1007/s11633-006-0348-6.<br />

Morgansen, K.A., Triplett, B.I., & Klein,<br />

D.J. 2007. Geometric methods for modeling<br />

<strong>and</strong> control of free-swimming fin-actuated<br />

underwater vehicles. IEEE T Robot. 23(6):<br />

1184-99. doi: 10.1109/LED.2007.911625.<br />

P<strong>and</strong>ya, S., Yang, Y., Jones, D.L., Engel, J.,<br />

& Liu, C. 2006. Multisensor processing<br />

algorithms for underwater dipole localization<br />

<strong>and</strong> tracking using MEMS artificial<br />

lateral-line sensors. EURASIP J Appl Signal<br />

Processing. 2006:76593. doi: 10.1155/<br />

ASP/2006/76593.<br />

Paquette, J.W., & Kim, K.J. 2004. Ionomeric<br />

electroactive polymer artificial muscle for<br />

naval applications. IEEE J Oceanic Eng. 29(3):<br />

729-37. doi: 10.1109/JOE.2004.833132.<br />

Reynolds-Fleming, J.V., Fleming, J.G., &<br />

Luettich, R.A. 2002. Portable autonomous<br />

vertical profiler for estuarine applications.<br />

Estuaries. 25:142-7. doi: 10.1007/BF02696058.<br />

Rudnick, D.L., Eriksen, C.C., Fratantoni,<br />

D.M., & Perry, M.J. 2004. Underwater gliders<br />

for ocean research. Mar Technol Soc J. 38:<br />

48-59. doi: 10.4031/002533204787522703.<br />

Sarles, S.A., & Leo, D.J. 2011. Hair cell<br />

sensing with encapsulated interface bilayers.<br />

In: Proceedings of Bioinspiration, <strong>Biomimetics</strong>,<br />

<strong>and</strong> Bioreplication. Proc. SPIE, eds. Martin-<br />

Palma, R.J., & Lakhtakia, A., 797509(1-11).<br />

San Diego, CA: SPIE.<br />

Science Daily. 2008. School of robofish communicate<br />

with each other in underwater robot<br />

teams. Science Daily. http://www.sciencedaily.<br />

com/releases/2008/06/080606105454.htm<br />

(accessed 5 July 2011).<br />

Shahinpoor, M., & Kim, K. 2001. Ionic<br />

polymer-metal composites: I. Fundamentals.<br />

Smart Mater Struct. 10:819-33. doi: 10.1088/<br />

0964-1726/10/4/327.<br />

Shatara, S., & Tan, X. 2010. An efficient,<br />

time-of-flight-based underwater acoustic<br />

ranging system for small robotic fish. IEEE<br />

J Oceanic Eng. 35(4):837-46. doi: 10.1109/<br />

JOE.2010.2060810.<br />

Sherman, J., Davis, R.E., Owens, W.B., &<br />

Valdes, J. 2001. The autonomous underwater<br />

glider “spray”. IEEE J Oceanic Eng.<br />

26:437-46. doi: 10.1109/48.972076.<br />

Smithers, M. 2011. Designing future<br />

autonomous underwater vehicles. Defense<br />

Tech Briefs. Available at http://www.<br />

defensetechbriefs.com/component/content/<br />

article/9075 (accessed 5 July 2011).<br />

Tangorra, J., Anquetil, P., Fofonoff, T.,<br />

Chen, A., Del Zio, M., & Hunter, I. 2007.<br />

July/August 2011 Volume 45 Number 4 39


The application of conducting polymers to a<br />

biorobotic fin propulsor. Bioinspir Biomim.<br />

2:S6-17. doi: 10.1088/1748-3182/2/2/S02.<br />

Triantafyllou, M.S., & Triantafyllou, G.S.<br />

1995. An efficient swimming machine.<br />

Sci Amer. 272:64-71. doi: 10.1038/<br />

scientificamerican0395-64.<br />

Verzijlenberg, B., & Jenkin, M. 2010.<br />

Swimming with robots: Human robot<br />

communication at depth. In: Proceedings of<br />

the 2010 IEEE/RSJ International Conference<br />

on Intelligent Robots <strong>and</strong> Systems.<br />

4023-28. Taipei, Taiwan: IEEE. doi: 10.1109/<br />

IROS.2010.5652751.<br />

Webb, D.C., Simonetti, P.J., & Jones, C.P.<br />

2001. “SLOCUM”: An underwater glider<br />

propelled by environmental energy. IEEE J<br />

Oceanic Eng. 26:447-52. doi: 10.1109/<br />

48.972077.<br />

Yang, Y., Chen, J., Engel, J., P<strong>and</strong>ya, S.,<br />

Chen, N., Tucker, C., … Liu, C. 2006.<br />

Distant touch hydrodynamic imaging with<br />

an artificial lateral line. Proc Nat Acad<br />

Sci. 103(50):18891-5. doi: 10.1073/<br />

pnas.0609274103.<br />

Yang, Y., Nguyen, N., Chen, N., Lockwood,<br />

M., Tucker, C., Hu, H., … Jones, D.L. 2010.<br />

Artificial lateral line with biomimetic neuromasts<br />

to emulate fish sensing. Bioinspir<br />

Biomim. 5:016 001.<br />

40 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Robotic Models for Studying Undulatory<br />

Locomotion in Fishes<br />

AUTHORS<br />

George V. Lauder<br />

Jeanette Lim<br />

Ryan Shelton<br />

Museum of Comparative Zoology,<br />

Harvard University<br />

Chuck Witt<br />

Erik Anderson<br />

Department of Mechanical<br />

Engineering, Grove City College<br />

James L. Tangorra<br />

Department of Mechanical<br />

Engineering, Drexel University<br />

Introduction<br />

Fish moving through the water are<br />

capable of using a variety of locomotor<br />

modes. Some species swim nearly exclusively<br />

using their fins <strong>and</strong> generate<br />

propulsive <strong>and</strong> maneuvering forces<br />

using midline fins (dorsal <strong>and</strong> anal)<br />

or paired fins (pectoral <strong>and</strong> pelvic)<br />

(Drucker & Lauder, 1999, 2001;<br />

Hove et al., 2001; St<strong>and</strong>en, 2008;<br />

St<strong>and</strong>en & Lauder, 2007). Other<br />

species generate thrust primarily by<br />

activating body musculature to bend<br />

the body <strong>and</strong> generate waves passing<br />

from the head toward the tail (Gillis,<br />

1996; Jayne & Lauder, 1994, 1995c;<br />

Lauder & Tytell, 2006; Rome et al.,<br />

1993). The caudal or tail fin is generally<br />

considered as an extension of the<br />

body in most analyses of fish locomotion,<br />

but the tail also possesses a<br />

substantial array of intrinsic muscles<br />

that appear to stiffen the tail during<br />

steady forward swimming <strong>and</strong> generate<br />

a variety of complex tail conformations<br />

during maneuvering (Flammang<br />

ABSTRACT<br />

Many fish swim using body undulations to generate thrust <strong>and</strong> maneuver in<br />

three dimensions. The pattern of body bending during steady rectilinear locomotion<br />

has similar general characteristics in many fishes <strong>and</strong> involves a wave of increasing<br />

amplitude passing from the head region toward the tail. While great progress has<br />

been made in underst<strong>and</strong>ing the mechanics of undulatory propulsion in fishes, the<br />

inability to control <strong>and</strong> precisely alter individual parameters such as oscillation frequency,<br />

body shape, <strong>and</strong> body stiffness, <strong>and</strong> the difficulty of measuring forces on<br />

freely swimming fishes have greatly hampered our ability to underst<strong>and</strong> the fundamental<br />

mechanics of the undulatory mode of locomotion in aquatic systems. In this<br />

paper, we present the use of a robotic flapping foil apparatus that allows these parameters<br />

to be individually altered <strong>and</strong> forces measured on self-propelling flapping<br />

flexible foils that produce a wave-like motion very similar to that of freely swimming<br />

fishes. We use this robotic device to explore the effects of changing swimming<br />

speed, foil length, <strong>and</strong> foil-trailing edge shape on locomotor hydrodynamics, the<br />

cost of transport, <strong>and</strong> the shape of the undulating foil during locomotion. We<br />

also examine the passive swimming capabilities of a freshly dead fish body. Finally,<br />

we model fin-fin interactions in fishes using dual-flapping foils <strong>and</strong> show that thrust<br />

can be enhanced under correct conditions of foil phasing <strong>and</strong> spacing as a result of<br />

the downstream foil making use of vortical energy released by the upstream foil.<br />

Keywords: fish, robot, swimming, biomechanics<br />

& Lauder, 2008; Flammang &<br />

Lauder, 2009).<br />

Despite the large number of studies<br />

of fish undulatory locomotion<br />

over the last 20 years, there are still<br />

many unanswered questions about<br />

how the pattern of body deformation<br />

is generated, the effect of body stiffness<br />

on locomotor performance, what<br />

effect different tail shapes have on locomotor<br />

function, <strong>and</strong> how hydrodynamic<br />

interactions among different<br />

fins might influence the generation<br />

of swimming forces. Studies of freely<br />

swimming fishes have contributed<br />

enormously to our underst<strong>and</strong>ing of<br />

the mechanics of aquatic locomotion,<br />

but such an approach is necessarily<br />

limited to the behaviors voluntarily<br />

executed by living fishes. And measuring<br />

locomotor forces on freely<br />

swimming fishes is a challenging proposition.<br />

Furthermore, a wide array of<br />

interesting experimental manipulations,<br />

including changing the flexural<br />

stiffness of the body, altering the<br />

shape of the tail, <strong>and</strong> changing the<br />

spacing between fins, are clearly impossible<br />

to conduct in living animals.<br />

Robotics offers a complementary<br />

approach to studies of living fishes by<br />

allowing manipulation of variety of<br />

parameters such as flexural stiffness,<br />

aspect ratio, tail shape, <strong>and</strong> spacing between<br />

adjacent fins. A simple robotic<br />

flapping foil device can be used to generate<br />

undulatory locomotion in flexible<br />

fish-like materials, <strong>and</strong> forces can<br />

July/August 2011 Volume 45 Number 4 41


e measured <strong>and</strong> the effect of changes<br />

in body length, stiffness <strong>and</strong> tail shape<br />

can be quantified.<br />

The focus of this paper will be on<br />

undulatory locomotion in the water<br />

<strong>and</strong> the use of a robotic flapping apparatus<br />

to produce swimming in both<br />

flexible plastic foils <strong>and</strong> a passive<br />

freshly dead fish body. After a brief<br />

overview of undulatory locomotion<br />

in fishes, we discuss the design of a robotic<br />

controller for flapping flexible<br />

foils that allows measurement of selfpropelled<br />

speeds (SPS), forces, <strong>and</strong><br />

torques <strong>and</strong> the cost of transport associated<br />

with foils of different shapes <strong>and</strong><br />

stiffnesses. In addition, we address<br />

an important technical issue in studies<br />

of aquatic propulsion: the effect of<br />

swimming at non-SPS on body waveform,<br />

patterns of force production <strong>and</strong><br />

wake flow patterns. Finally, we present<br />

data on the swimming performance<br />

of passive fish bodies <strong>and</strong> discuss the<br />

future for studies of robotically controlled<br />

undulatory locomotion. Our<br />

overall aim is to introduce a number<br />

of case studies with data that show<br />

the utility of this approach for studying<br />

underwater propulsion <strong>and</strong> to present<br />

an overview of how such studies can<br />

provide new ideas <strong>and</strong> tests for current<br />

views of how fish swim.<br />

Overview of Undulatory<br />

Propulsion in Fishes<br />

When fish swim using their bodies<br />

<strong>and</strong> tail fin as the primary thrust generators,<br />

they pass a wave of bending<br />

down the body that increases in amplitude<br />

from the head toward the tail<br />

(Donley & Dickson, 2000; Gillis,<br />

1996; Jayne & Lauder, 1995a, 1995b;<br />

Liao, 2002; Long et al., 1994). This<br />

bending wave is produced by a wave<br />

of muscular activity that also moves<br />

posteriorly <strong>and</strong> is created by spinal<br />

cord <strong>and</strong> hindbrain pattern generators<br />

(Bone et al., 1978; Fetcho & Svoboda,<br />

1993; Fetcho, 1986; Shadwick &<br />

Gemballa, 2006). Muscular power<br />

to generate thrust is produced primarily<br />

by the segmented myotomal<br />

musculature in the posterior region of<br />

fish, especially during slow swimming<br />

(Johnson et al., 1994; Rome et al.,<br />

1993; Syme, 2006), <strong>and</strong> as speed increases<br />

more anterior body muscles<br />

are recruited to power locomotion.<br />

The division of the segmented body<br />

musculature into superficial “red” fibers<br />

<strong>and</strong> deeper <strong>and</strong> more complexly<br />

arranged “white” fibers also plays an<br />

important role in underst<strong>and</strong>ing the<br />

multiple “gaits” used by fishes; the relative<br />

roles of these two types of muscle<br />

fibers can change dramatically as fish<br />

change swimming speed <strong>and</strong> execute<br />

rapid locomotor behaviors such as<br />

fast-start escapes (Jayne & Lauder,<br />

1993; Tytell & Lauder, 2002).<br />

When fish swim slowly using body<br />

undulations, oscillation of the front<br />

third or so of the body is quite small<br />

even in species as diverse as eels, trout,<br />

<strong>and</strong> tuna (Donley & Dickson, 2000;<br />

Gillis, 1998; Lauder & Tytell, 2006),<br />

<strong>and</strong> a primary role of this reduced oscillation<br />

appears to be drag reduction<br />

by minimizing the frontal area of the<br />

fish that encounters oncoming flow.<br />

As swimming speed increases, the<br />

front region of fish shows increasingly<br />

large side-to-side oscillations, which is<br />

in part a reflection of the recruitment<br />

of body musculature in this more anterior<br />

region of the fish.<br />

Even when fish swim by undulatory<br />

propulsion using the production<br />

of traveling waves, other fins often<br />

are used too, <strong>and</strong> it is incorrect to suggest<br />

that fish locomotor modes represent<br />

completely distinct patterns of<br />

motion. For example, when trout or<br />

bluegill sunfish swim using body undulations,<br />

they are also actively using<br />

their dorsal <strong>and</strong> anal fins, which play<br />

a key role in balancing roll torques<br />

<strong>and</strong> generating thrust (Drucker &<br />

Lauder, 2001, 2005; St<strong>and</strong>en &<br />

Lauder, 2005, 2007). In addition,<br />

the pelvic fins play an important role<br />

in controlling body stability during<br />

locomotion in fishes (Harris, 1936,<br />

1938; St<strong>and</strong>en, 2008, 2010). Fishes<br />

also vary in tail shape, <strong>and</strong> the distinction<br />

between the externally symmetrical<br />

(homocercal) tail of teleost fishes<br />

<strong>and</strong> the asymmetrical tail in sharks<br />

<strong>and</strong> fish such as sturgeon (Lauder,<br />

1989, 2000) is well known. The heterocercal<br />

tail shape induces torques around<br />

the body center of mass that requires<br />

compensatory changes in body position<br />

to allow steady horizontal swimming<br />

(Liao & Lauder, 2000; Wilga<br />

& Lauder, 2002, 2004b).<br />

Although considerable progress has<br />

been made in studies of fish locomotion<br />

by investigating the kinematics,<br />

muscle activity, <strong>and</strong> hydrodynamics<br />

of live fishes swimming steadily <strong>and</strong><br />

maneuvering, there are many limits<br />

to studies of this kind. Perhaps the<br />

two greatest limitations to research<br />

on live fishes are (1) the considerable<br />

difficulty in measuring locomotor<br />

forces <strong>and</strong> torques produced by the<br />

bending body as fish swim freely <strong>and</strong><br />

(2) the inability to manipulate key<br />

variables such as body length, aspect<br />

ratio, tail shape, <strong>and</strong> stiffness. Without<br />

an ability to alter such key components<br />

that govern locomotor dynamics, we<br />

will be limited in our underst<strong>and</strong>ing<br />

of the factors that influence undulatory<br />

propulsion in the water.<br />

The purpose of this paper is to present<br />

a number of new case studies<br />

using a robotic flapping device for<br />

generating undulatory locomotion in<br />

engineered materials as a means of<br />

42 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


etter underst<strong>and</strong>ing how fish swim<br />

<strong>and</strong> the dynamics of undulatory propulsion.<br />

We discuss different examples<br />

to show the utility of this approach <strong>and</strong><br />

how use of simple flexible foils informs<br />

studies of fish locomotion <strong>and</strong> points<br />

to new avenues of research.<br />

Simple Robotic Models of<br />

Fish Undulatory Locomotion<br />

We have designed a robotic apparatus<br />

that produces controlled heave <strong>and</strong><br />

pitch motions of flapping foils. The<br />

most important features of this device<br />

are (1) that it can be set up to be selfpropelling<br />

(allowing locomotion by<br />

flapping foils at their natural swimming<br />

speed <strong>and</strong> not only at imposed<br />

speeds) <strong>and</strong> (2) that forces <strong>and</strong> torques<br />

can be measured on flapping foils<br />

during self-propelled swimming so<br />

that within-cycle patterns of force <strong>and</strong><br />

torque oscillation can be compared<br />

among foils with different shapes<br />

<strong>and</strong> stiffnesses at different swimming<br />

speeds. This apparatus was designed<br />

with two sets of flapping foils in series<br />

in order to be able to model, with foils,<br />

the interactions that can occur between<br />

fins of fishes that are arranged<br />

in series such as the dorsal <strong>and</strong> anal<br />

fins <strong>and</strong> the tail fin (discussed further<br />

in the section on fin-fin interactions<br />

below).<br />

A general description of the first<br />

generation of this apparatus is presented<br />

in Lauder et al. (2007). Briefly,<br />

heave <strong>and</strong> pitch motors <strong>and</strong> rotary<br />

encoders that allow readouts of foil position<br />

are mounted on a carriage above<br />

a recirculating flow tank. This carriage<br />

is supported on low friction air bearings,<br />

which allow the carriage to move<br />

in response to foil thrust <strong>and</strong> drag forces<br />

generated during flapping motions.<br />

The second generation version of this<br />

apparatus has an ATI Nano-17 sixaxis<br />

force/torque sensor mounted on<br />

the shaft supporting the foils (Figure<br />

1A). This permits measurement<br />

of three forces <strong>and</strong> three torques during<br />

self-propulsion at sample rates that<br />

allow quantification of within-cycle<br />

patterns of force production even during<br />

self-propulsion. In addition, a linear<br />

encoder mounted on the carriage<br />

allows a readout of carriage position,<br />

FIGURE 1<br />

<strong>and</strong> this is used by a Labview program<br />

to calculate SPS from data generated<br />

by a series of swimming tests at a<br />

range of speeds. Synchronizing signals<br />

from the LabView program controlling<br />

the heave <strong>and</strong> pitch motors are<br />

used to trigger data acquisition from<br />

theATIsensor<strong>and</strong>alsototrigger<br />

image acquisition from three synchronized<br />

Photron high-speed video<br />

Images of a variety of flexible foils <strong>and</strong> freshly dead trout (Oncorhynchus mykiss) attached to a<br />

robotic flapping foil apparatus (see Lauder et al., 2007, for details on the basic design). (A) Flexible<br />

foil actuated at its leading edge in heave <strong>and</strong> pitch, suspended in a recirculating flow tank. The red<br />

arrow points to an ATI Nano-17 6-axis force/torque transducer on the foil shaft. (B, C) Flexible foils<br />

with different trailing edge shapes are used to study the effect of tail shape on swimming performance.<br />

(D, E, F) Images of a trout held behind the head to allow imposition of heave <strong>and</strong> pitch motions to<br />

study passive body properties. (E) An image from a trout self-propelling under an imposed heave<br />

motion; the body waveform produced is very similar to that generated during swimming.<br />

July/August 2011 Volume 45 Number 4 43


cameras. Images of the flapping foils to<br />

quantify both foil motion <strong>and</strong> hydrodynamic<br />

flow patterns are thus<br />

synchronized with force <strong>and</strong> torque<br />

measurements on the foil <strong>and</strong> with<br />

the imposed heave <strong>and</strong> pitch motions<br />

on the foil.<br />

The overall goals of using a flapping<br />

foil robotic device are to simplify<br />

<strong>and</strong> control as much as possible patterns<br />

of motion imposed on flexible<br />

<strong>and</strong> rigid foils that swim through the<br />

water <strong>and</strong> to allow direct testing of<br />

the effect on propulsion of a variety<br />

of key factors relevant to underst<strong>and</strong>ing<br />

fish locomotion: tail shape (Figures<br />

1B <strong>and</strong> 1C), foil flexibility, <strong>and</strong><br />

interactions among fins. This robotic<br />

apparatus can also be used to examine<br />

the passive swimming capabilities of a<br />

freshly dead fish body (Figures 1D, 1E,<br />

<strong>and</strong> 1F ), <strong>and</strong> below we present data on<br />

the ability of fish bodies to passively<br />

propel <strong>and</strong> generate propulsive waveforms<br />

under imposed motions. Foils<br />

of various kinds are attached to a stainless<br />

steel s<strong>and</strong>wich bar (to hold the<br />

leading edge of flexible materials; Figures<br />

1A, 1B, <strong>and</strong> 1C) to a solid 8-mm<br />

shaft (for rigid NACA 0012 foils; see<br />

Lauder et al., 2007, <strong>and</strong> data shown<br />

in Figure 11), <strong>and</strong> flexible fish bodies<br />

aremountedinaholderthatisattached<br />

behind the head (Figures 1D<br />

<strong>and</strong> 1F). Holding systems are designed<br />

not to flex in response to imposed<br />

heave <strong>and</strong> pitch motions <strong>and</strong> to allow<br />

forces <strong>and</strong> torques generated by swimming<br />

foils to be transmitted to the<br />

force/torque sensor on the shaft. Holding<br />

systems such as the s<strong>and</strong>wich bar<br />

system do have their own drag, <strong>and</strong><br />

this drag is time-dependent due to<br />

the heaving <strong>and</strong> pitching motion of<br />

the holding system as the foils selfpropel.<br />

It is thus not possible to give<br />

asinglevalueforthedragofthefoil<br />

holding system. Since all foil comparisons<br />

within a single experimental type<br />

used the same holding system <strong>and</strong> were<br />

treated identically, we do not present<br />

data on the performance of the holding<br />

apparatus alone.<br />

Sample data from a self-propelled<br />

flapping foil actuated in heave only<br />

are shown in Figure 2. Monitoring<br />

foil shaft position <strong>and</strong> measuring forces<br />

in the X (upstream-downstream) <strong>and</strong><br />

Y (side to side) directions allows calculation<br />

of the heave velocity <strong>and</strong> other<br />

derived quantities such as the instantaneous<br />

power required by the foil to<br />

swim <strong>and</strong> the coefficients of thrust<br />

<strong>and</strong> power. The cost of transport is calculated<br />

by measuring the foil mass <strong>and</strong><br />

dividing the cost/meter by this value<br />

for each foil (see Table 1). Typical<br />

peak thrust coefficients for highly flexible<br />

foils (flexural stiffness in the range<br />

of 10 −4 to 10 −6 Nm 2 ) are in the range<br />

of±0.2,lowerthantypicalforrigid<br />

foils, but these flexible foils nonetheless<br />

are capable of self-propulsion<br />

at speeds of 10-30 cm s −1 .<br />

FIGURE 2<br />

A key technical issue arises in studies<br />

of flapping foil propulsion that<br />

hope to imitate the self-propelled condition<br />

achieved by swimming fishes:<br />

foils that are not self-propelling may<br />

exhibit patterns of thrust oscillation<br />

that are not centered around zero.<br />

Any foil that is truly self-propelling<br />

(<strong>and</strong> not being dragged through the<br />

wateratspeedsslowerorfasterthan<br />

it would naturally move) should generate<br />

thrust in an oscillatory pattern, <strong>and</strong><br />

thrust integrated over a single flapping<br />

cycle should equal zero. Graphs in the<br />

literature of thrust coefficients during<br />

foil-based locomotion that are not<br />

centered around zero indicate that<br />

the foil was being towed above or<br />

below the SPS <strong>and</strong> do not reflect the<br />

self-propelled condition. Data from<br />

recordings of foil forces generated<br />

during self-propulsion <strong>and</strong> at speeds<br />

below <strong>and</strong> above the SPS are shown<br />

in Figure 3. During self-propelled<br />

swimming, the thrust coefficient has<br />

a mean of zero over each flapping<br />

Sample data from a self-propelling flexible plastic foil (flexural stiffness = 9.2 × 10 −5 Nm 2 ) actuated<br />

in heave at the leading edge at 2-Hz frequency. Heave position is monitored by rotary encoders,<br />

<strong>and</strong> X <strong>and</strong>Y forces are measured by a force transducer mounted on the foil shaft. From the data on<br />

foil position, force, <strong>and</strong> velocity, we make calculations of the instantaneous power, <strong>and</strong> dimensionless<br />

thrust <strong>and</strong> power coefficients. The dashed lines indicate zero for each trace.<br />

44 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


TABLE 1<br />

Locomotor properties of flexible plastic foils of three lengths while self-propelling.<br />

Foil length SPS (m/s) Reynolds number Strouhal number Work/cycle (mJ) Cost/meter (mJ/m) Cost of transport (mJ/g/m)<br />

20 0.105 21,114 0.831 2.248 42.59 106.47<br />

25 0.101 25,250 0.775 2.257 44.69 89.39<br />

35 0.091 31,850 0.349 2.233 49.08 70.11<br />

The three foils were made of the same material with a flexural stiffness of 3.1 × 10 −6 Nm 2 . These highly flexible foils propel at relatively high Strouhal numbers at<br />

shorter lengths.<br />

Reynolds <strong>and</strong> Strouhal numbers are dimensionless.<br />

Foils were actuated at their leading edge with ±1 cm heave, no pitch, at 2 Hz.<br />

Foil span was 6.8 cm for all three foils.<br />

St<strong>and</strong>ard errors for all parameters ranged from 0.3% to 1.5% of the mean values in each column.<br />

cycle. When foils are forced to swim<br />

above their SPS, the thrust coefficient<br />

curves shift above the zero baseline,<br />

<strong>and</strong> when forced swimming occurs at<br />

speeds below the SPS, data are shifted<br />

below the baseline.<br />

Change in foil swimming speed<br />

above <strong>and</strong> below the SPS can also<br />

have dramatic effects on the kinematics<br />

of the foil (Lauder et al., 2011) <strong>and</strong><br />

on the hydrodynamic wakes displayed<br />

by swimming foils. Figure 4 shows<br />

FIGURE 3<br />

the shape observed during swimming<br />

of a flexible foil (flexural stiffness,<br />

3.1 × 10 −6 Nm 2 )<strong>and</strong>thehydrodynamic<br />

wakes that result from swimming<br />

at, below, <strong>and</strong> above the average<br />

SPS. During swimming at an imposed<br />

speed below the SPS, the trailing edge<br />

ofthefoilhasalargeamplitude<strong>and</strong><br />

irregular motion that produces a wide<br />

bifurcating wake with separate momentum<br />

jets to each side (Figure 4A).<br />

At the SPS where the foil is allowed to<br />

Graph showing the dimensionless thrust coefficient versus time for a flexible plastic foil (flexural<br />

stiffness = 9.2 × 10 −5 Nm 2 ), 20 cm long, 6.8 cm high, actuated in heave ±1 cm at the leading<br />

edge at 2-Hz frequency. The red curve shows data for the self-propelled condition (18.8 cm/s),<br />

during which the thrust coefficient integrated over a single flapping cycle equals zero. Note that<br />

when experiments are done under non-self-propelling conditions (green <strong>and</strong> blue curves) the<br />

plots shift up or down so that the integrated coefficient over a flapping cycle is no longer zero.<br />

Data shown have been digitally filtered with a b<strong>and</strong>pass filter. Strouhal number for this experiment<br />

= 0.3 at the SPS (red curve). (Color versions of figures available online at: http://www.<br />

ingentaconnect.com/content/mts/mtsj/2011/00000045/00000004.)<br />

swim freely with no imposed constraints<br />

on speed, the foil bends into<br />

a regular ribbon-like pattern with<br />

a fish-likebodywake(seeNauen&<br />

Lauder, 2002a, 2002b) <strong>and</strong> alternating<br />

centers of vorticity with a fluid jet that<br />

me<strong>and</strong>ers in between these vortical<br />

centers (Figure 4B). Above the SPS<br />

where the external free-stream flow<br />

is increased above SPS <strong>and</strong> the foil is<br />

forced to swim against the increased<br />

flow, the foil shape exhibits large<br />

amplitude wave-like motion <strong>and</strong> a<br />

substantial drag-like wake with fluid<br />

velocities below that of the free stream<br />

in the wake (Figure 4C). These changing<br />

patterns of foil kinematics <strong>and</strong><br />

hydrodynamics during swimming<br />

that are not under conditions of selfpropulsion<br />

are most easily seen in highly<br />

flexible materials (flexural stiffnesses in<br />

the range of 10 −3 to 10 −6 Nm 2 ) where<br />

the fluid-structure interaction is most<br />

evident visually. These flexible materials<br />

which have flexural stiffnesses similar<br />

to those of fish (McHenry et al., 1995)<br />

show fish-like propulsion <strong>and</strong> deform<br />

into wave-like patterns with a fish-like<br />

wake (Figure 4B) when allowed to selfpropel<br />

in a low-friction system.<br />

The structure of the wake behind<br />

flapping flexible foils has a significant<br />

three-dimensional component due<br />

to the finite chord <strong>and</strong> span, <strong>and</strong> we<br />

July/August 2011 Volume 45 Number 4 45


FIGURE 4<br />

Hydrodynamics of propulsion in a flexible foil (flexural stiffness = 3.1 × 10 −6 Nm 2 ) swimming<br />

below, at, <strong>and</strong> above its SPS (8.55 cm/s). The foil was actuated at the leading edge with amplitude<br />

±0.5 cm, no pitch, at 3Hz, <strong>and</strong> is 20 cm long; Strouhal number at the SPS is 0.83 (see Table 1). The<br />

bottom margin of the foil is marked in white. Yellow arrows indicate water velocity; red color indicates<br />

counterclockwise vorticity; blue color indicates clockwise vorticity. The panels on the left<br />

show the whole foil swimming, while the matched panels on the right show a close-in view of the<br />

wake structure. The effect of swimming at a non SPS is dramatic, both on foil shape <strong>and</strong> on wake<br />

structure. In (A) the foil swam at an imposed speed of 4.75 cm/s, <strong>and</strong> in (C) the foil swam at an<br />

imposed speed of 25.7 cm/s.<br />

quantified this aspect of foil locomotor<br />

dynamics using the volumetric flow<br />

visualization system described by<br />

Flammang et al. (2011a, 2011b) <strong>and</strong><br />

Troolin <strong>and</strong> Longmire (2010). Figure<br />

5 shows how each of the centers<br />

of vorticity in the ribbon-like pattern<br />

shown in Figure 4B actually represent<br />

vortical columns on each side of the<br />

foil, which connect to each other above<br />

<strong>and</strong> below the foil. These inter-column<br />

connections occur both to upstream<br />

<strong>and</strong> downstream columns on the<br />

same side <strong>and</strong> also across the foil to<br />

vortical columns on the opposite side<br />

(Figure 5B). To date no other studies<br />

have provided three-dimensional<br />

wake snapshots during self-propulsion<br />

in highly flexible foils, but such studies<br />

in the future may reveal interesting<br />

patterns of wake interaction among<br />

components of the foil <strong>and</strong> may contribute<br />

to explaining the dynamics of<br />

propulsion in the flexing bodies.<br />

These results also suggest that at<br />

least some of the diversity of fish<br />

wakes reported in the literature results<br />

from situations in which the fishes<br />

were not swimming steadily, either<br />

in still water or against imposed<br />

flows, as wakes that look like those<br />

in Figures 4A <strong>and</strong> 4C are frequently<br />

presented. Fish commonly accelerate<br />

<strong>and</strong> execute small maneuvers during<br />

locomotion <strong>and</strong> considerable effort is<br />

needed to ensure that kinematics <strong>and</strong><br />

wake flow patterns are taken at moments<br />

when the fish is swimming<br />

steadily. Even small accelerations can<br />

substantially change fish wake flows<br />

(Tytell, 2004), <strong>and</strong> data from the<br />

flapping foil robot here illustrate that<br />

these kinematic <strong>and</strong> hydrodynamic alterations<br />

can be reproduced with flexible<br />

foils.<br />

One of the most straightforward<br />

questions that could be asked about<br />

undulatory propulsion <strong>and</strong> one that<br />

is difficult to study with live fishes is<br />

the effect of changing length alone on<br />

locomotor performance. We clearly cannot<br />

alter fish length experimentally<br />

<strong>and</strong> expect reasonable swimming performance,<br />

<strong>and</strong> comparing fish of<br />

different lengths (while useful for studies<br />

of scaling) does not account for the<br />

many other changes in the musculature<br />

<strong>and</strong> skeleton that occur as fish<br />

grow. Table 1 shows data obtained<br />

from three flexible foils of different<br />

46 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 5<br />

Volumetric flow visualization using the V3V technique (see Flammang et al., 2011a, 2011b) to<br />

image the 3D wake structure behind the flexible foil shown in Figure 4B. The trailing edge of<br />

the foil is located at the −60 mm position on the x-axis, <strong>and</strong> the fluid structures shown are all<br />

in the wake of the foil. This foil was self-propelling <strong>and</strong> was actuated using the same parameters<br />

shown for Figure 4B. (A) The flexible foil achieved a ribbon-like shape <strong>and</strong> columns of vorticity<br />

extend vertically on either side of the foil. Vorticity is isosurfaced at a value of 3.1 (blue surface),<br />

<strong>and</strong> a horizontal slice through the wake is shown to correspond to that shown using 2D piv in<br />

Figure 4B (green plane with velocity vectors). (B) The vortical columns on each side of the<br />

ribbon-like foil motion connect to each other across the top <strong>and</strong> bottom of the same side (yellow<br />

arrows) <strong>and</strong> opposite sides (red arrows).<br />

The work per cycle stays nearly constant,<br />

as does the cost per meter<br />

(Table 1), but the cost of transport<br />

decreases substantially as the increased<br />

mass of the longer foils does not result<br />

in increased energy requirements for<br />

propulsion. Foils such as these that<br />

are composed of a very flexible material<br />

self-propel at shorter lengths with a<br />

Strouhal number that is quite large<br />

relative to most self-propelling bodies.<br />

At longer lengths, the Strouhal number<br />

approaches that of many swimming<br />

fishes or rigid foils (Table 1).<br />

Figure 6 shows changes in shape<br />

of self-propelling foils made of the<br />

same material (flexural stiffness, 3.1 ×<br />

10 −6 Nm 2 ) that occur due to change<br />

FIGURE 6<br />

Graphs to show the effect of length on the<br />

shape (amplitude envelope) of a flexible foil<br />

(flexural stiffness = 3.1 × 10 −6 Nm 2 )swimming<br />

at its SPS. This foil was 6.85 cm in<br />

chord, <strong>and</strong> of varying length, given in the<br />

color-coded legend. The leading edge was actuated<br />

at 2 Hz <strong>and</strong> with amplitude of ±1 cm.<br />

Each plot shows the shape of the foil during<br />

self-propulsion as indicated by the peak-topeak<br />

amplitude of the sideways flapping motion.<br />

(A, B) Foil shapes as the absolute distance<br />

along the foil <strong>and</strong> as percentage of the total foil<br />

length, respectively.<br />

lengths made of the same material <strong>and</strong><br />

actuated in heave only at the leading<br />

edge. Altering the length of this flexible<br />

swimming foil from 20 to 35 cm produces<br />

only minor changes in the SPS<br />

(from10toabout9cms −1 )<strong>and</strong><br />

hence in Reynolds number but dramatically<br />

lowers the Strouhal number<br />

(from0.83to0.34;Table1)dueto<br />

changes in foil trailing edge amplitude.<br />

July/August 2011 Volume 45 Number 4 47


in length. Longer foils show peak excursions<br />

at the same locations as shorter<br />

foils (Figure 6A) but amplitudes that<br />

are lower. Each of the foils of this material<br />

generates a ribbon-like shape<br />

when self-propelling with a consistent<br />

wave-like pattern down its length (see<br />

Figure 4B). When the amplitude of<br />

each foil as a percentage of foil length<br />

is plotted (Figure 6B), changes in<br />

amplitude are more evident with<br />

side-to-side excursion amplitudes decreasing<br />

as length increases while the<br />

wave-like pattern is retained. The<br />

shortest foils have high amplitudes<br />

near the trailing edge <strong>and</strong> an amplitude<br />

envelope that grows along the foil<br />

while the longer foils display a tapering<br />

amplitude envelope (compare blue <strong>and</strong><br />

black curves in Figure 6B). These data<br />

show that length alone can have significant<br />

effects on locomotor efficiency<br />

<strong>and</strong> foil kinematics <strong>and</strong> that, all other<br />

things held constant, foil length increases<br />

on the order of 100% can provide<br />

reduced costs of transport.<br />

The Effect of Trailing<br />

Edge Shape on<br />

Swimming Performance<br />

The shape of the trailing tail edge<br />

of swimming fishes has been the subject<br />

of considerable discussion in the<br />

literature, with various authors considering<br />

the advantages or disadvantages<br />

of fish tails with symmetrical,<br />

asymmetrical, or forked shapes (Affleck,<br />

1950; Aleev, 1969; Lauder, 1989;<br />

Plaut, 2000; Thomson, 1971, 1976;<br />

Wilga & Lauder, 2004a). One advantage<br />

of a robotic flapping foil approach<br />

is that the trailing edge of a flexible<br />

flapping foil can be altered <strong>and</strong> a variety<br />

of configurations constructed that<br />

allow the effect of trailing edge shape<br />

alone on locomotor performance to<br />

FIGURE 7<br />

Propulsion by flexible foils of different shapes (material flexural stiffness = 3.1 × 10 −4 Nm 2 ) swimming<br />

at their SPS. Each foil was actuated at ±1 cm heave at 2 Hz. Error bars are ±2 SE. Strouhal<br />

numbers for these experiments range from 0.2 to 0.3. Foils were constructed to be of differing<br />

shapes <strong>and</strong> areas as follows: foil 1= square trailing edge, area = 131.2 cm 2 , dimensions = 6.85 cm ×<br />

19.15 cm; foil 2 = angled trailing edge, same length as foil 1, area = 107.71 cm 2 ; foil 3 = angled<br />

trailing edge, same area as foil 1; foil 4 = angled trailing edge: same length <strong>and</strong> area as foil 1; foil 5 =<br />

forked trailing edge: same area as foil 1.<br />

be investigated. We designed five different<br />

flexible foils (material flexural<br />

stiffness = 3.1 × 10 −4 Nm 2 ) that control<br />

for total length <strong>and</strong> foil area <strong>and</strong><br />

allow different tail shapes to be compared<br />

for the effect of these changes<br />

on swimming speed (Figure 7). Foil 1<br />

corresponds to a highly abstracted<br />

“trout-like” body shape with a mostly<br />

vertical trailing tail edge, while foils 3<br />

<strong>and</strong> 4 present a shark-like tail trailing<br />

edge shape. Many fish have forked<br />

tails, <strong>and</strong> this shape is represented by<br />

foil 5.<br />

The fastest swimming foil (Figure<br />

7, foil 4) is the one with the<br />

most area near the axis of actuation,<br />

even though it has the same area as<br />

several of the other foils. This result<br />

corresponds with our previous results<br />

showing that foils with higher aspect<br />

ratios <strong>and</strong> hence more material near<br />

the actuator swim significantly faster<br />

(Lauderetal.,2011).Interestingly,<br />

the foil with the angled trailing edge<br />

<strong>and</strong> the shark tail shape (Figure 7,<br />

foil 3) swims significantly faster (P <<br />

0.003) than a foil with the same<br />

area but a straight trailing edge (Figure<br />

7, foil 1). The foil shape with<br />

the significantly lowest swimming<br />

speed was the notched shape (Figure<br />

7, foil 5) even though it possesses<br />

the same area as foils 1 <strong>and</strong> 3. The reduced<br />

performance of the notched<br />

shape may be due to bending of the<br />

upper <strong>and</strong> lower “lobes” of the tail<br />

during the flapping motion, <strong>and</strong> kinematic<br />

data obtained for these foils<br />

does show that each lobe of this shape<br />

twists during the flapping cycle. Twisting<br />

of the tail could be reduced by introducing<br />

stiffening elements, <strong>and</strong> this<br />

may be one reason that fishes with<br />

high-performance tail shapes such as<br />

tuna possess substantial stiffening of<br />

both the upper <strong>and</strong> lower tail lobes<br />

(Fierstine & Walters, 1968; Westneat<br />

& Wainwright, 2001).<br />

Although the angled foil shape<br />

swims significantly faster than a foil<br />

of the same area with a straight trailing<br />

edge (Figure 7: compare foils 1 <strong>and</strong> 3),<br />

more power is required for this foil to<br />

swim at this (self-propelled) speed<br />

(Figure 8A). The foils were made of<br />

48 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 8<br />

Graphs of power consumed (A) <strong>and</strong> the cost of<br />

transport (B) of foils 1 <strong>and</strong> 3 (see Figure 7) selfpropelling.<br />

Both foils have the same surface<br />

area (131.2 cm 2 ) <strong>and</strong> were actuated with<br />

±1 cm heave at 2 Hz. Power values are significantly<br />

different at P =0.003.Costoftransport<br />

values are not significantly different at P =0.07.<br />

thesamematerial<strong>and</strong>havethesame<br />

area, so the cost of transport can be<br />

calculated in mJ/m. Figure 8B shows<br />

that there is no significant difference<br />

(P = 0.07) between the foils in cost of<br />

transport, although there is a trend in<br />

the data toward the angled foil edge<br />

costing less per meter resulting in the<br />

marginally non-significant difference.<br />

Further experiments on both foils<br />

will be needed to extend this result<br />

<strong>and</strong> to determine if in fact there is a<br />

small difference in cost of transport between<br />

these two foil types.<br />

fishes that minimizes the complexity<br />

of the foil <strong>and</strong> maintains constant material<br />

properties along the foil length,<br />

fish bodies are clearly different in<br />

showing changing material properties<br />

from head to tail (McHenry et al.,<br />

1995). To better underst<strong>and</strong> the locomotor<br />

properties of the passive fish<br />

body alone, we used freshly dead rainbow<br />

trout (Oncorhynchus mykiss) <strong>and</strong><br />

attached the body to the robotic flapping<br />

foil apparatus (Figures 1D, 1E,<br />

<strong>and</strong> 1F). We took care to ensure<br />

that rigor mortis had not set in during<br />

the experiments <strong>and</strong> to ensure that<br />

the body had thus not stiffened during<br />

the time the flapping trials were<br />

conducted. By actuating the passive<br />

trout body in heave <strong>and</strong> pitch in various<br />

combinations just behind the<br />

head,wewereabletoconstructa<br />

swimming performance surface for<br />

the passive fish body (Figure 9).<br />

Body waveforms that are remarkably<br />

like those occurring in live trout<br />

were generated by the passive fish<br />

bodies. These data show that heave<br />

amplitude overall has the greater effect<br />

on swimming performance than<br />

FIGURE 9<br />

changes in pitch alone. At any given<br />

heave value increases in pitch further<br />

increase swimming speed, but at<br />

higher heave amplitudes near ±2 cm,<br />

changes in pitch only produce modest<br />

increases in SPS (Figure 9). These<br />

data provide an interesting comparison<br />

to our previous results showing<br />

the effects of heave <strong>and</strong> pitch actuation<br />

on flexible foil propulsion (Lauder<br />

et al., 2011). In that study adding<br />

pitch motions to baseline heave actuation<br />

for foils of varying flexural stiffness<br />

did not produce significant<br />

increases in foil SPS but did allow<br />

stiffer foils to maintain a relatively<br />

high swimming speed that would<br />

have declined with heave actuation<br />

only.<br />

As driving frequency increases, the<br />

tail beat amplitude of the passive flapping<br />

trout body remains relatively<br />

constant from 0.1 to 2.0 Hz before<br />

increasing steadily to a peak at<br />

3.5 Hz (Figure 10). These tail beat<br />

amplitude values are very similar to<br />

those observed in live trout swimming<br />

under a variety of locomotor conditions<br />

(Liao et al., 2003a; Webb,<br />

Performance surface of a freshly dead trout (Oncorhynchus mykiss) attached to a robotic controller<br />

(see Figures 1D, 1E, 1F) driven at 2 Hz at a variety of heave <strong>and</strong> pitch amplitudes. This trout was<br />

25.3 cm in total length. The graph shows how SPS of the passive trout body varies with different<br />

pitch <strong>and</strong> heave actuation parameters.<br />

Undulatory Locomotion<br />

of a Passive Fish Body<br />

Although flapping flexible foils<br />

provide a reasonable <strong>and</strong> simple<br />

model for undulatory propulsion in<br />

July/August 2011 Volume 45 Number 4 49


FIGURE 10<br />

Graph of tail-tip amplitude versus heave actuation frequency for a freshly dead trout (Oncorhynchus<br />

mykiss) attached to a robotic controller driving the passive body at a variety of frequencies. This trout<br />

was 25.3 cm in total length <strong>and</strong> was actuated with a constant heave (±2 cm) <strong>and</strong> pitch (±5°). Error<br />

bars are ±1 SE of the mean.<br />

1971; Webb et al., 1984) <strong>and</strong> increases<br />

in tail beat amplitude of the<br />

magnitude shown here for the passive<br />

trout body are similar to those observed<br />

previously as fish increase<br />

swimming speed <strong>and</strong> alter both frequency<br />

(primarily) <strong>and</strong> amplitude<br />

(to a lesser extent) of the tail beat.<br />

Although during routine undulatory<br />

swimming fish bodies are not<br />

passive <strong>and</strong> are certainly stiffened by<br />

body musculature as fish swim, for<br />

all but the fastest swimming speeds<br />

locomotion is powered by red muscle<br />

fibers, which can make up a rather<br />

small percentage of body mass<br />

(around 1.5% in largemouth bass;<br />

see Johnson et al., 1994). The bulk<br />

of the locomotor musculature is composed<br />

of white fibers that are not activated<br />

until the fastest swimming<br />

speeds are needed (Jayne & Lauder,<br />

1994, 1995a). Thus, the red muscle<br />

fibers can be considered as acting to<br />

bend a mostly passive fish body composed<br />

of white muscle fibers <strong>and</strong> associated<br />

skeletal tissues that may not be<br />

much different in flexural stiffness<br />

from the passive bodies studied here.<br />

Also, data on the propulsion of passive<br />

fish bodies are relevant to fishes<br />

swimming in turbulent flows. Trout<br />

swimming in a vortex street have been<br />

shown to greatly alter body kinematics<br />

<strong>and</strong> to utilize vortical energy shed from<br />

objects in the flow (Liao, 2004; Liao<br />

et al., 2003b). The amplitude of the<br />

center of mass oscillation by trout<br />

swimming in the Karman gait is up to<br />

±2 cm for a 10-cm-long trout, which<br />

corresponds on a length-specific basis<br />

to the middle region of the performance<br />

surface in Figure 9 (Liao et al., 2003a).<br />

By greatly reducing muscle activity<br />

when trout enter a vortex street, the<br />

body becomes largely passive <strong>and</strong> deforms<br />

in response to the oncoming<br />

flows. Trouts are able to maintain<br />

their position in such flows entirely passively<br />

<strong>and</strong> allow their bodies to extract<br />

energy from the oncoming flow <strong>and</strong><br />

generate thrust. This phenomenon<br />

was further demonstrated in a study<br />

using passive foils <strong>and</strong> freshly dead<br />

fish bodies to show that the passive<br />

trout body alone in a vortex wake can<br />

generate sufficient thrust to maintain<br />

position in flow (Beal et al., 2006).<br />

Fin-Fin Hydrodynamic<br />

Interactions<br />

One of the most intriguing aspects<br />

of fish functional design is the arrangement<br />

of two fins in series that could<br />

allow enhanced locomotor efficiency<br />

through hydrodynamic interactions<br />

between the fins. For example, the<br />

dorsal fin<strong>and</strong>theanalfininmostfishes<br />

are located upstream of the caudal fin,<br />

<strong>and</strong> the wakes shed from these fins<br />

could interact with the tail fin during<br />

locomotion (Drucker & Lauder, 2001,<br />

2005; Tytell, 2006). Undulatory locomotion<br />

using the body also causes the<br />

attached median fins to oscillate from<br />

side to side, <strong>and</strong> in the fish species studied<br />

so far these fins have been shown also<br />

to be actively oscillated by intrinsic fin<br />

musculature (Jayne et al., 1996) <strong>and</strong><br />

thus can generate thrust on their own.<br />

The possibility of fin-fin hydrodynamic<br />

interactions during locomotion<br />

has been explored in a number of<br />

experimental papers on living fishes<br />

(Drucker & Lauder, 2001, 2005;<br />

St<strong>and</strong>en, 2008; St<strong>and</strong>en & Lauder,<br />

2007; Tytell, 2006; Webb & Keyes,<br />

1981) as well as using computational<br />

approaches (Akhtar et al., 2007; Weihs<br />

et al., 2006). Akhtar et al. (2007) used<br />

data on the kinematics of the bluegill<br />

dorsal fin <strong>and</strong> tail from Drucker <strong>and</strong><br />

Lauder (2001) <strong>and</strong> performed a twodimensional<br />

computational analysis of<br />

the effect of having the fins in series<br />

<strong>and</strong> found that vortex shedding from<br />

the dorsal fin can increase thrust of the<br />

tail <strong>and</strong> that the amount of this thrust<br />

increase depends on the phasing of<br />

dorsal fin <strong>and</strong>tailmotion.<br />

In order to experimentally evaluate<br />

hydrodynamic fin-fin interactions<br />

using an apparatus in which phasing<br />

<strong>and</strong> the distance between fins can be<br />

experimentally manipulated, we used<br />

our robotic flapping foil apparatus<br />

in the dual-foil configuration with<br />

50 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


upstream <strong>and</strong> downstream foils. Foils<br />

were rigid aluminum plates in a<br />

NACA 0012 airfoil shape, <strong>and</strong> each<br />

foil could be moved in heave <strong>and</strong><br />

pitch, <strong>and</strong> both foils were driven from<br />

a common carriage mounted on air<br />

bearings as described above. Foils were<br />

moved according to the parameters<br />

measured for bluegill sunfish dorsal<br />

<strong>and</strong> caudal fins (Drucker & Lauder,<br />

2001) <strong>and</strong> used by Akhtar et al.<br />

(2007) for their computational study<br />

of this problem: the upstream foil was<br />

moved with heave amplitude of<br />

2.5 cm, the downstream foil at<br />

3.5 cm heave amplitude. Pitch amplitude<br />

for both foils was ±20°, <strong>and</strong> the<br />

frequency for both foils was 1.7 Hz,<br />

corresponding to the frequency of fin<br />

flapping in the bluegill sunfish model<br />

case (Drucker & Lauder, 2001).<br />

Increases in SPS as a result of<br />

changing foil phase <strong>and</strong> distance reflect<br />

thrust enhancement beyond the<br />

thrust achieved by the two foils operating<br />

separately (assessed by offsetting<br />

the foils from each other so that the<br />

downstream foil was no longer in<br />

the wake of the upstream foil). A general<br />

description of the dual-foil configuration<br />

<strong>and</strong> images of flows<br />

around <strong>and</strong> between the foils are presented<br />

in Lauder et al. (2007). Data<br />

for SPS were plotted over a range of<br />

phase differences between the upstream<br />

<strong>and</strong> downstream foils <strong>and</strong><br />

polynomial fits to these data points<br />

were used to determine the change<br />

in SPS with phase (e.g., Figure 11).<br />

Here we present the results of experiments<br />

measuring the SPS of dual<br />

foil flapping. Figure 11a shows the effect<br />

of changing the phase of sinusoidal<br />

motion of the downstream<br />

foil relative to the upstream foil on<br />

SPS for three different spacings between<br />

the foils. In each case, there is<br />

a clear peak in swimming speed, <strong>and</strong><br />

FIGURE 11<br />

Robotic model of fish fin hydrodynamic interactions. Dual NACA 0012 flapping foils in series are<br />

driven by a robotic flapping foil apparatus. Details of this device <strong>and</strong> of the experimental setup are<br />

given in Lauder et al. (2007). (A) SPS plotted against the phase difference between the upstream<br />

<strong>and</strong> downstream foils. Foils were arranged with 0 side-to-side offset (in cm) <strong>and</strong> are thus moving<br />

in line with each other with the downstream foil at varying chord length separations from the<br />

upstream foil: 0.5, 1, <strong>and</strong> 2 chord lengths separation; foil chord length was 6.85 cm. Further<br />

experimental details are provided in the text. (B) Effect of changing the foil offsets (in cm) so<br />

that the downstream foil is not moving in the direct wake of the upstream foil. The one chord length<br />

spacing with zero offset curve (dark blue) is the same as in panel (A) <strong>and</strong> is shown for reference.<br />

The other three plots show the effect of a 3.5 cm offset of the midline motion of the downstream<br />

foil relative to the upstream foil, removing it from the upstream foil wake.<br />

as the distance between the foils is increased<br />

the peak swimming speed<br />

shifts toward a larger phase lag of<br />

the downstream foil. Interestingly,<br />

the maximal swimming speed of the<br />

two foils together does not change significantly<br />

as the distance between foils<br />

changes, <strong>and</strong> alterations in phasing<br />

between the foils can thus be used<br />

to compensate for changes in spacing.<br />

For each interfoil spacing, however,<br />

there are phase relationships that significantly<br />

reduce swimming performance,<br />

indicating that thrust of the<br />

entire two-foil system is sensitive to<br />

phase relationships between two foils<br />

flapping. These data correspond very<br />

well to the computational results of<br />

Akhtar et al. (2007), which showed<br />

peak thrust enhancement at a phase<br />

July/August 2011 Volume 45 Number 4 51


of about 40°, very similar to our data<br />

(Figure11A,bluecurve)wherethe<br />

plateau around the SPS peak includes<br />

the 40° value.<br />

Figure 11B illustrates the changes<br />

in propulsion that result from moving<br />

the downstream foil to the side by an<br />

offset of 3.5 cm to move it out of the<br />

wake of the upstream foil. This effectively<br />

creates a t<strong>and</strong>em foil configuration<br />

in which fluid dynamic<br />

interactions between the foils are<br />

minimized, although not completely<br />

eliminated. Comparison of the three<br />

offset curves to the zero offset curve<br />

shows that offsetting the foils reduces<br />

both the SPS <strong>and</strong> the effect of foil<br />

phasing. Offset plots show reduced<br />

effects of foil phasing (with lower<br />

maxima <strong>and</strong> higher minima) <strong>and</strong> less<br />

distinct overall peaks in SPS, showing<br />

that the downstream foil was not able<br />

to improve swimming speed of the<br />

two foils together when removed<br />

from the upstream foil wake.<br />

Computational work <strong>and</strong> preliminary<br />

flow visualization (Lauder et al.,<br />

2007) data indicate that the behavior<br />

of t<strong>and</strong>em foils can be explained in<br />

terms of how the downstream foil interacts<br />

with fluid structures generated by<br />

the upstream foil. The shifting of the<br />

peak SPS with increased spacing of<br />

the t<strong>and</strong>em foils can be explained by<br />

the fact that, for larger spacings, the convection<br />

of those structures to the downstream<br />

foil takes longer. That time, tc,is<br />

simply the spacing between the foils, s,<br />

divided by the convection velocity, Uc.<br />

The increase in phase lag, Δϕ, needed<br />

so that the downstream foil meets the<br />

fluid structures at the right time is simply<br />

ðS 2 S 1 Þ<br />

Δϕ ¼ 2πf<br />

U c<br />

where f is the frequency of flapping<br />

<strong>and</strong> s 2 <strong>and</strong> s 1 are two different spacings.<br />

If we assume that Uc is close to the<br />

SPS, this equation estimates the Δϕ<br />

between the peak SPS well for the<br />

three spacings we used. The equation<br />

predicts Δϕ = 0.648 radians, or 37°,<br />

for a spacing difference of 0.5 chord<br />

lengths, <strong>and</strong> 74° for a difference of<br />

one chord length. The Δϕ from our<br />

data are approximately 30°<strong>and</strong> 75° for<br />

the corresponding spacing differences.<br />

The nearness of the foils <strong>and</strong>/or higher<br />

convection velocities at the smaller<br />

spacings may explain the lower than<br />

predicted Δϕ for the peaks in SPS in<br />

these cases. Overall, the agreement is<br />

very good considering that the peaks<br />

are somewhat broad <strong>and</strong> the equation<br />

for Δϕ is a simplification of the fluid<br />

dynamics.<br />

The Future of<br />

Undulatory Biorobotics<br />

In this paper, we use a robotic tool<br />

for investigating a variety of phenomena<br />

relating to undulatory propulsion<br />

in fishes <strong>and</strong> present experimental<br />

data that would be difficult if not impossible<br />

to obtain from studying live<br />

animals. The promise of robotic models<br />

for studying the biomechanics of<br />

locomotion in fishes has just begun<br />

to be realized (Curet et al., 2011;<br />

Long et al., 2006, 2010; Tangorra<br />

et al., 2010, 2011), <strong>and</strong> fundamental<br />

questions relating to the mechanics of<br />

undulatory propulsion remain to be<br />

addressed. In particular, key unresolved<br />

issues are the extent to which<br />

changes in body stiffness during propulsion<br />

affect locomotor performance<br />

(see Long & Nipper, 1996) <strong>and</strong> how<br />

active modulation of stiffness during<br />

an undulatory cycle <strong>and</strong> across<br />

changes in swimming speed are<br />

achieved <strong>and</strong> affect propulsive speed<br />

<strong>and</strong> efficiency.<br />

To address these questions, a new<br />

generation of robotic undulatory devices<br />

will be needed that allow for<br />

controlled modulation of body stiffness<br />

<strong>and</strong> the phasing of stiffness<br />

changes with undulatory cycles of<br />

compression <strong>and</strong> tension on the<br />

bending fish or foil body. An additional<br />

arena that is key to making<br />

progress in underst<strong>and</strong>ing undulatory<br />

mechanics is the ability to perturb the<br />

locomotor system to assess how stiffness<br />

of the body relates to the ability<br />

to recover from perturbations. There<br />

have been very few studies of perturbations<br />

of undulatory locomotor systems<br />

(see Webb, 2004), <strong>and</strong> yet fishes<br />

oftenswiminchallenginghydrodynamic<br />

environments in which they<br />

are forced to recover from impulsive<br />

challenges to their undulatory<br />

pattern.<br />

Acknowledgments<br />

This work was supported the Office<br />

of Naval Research grant N00014-<br />

09-1-0352 on fin neuromechanics<br />

monitored by Dr. Thomas McKenna<br />

<strong>and</strong> by the National Science Foundation<br />

grant EFRI-0938043. We<br />

thank the members of the Lauder<br />

<strong>and</strong> Tangorra labs for many helpful<br />

discussions on fish fins <strong>and</strong> flexible<br />

flapping foil propulsion <strong>and</strong> Nate<br />

Jackson for his assistance with the<br />

dual-flapping foil experiments. Many<br />

thanks to Brooke Flammang, Tyson<br />

Str<strong>and</strong>, <strong>and</strong> Dan Troolin for assistance<br />

with the V3V volumetric flow<br />

imaging experiments on the undulating<br />

plastic foil.<br />

Lead Author:<br />

George V. Lauder<br />

The Museum of<br />

Comparative Zoology<br />

52 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


26 Oxford Street, Harvard University<br />

Cambridge, MA 02138<br />

Email: glauder@oeb.harvard.edu<br />

References<br />

Affleck, R.J. 1950. Some points in the<br />

function, development, <strong>and</strong> evolution of the<br />

tail in fishes. Proc Zool Soc Lond. 120:349-68.<br />

doi: 10.1111/j.1096-3642.1950.tb00954.x.<br />

Akhtar, I., Mittal, R., Lauder, G.V., &<br />

Drucker, E. 2007. Hydrodynamics of a<br />

biologically inspired t<strong>and</strong>em flapping foil<br />

configuration. Theor Comp Fluid Dyn.<br />

21:155-70. doi: 10.1007/s00162-007-0045-2.<br />

Aleev, Y.G. 1969. Function <strong>and</strong> Gross<br />

Morphology in Fish. Jerusalem: Keter Press.<br />

Beal, D.N., Hover, F.S., Triantafyllou,<br />

M.S., Liao, J., & Lauder, G.V. 2006.<br />

Passive propulsion in vortex wakes.<br />

J Fluid Mech. 549:385-402. doi: 10.1017/<br />

S0022112005007925.<br />

Bone, Q., Kiceniuk, J., & Jones, D.R. 1978.<br />

On the role of the different fibre types in fish<br />

myotomes at intermediate swimming speeds.<br />

Fish Bull. 76:691-9.<br />

Curet, O.M., Patankar, N.A., Lauder, G.V.,<br />

& MacIver, M.A. 2011. Aquatic manoeuvering<br />

with counter-propagating waves: A novel<br />

locomotive strategy. J R Soc Interface.<br />

8:1041-50. doi: 10.1098/rsif.2010.0493.<br />

Donley, J., & Dickson, K.A. 2000. Swimming<br />

kinematics of juvenile Kawakawa tuna<br />

(Euthynnus affinis) <strong>and</strong> chub mackerel (Scomber<br />

japonicus). J Exp Biol. 203:3103-16.<br />

Drucker, E.G., & Lauder, G.V. 1999.<br />

Locomotor forces on a swimming fish: Threedimensional<br />

vortex wake dynamics quantified<br />

using digital particle image velocimetry.<br />

J Exp Biol. 202:2393-412.<br />

Drucker, E.G., & Lauder, G.V. 2001.<br />

Locomotor function of the dorsal fin in teleost<br />

fishes: Experimental analysis of wake forces<br />

in sunfish. J Exp Biol. 204:2943-58.<br />

Drucker, E.G., & Lauder, G.V. 2005.<br />

Locomotor function of the dorsal fin in<br />

rainbow trout: Kinematic patterns <strong>and</strong> hydrodynamic<br />

forces. J Exp Biol. 208:4479-94.<br />

doi: 10.1242/jeb.01922.<br />

Fetcho, J., & Svoboda, K.R. 1993. Fictive<br />

swimming elicited by electrical stimulation<br />

of the midbrain in goldfish. J Neurophysiol.<br />

70:764-80.<br />

Fetcho, J.R. 1986. The organization of the<br />

motoneurons innervating the axial musculature<br />

of vertebrates: I. Goldfish (Carassius<br />

auratus) <strong>and</strong> mudpuppies (Necturus maculosus).<br />

J Comp Neurol. 249:521-50. doi: 10.1002/<br />

cne.902490408.<br />

Fierstine, H.L., & Walters, V. 1968. Studies<br />

in locomotion <strong>and</strong> anatomy of scombroid<br />

fishes. Mem Southern Calif Acad Sci. 6:1-31.<br />

Flammang, B.E., & Lauder, G.V. 2008.<br />

Speed-dependent intrinsic caudal fin muscle<br />

recruitment during steady swimming in<br />

bluegill sunfish, Lepomis macrochirus. J Exp<br />

Biol. 211:587-98. doi: 10.1242/jeb.012096.<br />

Flammang, B.E., & Lauder, G.V. 2009.<br />

Caudal fin shape modulation <strong>and</strong> control<br />

during acceleration, braking <strong>and</strong> backing<br />

maneuvers in bluegill sunfish, Lepomis<br />

macrochirus. J Exp Biol. 212:277-86.<br />

doi: 10.1242/jeb.021360.<br />

Flammang, B.E., Lauder, G.V., Troolin,<br />

D.R., & Str<strong>and</strong>, T. 2011a. Volumetric<br />

imaging of shark tail hydrodynamics reveals<br />

a three-dimensional dual-ring vortex wake<br />

structure. P Roy Soc B. 278. doi: 10.1098/<br />

rspb.2011.0489.<br />

Flammang, B.E., Lauder, G.V., Troolin,<br />

D.R., & Str<strong>and</strong>, T.E. 2011b. Volumetric<br />

imaging of fish locomotion. Biol Lett.<br />

doi: 10.1098/rsbl.2011.0282.<br />

Gillis, G.B. 1996. Undulatory locomotion<br />

in elongate aquatic vertebrates: Anguilliform<br />

swimming since Sir. James Gray. Am Zool.<br />

36:656-65.<br />

Gillis, G.B. 1998. Environmental effects on<br />

undulatory locomotion in the American eel<br />

Anguilla rostrata: Kinematics in water <strong>and</strong><br />

on l<strong>and</strong>. J Exp Biol. 201:949-61.<br />

Harris, J.E. 1936. The role of the fins in the<br />

equilibrium of the swimming fish: I. Wind<br />

tunnel tests on a model of Mustelus canis<br />

(Mitchell). J Exp Biol. 13:476-93.<br />

Harris, J.E. 1938. The role of the fins in the<br />

equilibrium of the swimming fish: II. The role<br />

of the pelvic fins. J Exp Biol. 16:32-47.<br />

Hove, J.R., O’Bryan, L.M., Gordon, M.S.,<br />

Webb, P.W., & Weihs, D. 2001. Boxfishes<br />

(Teleostei: Ostraciidae) as a model system for<br />

fishes swimming with many fins: Kinematics.<br />

J Exp Biol. 204:1459-71.<br />

Jayne, B.C., & Lauder, G.V. 1993. Red <strong>and</strong><br />

white muscle activity <strong>and</strong> kinematics of the<br />

escape response of the bluegill sunfish during<br />

swimming. J Comp Physiol A. 173:495-508.<br />

doi: 10.1007/BF00193522.<br />

Jayne, B.C., & Lauder, G.V. 1994. How<br />

swimming fish use slow <strong>and</strong> fast muscle fibers:<br />

Implications for models of vertebrate muscle<br />

recruitment. J Comp Physiol A. 175:123-31.<br />

doi: 10.1007/BF00217443.<br />

Jayne, B.C., & Lauder, G.V. 1995a. Are<br />

muscle fibers within fish myotomes activated<br />

synchronously Patterns of recruitment within<br />

deep myomeric musculature during swimming<br />

in largemouth bass. J Exp Biol.<br />

198:805-15.<br />

Jayne, B.C., & Lauder, G.V. 1995b. Red<br />

muscle motor patterns during steady swimming<br />

in largemouth bass: Effects of speed<br />

<strong>and</strong> correlations with axial kinematics.<br />

J Exp Biol. 198:1575-87.<br />

Jayne, B.C., & Lauder, G.V. 1995c.<br />

Speed effects on midline kinematics during<br />

steady undulatory swimming of largemouth<br />

bass, Micropterus salmoides. JExpBiol.<br />

198:585-602.<br />

Jayne, B.C., Lozada, A., & Lauder, G.V.<br />

1996. Function of the dorsal fin in bluegill<br />

sunfish: Motor patterns during four locomotor<br />

behaviors. J Morphol. 228:307-26.<br />

doi:10.1002/(SICI)1097-4687(199606)<br />

228:33.0.CO;2-Z.<br />

Johnson, T.P., Syme, D.A., Jayne, B.C.,<br />

Lauder, G.V., & Bennett, A.F. 1994. Modeling<br />

red muscle power output during steady <strong>and</strong> unsteady<br />

swimming in largemouth bass (Micropterus<br />

salmoides). Am J Physiol. 267:R481-8.<br />

July/August 2011 Volume 45 Number 4 53


Lauder, G., Madden, P.G.A., Tangorra, J.,<br />

Anderson, E., & Baker, T.V. 2011.<br />

Bioinspiration from fish for smart material<br />

design <strong>and</strong> function. Smart Mater Struct.<br />

in press.<br />

Lauder, G.V. 1989. Caudal fin locomotion<br />

in ray-finned fishes: Historical <strong>and</strong> functional<br />

analyses. Am Zool. 29:85-102.<br />

Lauder, G.V. 2000. Function of the caudal<br />

fin during locomotion in fishes: Kinematics,<br />

flow visualization, <strong>and</strong> evolutionary patterns.<br />

Am Zool. 40:101-22. doi: 10.1668/0003-<br />

1569(2000)040[0101:FOTCFD]2.0.CO;2.<br />

Lauder, G.V., Anderson, E.J., Tangorra, J.,<br />

& Madden, P.G.A. 2007. Fish biorobotics:<br />

Kinematics <strong>and</strong> hydrodynamics of selfpropulsion.<br />

J Exp Biol. 210:2767-80.<br />

doi: 10.1242/jeb.000265.<br />

Lauder, G.V., & Tytell, E.D. 2006. Hydrodynamics<br />

of undulatory propulsion. In:<br />

Fish Biomechanics. Volume 23 in Fish<br />

Physiology, eds. Shadwick, R.E., & Lauder,<br />

G.V., 425-68. San Diego: Academic Press.<br />

Liao, J. 2002. Swimming in needlefish<br />

(Belonidae): Anguilliform locomotion with<br />

fins. J Exp Biol. 205:2875-84.<br />

Liao, J. 2004. Neuromuscular control of trout<br />

swimming in a vortex street: Implications for<br />

energy economy during the Karman gait.<br />

J Exp Biol. 207:3495-506. doi: 10.1242/<br />

jeb.01125.<br />

Liao, J., Beal, D.N., Lauder, G.V., &<br />

Triantafyllou, M.S. 2003a. The Kármán gait:<br />

Novel body kinematics of rainbow trout<br />

swimming in a vortex street. J Exp Biol.<br />

206:1059-73. doi: 10.1242/jeb.00209.<br />

Liao, J., & Lauder, G.V. 2000. Function<br />

of the heterocercal tail in white sturgeon: Flow<br />

visualization during steady swimming <strong>and</strong><br />

vertical maneuvering. J Exp Biol. 203:3585-94.<br />

Liao, J.C., Beal, D.N., Lauder, G.V., &<br />

Triantafyllou, M.S. 2003b. Fish exploiting<br />

vortices decrease muscle activity. Science.<br />

302:1566-9. doi: 10.1126/science.1088295.<br />

Long, J.H., Jr., Koob, T.J., Irving, K.,<br />

Combie, K., Engel, V., Livingston, N., …<br />

Schumacher, J. 2006. Biomimetic evolutionary<br />

analysis: Testing the adaptive value of<br />

vertebrate tail stiffness in autonomous swimming<br />

robots. J Exp Biol. 209:4732-46.<br />

doi: 10.1242/jeb.02559.<br />

Long, J.H., McHenry, M.J., & Boetticher,<br />

N.C. 1994. Undulatory swimming: How<br />

traveling waves are produced <strong>and</strong> modulated<br />

in sunfish (Lepomis gibbosus). J Exp Biol.<br />

192:129-45.<br />

Long, J.H., & Nipper, K.S. 1996. The<br />

importance of body stiffness in undulatory<br />

propulsion. Am Zool. 36:678-694.<br />

Long, J.H., Porter, M.E., Root, R.G., &<br />

Liew, C.W. 2010. Go reconfigure: How<br />

fish change shape as they swim <strong>and</strong> evolve.<br />

Integr Comp Biol. 50:1120-39. doi: 10.1093/<br />

icb/icq066.<br />

McHenry, M.J., Pell, C.A., & Long, J.A.<br />

1995. Mechanical control of swimming speed:<br />

Stiffness <strong>and</strong> axial wave form in undulating<br />

fish models. J Exp Biol. 198:2293-305.<br />

Nauen, J.C., & Lauder, G.V. 2002a. Hydrodynamics<br />

of caudal fin locomotion by chub<br />

mackerel, Scomber japonicus (Scombridae).<br />

J Exp Biol. 205:1709-24.<br />

Nauen, J.C., & Lauder, G.V. 2002b.<br />

Quantification of the wake of rainbow trout<br />

(Oncorhynchus mykiss) using three-dimensional<br />

stereoscopic digital particle image velocimetry.<br />

J Exp Biol. 205:3271-9.<br />

Plaut, I. 2000. Effects of fin size on swimming<br />

performance, swimming behavior <strong>and</strong><br />

routine activity of zebrafish Danio rerio.<br />

J Exp Biol. 203:813-20.<br />

Rome, L.C., Swank, D., & Corda, D. 1993.<br />

How fish power swimming. Science. 261:<br />

340-3. doi: 10.1126/science.8332898.<br />

Shadwick, R., & Gemballa, S. 2006.<br />

Structure, kinematics, <strong>and</strong> muscle dynamics in<br />

undulatory swimming. In: Fish Biomechanics.<br />

Volume 23 in Fish Physiology, eds. Shadwick,<br />

R.E., & Lauder, G.V., 241-280. San Diego:<br />

Academic Press.<br />

St<strong>and</strong>en, E.M. 2008. Pelvic fin locomotor<br />

function in fishes: Three-dimensional kinematics<br />

in rainbow trout (Oncorhynchus mykiss).<br />

J Exp Biol. 211:2931-42. doi: 10.1242/<br />

jeb.018572.<br />

St<strong>and</strong>en, E.M. 2010. Muscle activity <strong>and</strong><br />

hydrodynamic function of pelvic fins in<br />

trout (Oncorhynchus mykiss). J Exp Biol.<br />

213:831-41. doi: 10.1242/jeb.033084.<br />

St<strong>and</strong>en, E.M., & Lauder, G.V. 2005. Dorsal<br />

<strong>and</strong> anal fin function in bluegill sunfish<br />

(Lepomis macrochirus): Three-dimensional<br />

kinematics during propulsion <strong>and</strong> maneuvering.<br />

J Exp Biol. 205:2753-63. doi: 10.1242/<br />

jeb.01706.<br />

St<strong>and</strong>en, E.M., & Lauder, G.V. 2007.<br />

Hydrodynamic function of dorsal <strong>and</strong> anal<br />

fins in brook trout (Salvelinus fontinalis).<br />

J Exp Biol. 210:325-39. doi: 10.1242/<br />

jeb.02661.<br />

Syme, D.A. 2006. Functional properties of<br />

skeletal muscle. In: Fish Biomechanics.<br />

Volume 23 in Fish Physiology, eds. Shadwick,<br />

R.E., & Lauder, G.V., 179-240. San Diego:<br />

Academic Press.<br />

Tangorra, J., Phelan, C., Esposito, C., &<br />

Lauder, G. 2011. Biorobotic models of highly<br />

deformable fins for studies of the mechanics<br />

<strong>and</strong> control of fin forces. Integr Comp Biol.<br />

in press. doi: 10.1093/icb/icr036.<br />

Tangorra, J.L., Lauder, G.V., Hunter, I.,<br />

Mittal, R., Madden, P.G., & Bozkurttas, M.<br />

2010. The effect of fin ray flexural rigidity<br />

on the propulsive forces generated<br />

by a biorobotic fish pectoral fin. J Exp<br />

Biol. 213:4043-54. doi: 10.1242/<br />

jeb.048017.<br />

Thomson, K.S. 1971. The adaptation <strong>and</strong><br />

evolution of early fishes. Q Rev Biol. 46:<br />

139-66. doi: 10.1086/406831.<br />

Thomson, K.S. 1976. On the heterocercal<br />

tail in sharks. Paleobiology. 2:19-38.<br />

Troolin, D., & Longmire, E. 2010. Volumetric<br />

velocity measurements of vortex rings from<br />

inclined exits. Exp Fluids. 48:409-20.<br />

doi: 10.1007/s00348-009-0745-z.<br />

Tytell, E.D. 2004. Kinematics <strong>and</strong> hydrodynamics<br />

of linear acceleration in eels,<br />

54 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Anguilla rostrata. P Roy Soc Lond B.<br />

271:2535-40. doi: 10.1098/rspb.2004.2901.<br />

Tytell, E.D. 2006. Median fin function<br />

in bluegill sunfish, Lepomis macrochirus:<br />

Streamwise vortex structure during steady<br />

swimming. J Exp Biol. 209:1516-34.<br />

doi: 10.1242/jeb.02154.<br />

Wilga, C.D., & Lauder, G.V. 2004b.<br />

Hydrodynamic function of the shark’s tail.<br />

Nature. 430:850. doi: 10.1038/430850a.<br />

Tytell, E.D., & Lauder, G.V. 2002. The<br />

c-start escape response of Polypterus senegalus:<br />

Bilateral muscle activity <strong>and</strong> variation during<br />

stage 1 <strong>and</strong> 2. J Exp Biol. 205:2591-603.<br />

Webb, P.W. 1971. The swimming energetics<br />

of trout: I. Thrust <strong>and</strong> power output at<br />

cruising speeds. J Exp Biol. 55:489-520.<br />

Webb, P.W. 2004. Response latencies to<br />

postural disturbances in three species of<br />

teleostean fishes. J Exp Biol. 207:955-61.<br />

doi: 10.1242/jeb.00854.<br />

Webb, P.W., & Keyes, R.S. 1981. Division<br />

of labor between median fins in swimming<br />

dolphin (Pisces: Coryphaenidae). Copeia.<br />

1981:901-04. doi: 10.2307/1444198.<br />

Webb, P.W., Kostecki, P.T., & Stevens,<br />

E.D. 1984. The effect of size <strong>and</strong> swimming<br />

speed on the locomotor kinematics of rainbow<br />

trout. J Exp Biol. 109:77-95.<br />

Weihs, D., Ringel, M., & Victor, M. 2006.<br />

Aerodynamic interactions between adjacent<br />

slender bodies. AIAA J. 44:481-4.<br />

doi: 10.2514/1.18902.<br />

Westneat, M., & Wainwright, S.A. 2001.<br />

Mechanical design for swimming: Muscle,<br />

tendon, <strong>and</strong> bone. In: Tuna: Physiology,<br />

Ecology, <strong>and</strong> Evolution, eds. Block, B., &<br />

Stevens, E.D., 271-311. San Diego: Academic<br />

Press. doi: 10.1016/S1546-5098(01)19008-4.<br />

Wilga, C.D., & Lauder, G.V. 2002. Function<br />

of the heterocercal tail in sharks: Quantitative<br />

wake dynamics during steady horizontal<br />

swimming <strong>and</strong> vertical maneuvering.<br />

J Exp Biol. 205:2365-74.<br />

Wilga, C.D., & Lauder, G.V. 2004a.<br />

Biomechanics of locomotion in sharks, rays<br />

<strong>and</strong> chimeras. In: Biology of Sharks <strong>and</strong> Their<br />

Relatives, eds. Carrier, J.C., Musick, J.A.,<br />

& Heithaus, M.R., 139-64. Boca Raton:<br />

CRC Press.<br />

July/August 2011 Volume 45 Number 4 55


PAPER<br />

Thrust Production in Highly Flexible Pectoral<br />

Fins: A Computational Dissection<br />

AUTHORS<br />

Srinivas Ramakrishnan 1<br />

ANSYS, Inc.<br />

Meliha Bozkurttas<br />

Franklin W. Olin<br />

College of Engineering<br />

Rajat Mittal<br />

Department of Mechanical Engineering,<br />

Johns Hopkins University<br />

George V. Lauder<br />

Department of Organismic<br />

<strong>and</strong> Evolutionary Biology,<br />

Harvard University<br />

Introduction<br />

Robust design based on natural<br />

systems is a significant engineering<br />

challenge. Evolution-based design is<br />

inherently a multi-objective optimization<br />

problem. Natural selection puts<br />

pressure on organisms to produce<br />

locomotion abilities that balance competing<br />

requirements of speed, efficiency,<br />

<strong>and</strong> effectiveness. The goals<br />

of ongoing research efforts are to elucidate<br />

the competing requirements that<br />

have enabled the evolution of highly<br />

maneuverable propulsion/locomotion<br />

at low speeds. Prominent natural systems<br />

of interest are flapping flight in<br />

air <strong>and</strong> aquatic locomotion <strong>and</strong> a common<br />

feature among these systems is<br />

the presence of highly compliant control<br />

surfaces. Organisms that employ<br />

these models of locomotion appear to<br />

exploit the flexibility of their wings/<br />

1 Work presented here was done while the author<br />

was a postdoctoral scientist at The George Washington<br />

University prior to joining Ansys Inc.<br />

ABSTRACT<br />

Bluegill sunfish pectoral fins represent a remarkable success in evolutionary<br />

terms as a means of propulsion in challenging environments. Attempts to mimic<br />

their design in the context of autonomous underwater vehicles have overwhelmingly<br />

relied on the analysis of steady swimming. Experimental observations of<br />

maneuvers reveal that the kinematics of fin <strong>and</strong> wake dynamics exhibit characteristics<br />

that are distinctly different from steady swimming. We present a computational<br />

analysis that compares, qualitatively <strong>and</strong> quantitatively, the wake hydrodynamics<br />

<strong>and</strong> performance of the bluegill sunfish pectoral fin for two modes of swimming:<br />

steady swimming <strong>and</strong> a yaw turn maneuver. It is in this context that we comment on<br />

the role that flexibility plays in the success of the pectoral fin as a versatile propulsor.<br />

Specifically, we assess the performance of the fin by conducting a “virtual dissection”<br />

where only a portion of fin is retained. Approximately 90% of peak thrust for<br />

steady swimming is recovered using only the dorsal half. This figure drops to 70%<br />

for the yaw turn maneuver. Our findings suggest that designs based on fin analysis<br />

that account for various locomotion modes can lead to more robust performance<br />

than those based solely on steady swimming.<br />

Keywords: computational fluid dynamics (CFD), immersed boundary methods<br />

(IBM), bluegill sunfish, biological locomotion<br />

fins to achieve high maneuverability<br />

at low speeds. This paper presents the<br />

analysis of one such control surface:<br />

the bluegill sunfish pectoral fin.<br />

Atypicalsunfish pectoral fin<br />

consists of 14 fin raysasshownin<br />

Figure 1. We see the fin raysnumbered<br />

sequentially starting from the<br />

dorsal edge (ray 1) to the ventral edge<br />

(ray 14). These rays support an asymmetric<br />

planform shape for the pectoral<br />

fin. Figure 3 shows different frames of<br />

the sunfish executing a maneuver from<br />

a ventral view. The motion of the pectoral<br />

fin <strong>and</strong> body are captured using<br />

multiple high-speed video cameras<br />

simultaneously operating at 250 or<br />

more frames per second with a 1024 ×<br />

1024 resolution (Lauder et al., 2006).<br />

The wing surface is digitized at about<br />

300 spatial locations at several points<br />

during the fin cycle.Thus,thekinematics<br />

of the fin motionisacquired<br />

for the simulation. The collaboration<br />

with experimentalists (biologists <strong>and</strong><br />

FIGURE 1<br />

Bluegill sunfish pectoral fin consists of<br />

14 rays, which form the full planform. The<br />

dissected planform is interpolated from rays<br />

1–8 to investigate the flow <strong>and</strong> performance.<br />

56 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Bioinspired design paradigm.<br />

engineers), through a multi-disciplinary<br />

effort (Lauder et al., 2006; Mittal et al.,<br />

2006), has enabled high-fidelity data<br />

to be used in the computational analysis<br />

(see Figure 2).<br />

It is clear from looking at the fin<br />

motion during the maneuver (see Figure<br />

3) that the kinematics of the fin<br />

involves both deformation <strong>and</strong> translation.<br />

This poses severe challenges for<br />

traditional body-fitted computational<br />

methods. Here, the immersed boundary<br />

method, with its ability to h<strong>and</strong>le<br />

complex deforming structures, enables<br />

us to undertake high-fidelity computational<br />

fluid dynamics (CFD) analysis of<br />

the pectoral fin hydrodynamics(see<br />

Computational Methodology). It has<br />

been used to gain valuable insight into<br />

pectoral fin hydrodynamics in steady<br />

swimming (Bozkurttas et al., 2009;<br />

Dong et al., 2010). The experimentally<br />

obtained steady swimming kinematics<br />

was analyzed, <strong>and</strong> an efficient reconstruction<br />

of the kinematics using proper<br />

orthogonal decomposition (POD)<br />

was obtained. The POD modes using<br />

a combination of the first three modes<br />

(hereafterreferredtoasMode1+2+3)<br />

were successful in reproducing two<br />

thirds of the full fin kinematics. More<br />

significantly, this combination of<br />

modes was found to retain 92%<br />

of the thrust produced using the actual<br />

kinematics (Bozkurttas, 2007;<br />

Bozkurttas et al., 2009). Further, detailed<br />

analysis of the pressure distribution<br />

over the full fin surface (rays<br />

1-14) during steady swimming also revealed<br />

that most of the thrust was produced<br />

by the dorsal part mainly around<br />

the spanwise tip region (Bozkurttas<br />

et al., 2009; Dong et al., 2010) (see Figure<br />

5). Since different sections of the<br />

pectoral fin trace different trajectories<br />

during a fin stroke, the contribution<br />

of each region of the fin to its overall<br />

performance may not be uniform. Naturally,<br />

this leads us to the central theme<br />

of this paper, the idea of examining the<br />

thrust production of different sections<br />

of the fin. The goal is to enable a virtual<br />

“dissection” or “ablation” of the pectoral<br />

fin dynamics <strong>and</strong> the effect of this<br />

ablation on the fin performance. It is<br />

expected that this will yield useful insight<br />

into the hydrodynamic function<br />

of the fin in various swimming modes.<br />

FIGURE 3<br />

A bluegill sunfish during a maneuver: ventral (bottom) view. Images are frames from a highspeed<br />

video. Note the differential motion of the left <strong>and</strong> right side fins. Top row: t/T =0,t/T =<br />

0.23, t/T = 0.30. Bottom row: t/T = 0.46, t/T = 0.70, t/T = 0.84.<br />

Computational<br />

Methodology<br />

We present a brief description of<br />

the Cartesian grid-based immersed<br />

boundary method for moving boundaries<br />

starting with the governing<br />

equations. The three-dimensional<br />

unsteady, viscous incompressible<br />

Navier-Stokes equations are given as<br />

∂u i<br />

¼ 0<br />

∂x i<br />

∂u i<br />

∂t þ ∂ u <br />

iu j<br />

¼ 1 ∂p<br />

þ ν ∂ <br />

∂u i<br />

∂x j ρ ∂x i ∂x j ∂x j<br />

ð1Þ<br />

where i; j =1,2,3,u i are the velocity<br />

component, p is the pressure, <strong>and</strong> ρ<br />

July/August 2011 Volume 45 Number 4 57


<strong>and</strong> ν are the fluid density <strong>and</strong> kinematicviscosity.Wehaveemployeda<br />

conventional notation where repeated<br />

indices imply summation.<br />

1. Numerical Method<br />

The Navier-Stokes equations<br />

(Eq. 1) are discretized using a cellcentered,<br />

collocated (non-staggered)<br />

arrangement of the primitive<br />

variables (u i , p). In addition to the<br />

cell-centered velocities (u i ), the<br />

face-centered velocities, U i ,are<br />

computed. A second-order Adams-<br />

Bashforth scheme is employed for<br />

the convective terms while the diffusion<br />

terms are discretized using<br />

an implicit Crank-Nicolson scheme<br />

whicheliminatestheviscousstability<br />

constraint. The spatial derivatives<br />

are computed using a<br />

second-order accurate central difference<br />

scheme. The equations<br />

are integrated in time using the<br />

fractional step method (Chorin,<br />

1967). In the first sub-step of this<br />

method, a modified momentum<br />

equation is solved <strong>and</strong> an intermediate<br />

velocity u* obtained. The second<br />

sub-step requires the solution<br />

of the pressure correction equation<br />

which is solved with the constraint<br />

that the final velocity u i<br />

n+1<br />

be divergence-free. This gives a<br />

Poisson equation for the pressure<br />

correction <strong>and</strong> a Neumann boundary<br />

condition imposed on this pressure<br />

correction at all boundaries.<br />

This Poisson equation is solved<br />

with a highly efficient geometric<br />

multigridmethodwhichemploys<br />

a Gauss-Siedel line-SOR smoother.<br />

Once the pressure correction is obtained,<br />

the pressure <strong>and</strong> velocity are<br />

updated (see Dong et al., 2006 <strong>and</strong><br />

Mittal et al., 2008, for additional<br />

details). These separately updated<br />

face velocities satisfy discrete mass<br />

conservation to machine accuracy<br />

<strong>and</strong> use of these velocities in estimating<br />

the non-linear convective flux<br />

leads to a more accurate <strong>and</strong> robust<br />

solution procedure. The advantage<br />

of separately computing the facecentered<br />

velocities was initially proposed<br />

by Zang et al. (1994) <strong>and</strong><br />

discussed in the context of the<br />

Cartesian grid methods in Ye et al.<br />

(1999) <strong>and</strong> Mittal et al. (2008).<br />

2. Immersed Boundary Treatment<br />

The immersed boundary method<br />

used here employs a multidimensional<br />

ghost cell methodology<br />

to impose the boundary conditions<br />

on the immersed boundary. The<br />

current solver is designed from the<br />

start for fast, efficient, <strong>and</strong> accurate<br />

solution of flows with complex<br />

three-dimensional, moving boundaries.<br />

Also, the current method is<br />

a “sharp“ interface method in that<br />

the boundary conditions on the<br />

immersed boundary are imposed<br />

at the precise location of the immersed<br />

body, <strong>and</strong> there is no spurious<br />

spreading of boundary forcing<br />

into the fluid as what usually occurs<br />

with diffuse interface methods<br />

(Mittal & Iaccarino, 2005).<br />

3. Geometric Representation<br />

of Immersed Boundary<br />

The current method is designed to<br />

simulate flows over arbitrarily complex<br />

2D <strong>and</strong> 3D immersed stationary<br />

<strong>and</strong> moving boundaries <strong>and</strong> the<br />

approach chosen to represent the<br />

boundary surface should be flexible<br />

enough so as not to limit the type<br />

of geometries that can be h<strong>and</strong>led.<br />

A number of different approaches<br />

are available for representing the<br />

surface of the immersed boundary,<br />

including level sets (Osher &<br />

Sethian, 1988; Tran & Udaykumar,<br />

2004), <strong>and</strong> unstructured surface<br />

grids. In the current solver, we<br />

choose to represent the surface of<br />

the immersed boundary by an unstructured<br />

mesh with triangular elements.<br />

This approach is very well<br />

suited for the wide variety of engineering<br />

<strong>and</strong> biological configurations<br />

that are of interest to us <strong>and</strong><br />

is compatible with the immersed<br />

boundary methodology used in<br />

the current solver.<br />

4. Boundary Motion<br />

Boundary motion can be included<br />

into immersed boundary formulation<br />

with relative ease. In advancing<br />

the field equations from time level<br />

n to n + 1 in the case of a moving<br />

boundary, the first step is to move<br />

from its current location to the<br />

new location. This is accomplished<br />

by moving the nodes of the surface<br />

triangles with a known velocity.<br />

Thus, we employ the following<br />

equation to update the coordinates<br />

(X i ) of the surface element vertices,<br />

X nþ1<br />

i<br />

Δt<br />

X n<br />

i<br />

¼ V nþ1<br />

i<br />

ð2Þ<br />

where V i is the vertex velocity. The<br />

vertex velocity can either be prescribed<br />

or it can be computed<br />

from a dynamical equation if the<br />

body motion is coupled to the<br />

fluid. The next step is to determine<br />

the ghost cells for this new immersed<br />

boundary location <strong>and</strong><br />

recompute interpolation weights<br />

associated with the ghost point<br />

methodology. Subsequently, the<br />

flow equations, which are written<br />

in Eulerian form, are advanced<br />

in time. The general framework<br />

described above can, therefore, be<br />

considered as Eulerian-Lagrangian,<br />

wherein the immersed boundaries<br />

are explicitly tracked as surfaces in<br />

a Lagrangian mode, while the flow<br />

computations are performed on a<br />

fixed Eulerian mesh. Additional<br />

58 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


details regarding the current immersed<br />

boundary methodology<br />

may be found in Mittal et al. (2008).<br />

Computational Setup<br />

All simulations are conducted in<br />

a rectangular computational domain.<br />

The boundary conditions on the<br />

bounding box of the domain are freestream<br />

on the left (x direction), outflow<br />

on the right while the remaining<br />

boundaries (top <strong>and</strong> bottom ( y direction)<br />

<strong>and</strong> front <strong>and</strong> back (z direction))<br />

employ slip boundary conditions (see<br />

Figure 4). The fin surface <strong>and</strong> fish<br />

body are considered as no-slip boundaries.<br />

The fins are treated as deforming<br />

membranes while the body, where applicable,<br />

is treated as rigid body undergoing<br />

general motion. The Reynolds<br />

number in the present work is defined<br />

as Re = UL s /ν where U, L s ,<strong>and</strong>ν are<br />

the swimming velocity, spanwise fin<br />

length, <strong>and</strong> the kinematic viscosity of<br />

FIGURE 4<br />

water (ν =1.007×10 −6 m 2 s −1 at room<br />

temperature), respectively.<br />

Based on a swimming speed of<br />

1.1 body length per second, the<br />

Reynolds number for the steady swimming<br />

is 6300. However, a comparison<br />

of the force coefficients obtained<br />

at Re = 1440 with those at the experimental<br />

Reynolds number appear to be<br />

in good agreement both quantitatively<br />

<strong>and</strong> qualitatively (Bozkurttas, 2007).<br />

So, for computational expediency, we<br />

use the lower Reynolds number in the<br />

steady swimming analysis (Dong et al.,<br />

2010). As mentioned earlier, low dimensional<br />

model performance analyses<br />

have shown that Mode 1 + 2 + 3<br />

gait that accounts for 67% of the fin<br />

motion still produces 92% of the<br />

thrust (Bozkurttas et al., 2009). Therefore,<br />

in lieu of the experimentally<br />

extracted fin kinematics, this simplified<br />

model has been used here. The<br />

grid size in these simulations is 153 ×<br />

161 × 97, which is about 2.35 million<br />

Cartesian grid (4.8 million grid points) <strong>and</strong> unstructured mesh employed for yaw maneuver:<br />

(a) x-y plane section, (b) x-z plane section, (c) y-z plane section (strongside fin ontheleft<br />

<strong>and</strong> weakside on right of the body), <strong>and</strong> (d) unstructured surface mesh (pectoral fin only, number<br />

of nodes = 10,000, number of elements = 19,602).<br />

grid points. A domain size of 3.8L s ×<br />

4.5L s ×1.8L s is selected where L s is<br />

the span wise size of the fin. Comprehensive<br />

studies have been carried out<br />

to assess the effect of the grid resolution<br />

<strong>and</strong> domain size on the salient<br />

features of the flow <strong>and</strong> also to demonstrate<br />

the accuracy of the selected grid<br />

(Bozkurttas, 2007).<br />

The Reynolds number for the turning<br />

maneuver based on a freestream<br />

velocity of 0.5 body lengths per second<br />

is approximately 3500. The domain<br />

size employed for the maneuver is<br />

7.5L s ×5L s ×5L s . The pectoral fins<br />

<strong>and</strong> an idealized body, immersed in<br />

the computational grid, are shown in<br />

Figure 4. The nominal grid size used<br />

in the current simulation is 241 ×<br />

145 × 145 (see Figure 4). Finally, the<br />

domain size for the maneuver with<br />

just the strongside (outside) fin is<br />

4L s ×4L s ×4L s with a non-uniform<br />

grid using 128 points in all three<br />

dimensions.<br />

We note in passing that all the steady<br />

swimming cases <strong>and</strong> ablated fin simulations<br />

(for the maneuver) do not include<br />

the fish body. This is reasonable<br />

sincewehaveobservedthatthedifference<br />

in the thrust coefficients with<br />

<strong>and</strong> without the body is minimal. As<br />

we shall see shortly, the wake dynamics<br />

for both steady swimming <strong>and</strong> maneuver<br />

are dominated by vortex structures<br />

generated far from the fish body (see<br />

Figures 5 <strong>and</strong> 8). Thus, the interaction<br />

between the body <strong>and</strong> the fin hydrodynamics<br />

is minimal.<br />

The performance of the fin is evaluated<br />

using the computed force coefficients<br />

which are defined as,<br />

C T ¼<br />

2F x<br />

ρU∞ 2A fin<br />

; C L ¼ 2F y<br />

ρU∞ 2A ;<br />

fin<br />

C Z ¼<br />

2F z<br />

ρU 2 ∞ A fin<br />

ð3Þ<br />

July/August 2011 Volume 45 Number 4 59


FIGURE 5<br />

The anatomy of the principal vortex dynamics<br />

involved in steady swimming.<br />

where F x , F y <strong>and</strong> F z are the forces respectively<br />

in the streamwise (drag/<br />

thrust), vertical (lift), <strong>and</strong> spanwise<br />

(lateral) directions, A fin is the nominal<br />

fin area,<strong>and</strong>ρ is the density of the<br />

fluid. U ∞ is the forward swimming<br />

velocity. The force components are<br />

calculated by directly integrating the<br />

computed pressure <strong>and</strong> shear stress<br />

on the fin surface.<br />

Results<br />

Steady Swimming<br />

A snapshot of the vortex dynamics<br />

at the end of a steady swimming fin<br />

beatisshowninFigure5.Notethe<br />

FIGURE 6<br />

complex interaction of among vortices<br />

generated by the path traversed by the<br />

fin tip during a fin beat. Clearly, both<br />

adduction <strong>and</strong> abduction appear to<br />

produce distinct vortex structures.<br />

This is in stark contrast with a simple<br />

ring vortex created during the maneuver<br />

(see Figure 8). The time variations<br />

of the force coefficients (C T , C L <strong>and</strong><br />

C Z ) for three fin planforms are plotted<br />

in Figure 6. Note the presence of two<br />

distinct <strong>and</strong> comparable peaks corresponding<br />

to the adduction <strong>and</strong> abduction<br />

phases. This force signature<br />

bears the trademark of efficiency where<br />

the fin sustains net forward thrust<br />

throughout its fin beat. Clearly, the<br />

chordwise <strong>and</strong> spanwise compliance<br />

of the fin allows the simultaneous formation<br />

<strong>and</strong> persistence of two distinct<br />

vortex structures within a single fin<br />

beat. A rigid planform would lead to<br />

a more restrictive envelope for the fin<br />

tip path resulting in vortex dynamics<br />

that have stronger interactions detrimental<br />

to sustained net thrust production<br />

(see Akhtar et al., 2007).<br />

We now construct two different<br />

ablated fin models: one that contains<br />

only the rays 1-4 <strong>and</strong> one that contains<br />

rays 1-8 (see Figure 1). The motion of<br />

these dissected fins is precisely the<br />

same as that for the full fin <strong>and</strong>we<br />

carry out flow simulations for both of<br />

these cases. Examining the results from<br />

our virtual dissection, we notice that<br />

the dorsal half of the fin (rays 1-8) captures<br />

the two main peaks of the thrust<br />

<strong>and</strong> preserves 90% of the thrust production<br />

of the full fin planform. Consequently,<br />

the ventral contribution of<br />

the fin, represented by rays 9-14 in<br />

Figure 1, to the thrust production is<br />

found to be insignificant. Also, the<br />

planform interpolated from rays 1-4<br />

has a similar trend in thrust variation<br />

during the entire fin-beat cycle albeit<br />

with smaller amplitudes. Interestingly,<br />

it has two main peaks <strong>and</strong> even the two<br />

local peaks in the abduction phase as in<br />

the full fin case. This further reinforces<br />

the notion that the dorsal leading edge<br />

of the bluegill’s pectoral fin dominates<br />

the overall performance during steady<br />

swimming propulsion. This planform<br />

produces almost 40% of the thrust<br />

produced by the fish fin while undergoing<br />

Mode 1 + 2 + 3 gait. Finally,<br />

we observe similar tendencies for lift<br />

<strong>and</strong> spanwise force coefficients for the<br />

three planforms, except the case with<br />

just rays 1-4 where the values show attenuation.<br />

The key observation here is<br />

that the dorsal half of the pectoral fin<br />

Comparison of time variation of force coefficients for three different fin planforms (rays 1–4, rays 1–8, full planform) at Mode 1+2+3gait:(a)<br />

streamwise force, (b) vertical force, <strong>and</strong> (c) lateral force.<br />

60 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


(rays 1-8) is responsible for producing<br />

a majority of the thrust. These results<br />

bring into question the need for the<br />

ventral portion of the fin. We explore<br />

this in detail as we consider the case of<br />

the yaw turn maneuver.<br />

FIGURE 7<br />

Comparison of time variation of force coefficients: (a) strongside <strong>and</strong> (b) weakside.<br />

Yaw Turn Maneuver<br />

The evolution of wake structure<br />

from the strongside fin, that drives<br />

the maneuver, is shown from two vantage<br />

points: lateral (Figure 8 (a,c,e))<br />

<strong>and</strong>dorsal(Figure8(b,d,f)).The<br />

well-defined vortex ring formed during<br />

the outstroke (abduction) produces a<br />

lateral jet oriented normal to the fish<br />

body (see Figure 9). This type of vortex<br />

ring <strong>and</strong> associated lateral jet shown in<br />

Figures 8 <strong>and</strong> 9 have also been observed<br />

in experimental visualization<br />

(Drucker & Lauder, 2001). The peak<br />

lateral velocity is found to be greater<br />

than three times the freestream velocity.<br />

Consequently, the lateral forces<br />

developed are several times that observed<br />

in forward thrust for the steady<br />

swimming case (Bozkurttas, 2007).<br />

Preliminary estimates for stroke-averaged<br />

force coefficients ratio between<br />

lateral force in maneuvering ( ― C Z =<br />

6.1) to steady swimming thrust ( ― C T =<br />

1.29) is approximately 4 (( ‐ ) denotes<br />

averageoverstroke).Thisfactorisin<br />

reasonable agreement with the forces<br />

measured experimentally (Drucker &<br />

Lauder, 2001).<br />

Returning to Figure 7(a), we note<br />

that the C Z peak is reached between<br />

t/T = 0.15 <strong>and</strong> t/T = 0.3. Shortly thereafter,<br />

the C T peak occurs between<br />

t/T = 0.3 <strong>and</strong> t/T = 0.4. As expected,<br />

the first priority in the maneuver is<br />

to evade the stimulus (an obstacle or<br />

predator in the wild) by quickly generating<br />

a strong lateral force (maximum<br />

occurs at t/T = 0.2). Thereafter, the<br />

drag force developed in the streamwise<br />

direction is likely used to modulate the<br />

direction of the resultant force as the<br />

sunfish turns away from the stimulus.<br />

The evolving vortex ring, clearly seen<br />

in Figures 8(d) <strong>and</strong> 8(f ), continues to<br />

be oriented nearly parallel to the fish<br />

body. Consequently, the lateral jet orientation<br />

ensures that the maximum<br />

lateral force continues to act normal<br />

to the fish body for the duration of<br />

the maneuver. Here, the inherent flexibility<br />

of the pectoral fin structure <strong>and</strong><br />

the ability to continuously alter planform<br />

area is likely to be very useful.<br />

Finally, an examination of the force<br />

histories for the dissected fin reveals<br />

that the peak lateral thrust developed<br />

by the dorsal part (rays 1-8) is approximately<br />

70% of the total as opposed to<br />

30% for the ventral (rays 8-14) portion<br />

(see Figure 10 <strong>and</strong> Figure 11c). The<br />

streamwise drag is slightly more comparable,<br />

although the dorsal part peak is<br />

higher (see Figure 11a). Overall, while<br />

the dorsal portion contributes to the majority<br />

of lateral force production, the ratio<br />

of dorsal to ventral contribution appears<br />

to be more equitable than the steady<br />

swimming case.<br />

Conclusions<br />

A comparative analysis of the pectoral<br />

fin performance in steady swimming<br />

<strong>and</strong> yaw turn maneuver reveals<br />

that the dorsal part of the pectoral fin<br />

is responsible for the majority of force<br />

production. The chordwise <strong>and</strong> spanwise<br />

flexibility of the pectoral fin <strong>and</strong><br />

its ability to have them function either<br />

in concert or independently seems to<br />

enable the bluegill sunfish to achieve<br />

a variety of maneuvers. The virtual<br />

dissection reveals a significant loss of<br />

performance with maneuvering with<br />

respect to peak lateral thrust when<br />

the ventral portion is removed. Thus,<br />

a fin design using just the dorsal portion<br />

of the pectoral fin might perform<br />

as well as the full fin insteadyswimming<br />

but will not retain the same<br />

maneuverability. Hence, any effective<br />

design based on the pectoral fin that<br />

aims to preserve its performance over<br />

all locomotion mode needs to retain<br />

a greater portion of the fin thanthat<br />

suggested by steady swimming alone.<br />

The pectoral fins of fishes display a<br />

diversity of shapes (e.g., Drucker &<br />

Lauder, 2002; Thorsen & Westneat,<br />

2005), <strong>and</strong> although some general<br />

conclusions about correlations of fin<br />

shape with fish ecology have been possible<br />

(see Wainwright et al., 2002),<br />

there are very few data on functional<br />

regionalization of pectoral fins <strong>and</strong> on<br />

therolethatdifferentfin rays within<br />

July/August 2011 Volume 45 Number 4 61


FIGURE 8<br />

Formation of the vortex ring due to the strongside pectoral fin motion: (a), (c), <strong>and</strong> (e) are lateral views at t/T = 0.22, t/T = 0.49, <strong>and</strong> t/T = 0.66,<br />

respectively; (b), (d), <strong>and</strong> (f) are the corresponding dorsal views at t/T = 0.22, t/T = 0.49, <strong>and</strong> t/T = 0.66, respectively.<br />

62 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 9<br />

The strongside lateral jet associated with the vortex structures in Figure 8 (c) at t/T = 0.49.<br />

FIGURE 10<br />

Formation of the vortex ring due to the strongside pectoral fin motion: (a) full, (b) dorsal, <strong>and</strong><br />

(c) ventral portion of the fin sections.<br />

the pectoral fin might play in controlling<br />

locomotor performance. Taft et al.<br />

(2008) discussed functional regionalization<br />

during steady swimming in<br />

sculpin, but the role that different fin<br />

rays play during maneuvering behaviors<br />

has not previously been analyzed.<br />

The results presented here suggest that<br />

the ventral region of the fin playsan<br />

important role in modulating maneuvering<br />

forces, <strong>and</strong> future studies on the<br />

diversity of fish pectoral fin shapes<br />

could focus on the surface area <strong>and</strong><br />

mechanical properties of this region<br />

of the fin incorrelationwithmaneuvering<br />

performance. No data are currently<br />

available that would permit<br />

even general conclusions about the diversification<br />

of pectoral fin structure in<br />

relation to maneuvering capability,<br />

<strong>and</strong> this represents a new <strong>and</strong> very interesting<br />

direction for future work that<br />

integrates approaches from biomechanics<br />

<strong>and</strong> fluid dynamics with behavioral<br />

<strong>and</strong> ecological studies of fish<br />

locomotion.<br />

Acknowledgments<br />

This work was done while the first<br />

three authors were at The George<br />

FIGURE 11<br />

Comparison of forces produced on the dorsal <strong>and</strong> ventral halves of the strongside fin with respect to the full fin: (a) streamwise force, (b) vertical<br />

force, <strong>and</strong> (c) lateral force.<br />

July/August 2011 Volume 45 Number 4 63


Washington University, <strong>and</strong> the work<br />

wassupportedunderONR-MURI<br />

grant N00014-03-1-0897 monitored<br />

by Dr. Thomas McKenna.<br />

Lead Author:<br />

Srinivas Ramakrishnan<br />

ANSYS, Inc.<br />

10 Cavendish Court,<br />

Lebanon, NH 03766<br />

Email: srinivas.ramakrishnan@<br />

ansys.com<br />

References<br />

Akhtar, I., Mittal, R., Lauder, G.V., &<br />

Drucker, E.G. 2007. Hydrodynamics of a<br />

biologically inspired t<strong>and</strong>em flapping foil<br />

configuration. Theor Comp Fluid Dyn.<br />

21:155-70. doi: 10.1007/s00162-007-0045-2.<br />

Bozkurttas, M. 2007. Hydrodynamic performance<br />

of fish pectoral fins with application to<br />

autonomous underwater vehicles. Ph.D. Thesis,<br />

The George Washington University. p. 9.<br />

Bozkurttas, M., Mittal, R., Dong, H.,<br />

Lauder, G.V., & Madden, P. 2009. A lowdimensional<br />

models <strong>and</strong> performance scaling<br />

of a highly deformable fish pectoral fin.<br />

J Fluid Mech. 631:311-42. doi: 10.1017/<br />

S0022112009007046.<br />

Chorin, A.J. 1967. A numerical method<br />

for solving incompressible viscous flow<br />

problems. J Comput Phys. 2:1-9.<br />

doi: 10.1016/0021-9991(67)90037-X.<br />

Dong, H., Bozkurttas, M., Mittal, R., Lauder,<br />

G.V., & Madden, P. 2010. A Computational<br />

modelling <strong>and</strong> analysis of the hydrodynamics<br />

of a highly deformable fish pectoral fin.<br />

J Fluid Mech. 645:345-73. doi: 10.1017/<br />

S0022112009992941.<br />

Dong, H., Mittal, R., & Najjar, F.M. 2006.<br />

Wake topology <strong>and</strong> hydrodynamic performance<br />

of low-aspect-ratio flapping foils.<br />

J Fluid Mech. 566:309-43. doi: 10.1017/<br />

S002211200600190X.<br />

Drucker, E.G., & Lauder, G.V. 2001. Wake<br />

dynamics <strong>and</strong> fluid forces of turning maneuvers<br />

in sunfish. J Exp Biol. 204:431-42.<br />

Drucker, E.G., & Lauder, G.V. 2002. Experimental<br />

hydrodynamics of fish locomotion:<br />

Functional insights from wake visualization.<br />

Integr Comp Biol. 42:243-57. doi: 10.1093/<br />

icb/42.2.243.<br />

Lauder, G.V., Madden, P.G.A., Mittal, R.,<br />

Dong, H., & Bozkurttas, M. 2006. Locomotion<br />

with flexible propulsors: I. Experimental<br />

analysis of pectoral fin swimming in sunfish.<br />

Bioinspir Biomim. 1:35-41. doi: 10.1088/<br />

1748-3182/1/4/S04.<br />

Mittal, R., Dong, H., Bozkurttas, M., &<br />

Lauder, G.V. 2006. Locomotion with flexible<br />

propulsors: II. Computational modeling of<br />

pectoral fin swimming in sunfish. Bioinspir<br />

Biomim. 1:25-34. doi: 10.1088/1748-3182/<br />

1/4/S05.<br />

Mittal, R., Dong, H., Bozkurttas, M., Najjar,<br />

F.M., Vargas, A., & Von Loebbecke, A. 2008.<br />

A versatile sharp interface boundary method<br />

for incompressible flows with complex boundaries.<br />

J Comput Phys. 227:1-9. doi: 10.1016/<br />

j.jcp.2008.01.028.<br />

Mittal, R., & Iaccarino, G. 2005. Immersed<br />

boundary methods. Annu Rev Fluid Mech.<br />

37:239-61. doi: 10.1146/annurev.fluid.37.<br />

061903.175743.<br />

Osher, S., & Sethian, J.A. 1988. Fronts<br />

propagating with curvature-dependent.<br />

J Comput Phys. 79:12-49. doi: 10.1016/<br />

0021-9991(88)90002-2.<br />

Taft, N., Lauder, G.V., & Madden, P.G.<br />

2008. Functional regionalization of the<br />

pectoral fin of the benthic longhorn sculpin<br />

during station holding <strong>and</strong> swimming.<br />

J Zool Lond. 276:159-67. doi: 10.1111/<br />

j.1469-7998.2008.00472.x.<br />

Thorsen, D.H., & Westneat, M. 2005.<br />

Diversity of pectoral fin structure <strong>and</strong> function<br />

in fishes with labriform propulsion. J Morphol.<br />

263:133-50. doi: 10.1002/jmor.10173.<br />

Tran, L.B., & Udaykumar, H.S. 2004. A<br />

particle-level set-based sharp interface cartesian<br />

grid method for impact, penetration, <strong>and</strong><br />

void collapse. J Comput Phys. 193:469-510.<br />

doi: 10.1016/j.jcp.2003.07.023.<br />

Wainwright, P., Bellwood, D.R., & Westneat,<br />

M. 2002. Ecomorphology of locomotion in<br />

labrid fishes. Environ Biol Fish. 65:47-62.<br />

doi: 10.1023/A:1019671131001.<br />

Ye, T., Mittal, R., Udaykumar, H.S., &<br />

Shyy, W. 1999. An accurate Cartesian grid<br />

method for simulation of viscous incompressible<br />

flows with complex immersed boundaries.<br />

J Comput Phys. 156:209-40. doi: 10.1006/<br />

jcph.1999.6356.<br />

Zang, Y., Streett, R.L., & Koseff, J.R. 1994.<br />

A non-staggered fractional step method for<br />

time-dependent incompressible Navier-<br />

Stokes equations in curvilinear coordinates.<br />

J Comput Phys. 114:18-33. doi: 10.1006/<br />

jcph.1994.1146.<br />

64 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Learning From the Fins of Ray-Finned<br />

Fish for the Propulsors of Unmanned<br />

Undersea Vehicles<br />

AUTHORS<br />

James L. Tangorra<br />

Department of Mechanical<br />

Engineering, Drexel University<br />

Timo Gericke<br />

George V. Lauder<br />

Museum of Comparative Zoology,<br />

Harvard University<br />

Introduction<br />

Military <strong>and</strong> civilian studies have<br />

identified that two of the most significant<br />

technological obstacles to deploying<br />

unmanned undersea vehicles<br />

(UUVs) are energy <strong>and</strong> autonomy<br />

(Nicholson & Healey, 2008; Office<br />

of the Secretary of Defense, 2009).<br />

The energy <strong>and</strong> the rate it is used<br />

(power) limit the duration <strong>and</strong> distance<br />

of operations <strong>and</strong> bound the<br />

type of activity that can occur even<br />

for short periods. Autonomy defines<br />

the degree to which humans must supervise<br />

UUV operations <strong>and</strong> provides<br />

UUVs with the ability to react to<br />

external stimuli without human intervention.<br />

Among the enabling technologies<br />

that are critical for solving the<br />

challenges associated with energy <strong>and</strong><br />

autonomyaremoreeffectivepropulsors<br />

(Office of the Secretary Defense,<br />

2009). Propulsors are required to provide<br />

increased maneuverability, stealth,<br />

<strong>and</strong> endurance for the widespread range<br />

of missions envisioned for UUVs, from<br />

long duration sensing in the open ocean<br />

to mine countermeasures in very shallow,<br />

high-energy water.<br />

ABSTRACT<br />

Advanced propulsors are required to help unmanned undersea vehicles (UUVs)<br />

overcome major challenges associated with energy <strong>and</strong> autonomy. The fins of rayfinned<br />

fish provide an excellent model from which to develop propulsors that can<br />

create forces efficiently <strong>and</strong> drive a wide range of behaviors, from hover to lowspeed<br />

maneuvers to high-speed travel. Although much is known about the mechanics<br />

of fins, little is known about the fin’s sensorimotor systems or how fins<br />

are regulated in response to external disturbances. This information is crucial for<br />

implementing propulsive <strong>and</strong> control systems that exploit the same phenomena as<br />

the biological fins for efficiency, effectiveness, <strong>and</strong> autonomous regulation. Experiments<br />

were conducted to evaluate the in vivo response of the sunfish <strong>and</strong> its pectoral<br />

fins to vortex perturbations applied directly to the fish <strong>and</strong> to the fins. The fish<br />

<strong>and</strong> the fins responded actively to perturbations that disturbed the motion of the fish<br />

body. Surprisingly, perturbations that deformed the fins extensively did not cause a<br />

reaction from either the fins or the body. These results indicate that the response of the<br />

pectoral fins to large deformations is not reflexive <strong>and</strong> that fin motions are regulated<br />

when it is necessary to correct for disturbances to the motion of the fish. The results<br />

also demonstrate a benefit of compliance in propulsors, in that external perturbations<br />

can disturb the fins without having its impact be transferred to the fish body.<br />

Keywords: biorobotics, flapping fins, vortex pertubations, sensory-based control<br />

Fish are important biological models<br />

from which to learn methods of<br />

propulsion that are effective <strong>and</strong> efficient<br />

over a wide range of operating<br />

conditions. Bony fish, such as the<br />

bluegill sunfish (Lepomis macrochirus)<br />

<strong>and</strong> the swordfish (Xiphias gladius),<br />

are able to hover, swim <strong>and</strong> maneuver<br />

at low speeds, manipulate the orientation<br />

of their bodies, conduct acrobatics<br />

to escape or to attack prey, <strong>and</strong>, especially<br />

for the swordfish, sustain high<br />

swimming speeds. These behaviors<br />

can be accomplished in smooth water<br />

<strong>and</strong> in high-energy flows <strong>and</strong> relate<br />

directly to the behaviors desired for<br />

UUVs. The remarkable swimming<br />

abilities of these fish are due, in large<br />

part, to the fish having multiple, highly<br />

actuated, flexible fins that are able to<br />

create <strong>and</strong> to modulate large-magnitude<br />

forces.<br />

A great deal is known about the<br />

mechanisms that contribute to the<br />

production of hydrodynamic forces<br />

by flapping the fins <strong>and</strong> the fish<br />

body. Forces are created through the<br />

dynamic interaction of the fins, the<br />

body, <strong>and</strong> the fluid, which results in<br />

energy being added to, or taken from,<br />

the fluid. A review of seminal work<br />

that explains the way in which marine<br />

animals control vorticity is presented<br />

in Triantafyllou et al. (2002) <strong>and</strong><br />

July/August 2011 Volume 45 Number 4 65


Zhu et al. (2002). Numerical <strong>and</strong><br />

experimental studies of flexible fins<br />

with two-dimensional kinematics<br />

(heaving <strong>and</strong> pitching) include, but<br />

are in no way limited to, studies<br />

of McHenry (1995), Liu <strong>and</strong> Bose<br />

(1997), Prempraneerach et al. (2003),<br />

Triantafyllou et al. (2005), Fish et al.<br />

(2006), Lauder et al. (2006), Mittal<br />

et al. (2006), Lauder <strong>and</strong> Madden<br />

(2007), <strong>and</strong> Zhu <strong>and</strong> Shoele (2008).<br />

Recent studies that considered deformable<br />

fins with complex kinematics<br />

are presented in, for example, Shoele<br />

<strong>and</strong> Zhu (2009), Dong et al. (2010),<br />

<strong>and</strong> Tangorra et al. (2010).<br />

In contrast to our underst<strong>and</strong>ing<br />

of the mechanics of fins <strong>and</strong> of hydrodynamic<br />

forces, little is known about<br />

how fishes sense their interaction<br />

with the water <strong>and</strong> use sensory information<br />

to regulate the fins. Knowledge<br />

of fin sensorimotor control is critical<br />

if engineered systems are to take full<br />

advantage of the mechanisms used by<br />

fins to create forces efficiently <strong>and</strong> to<br />

react to changes in the environment.<br />

The focus of this paper will be on<br />

the pectoral fins of sunfish <strong>and</strong>, in particular,<br />

on how <strong>and</strong> when sunfish alter<br />

the use of the pectoral fins in response<br />

to external perturbations. We begin<br />

with an overview of pectoral fin swimming<br />

in sunfish <strong>and</strong> briefly present robotic<br />

fins that produce <strong>and</strong> modulate<br />

forces like the biological fins. A series<br />

of experiments where the biological<br />

fin is perturbed during steady swimming<br />

is then presented. These experiments<br />

address the response of the fins<br />

in the context of using the fins to control<br />

<strong>and</strong> stabilize the fish body.<br />

Ray-Finned Fish <strong>and</strong> Robots<br />

Sunfish Swimming<br />

The ability of the sunfish to control<br />

the magnitude <strong>and</strong> direction of its propulsive<br />

forces is due to its ability to<br />

modulate the kinematics, coordination,<br />

<strong>and</strong> mechanical properties of<br />

its fins <strong>and</strong> muscular tail (Tangorra<br />

et al., 2010, 2011). Hydrodynamic<br />

forces are created through an exchange<br />

of energy between the propulsive surfaces<br />

<strong>and</strong> the surrounding fluid. As the<br />

fish moves through the water, vortices<br />

develop along the body <strong>and</strong> fins, the<br />

propulsive structures bend <strong>and</strong> store<br />

energy, <strong>and</strong> the vortices are shed into<br />

the flow along with directed jets<br />

(Triantafyllou et al., 2002; Dong<br />

et al., 2010). The complex motions<br />

that cause this exchange of energy are<br />

the result of driven motions of the<br />

fin rays <strong>and</strong> a dynamic interaction of<br />

the deformable fin surfaces with the<br />

water. The forces created by fins are,<br />

therefore, modulated through changes<br />

to the kinematics of the fin <strong>and</strong> active<br />

adjustments of fin’s mechanical properties<br />

(Lauder et al., 2006; Mittal et al.,<br />

2006; Akhtar et al., 2007; Tangorra<br />

et al., 2010). The changes may be subtle,<br />

as in steady swimming where the<br />

stiffness of the fin rays is gradually increased<br />

with speed, but where the motions<br />

of the fins are approximately the<br />

same. Or the changes may be obvious,<br />

as when the fish interrupts a cyclic<br />

swimming pattern <strong>and</strong> uses a stiff, impulsive<br />

fin motion to slow the fish <strong>and</strong><br />

turn it away from an obstacle (Gottlieb<br />

et al., 2010).<br />

Ray-Finned Robotic Systems<br />

Robotic fins(Figure1)havebeen<br />

developed that produce motions,<br />

forces, <strong>and</strong> flows like the biological<br />

fins (Tangorra, Davidson et al., 2007;<br />

Phelan, Tangorra et al., 2010; Tangorra,<br />

Lauder et al., 2010). These fins were<br />

designed originally as physical models<br />

with which to conduct experimental<br />

studies that would have been difficult<br />

to conduct with the living fish<br />

FIGURE 1<br />

Biorobotic models of the sunfish pectoral fin<br />

(A) <strong>and</strong> caudal fin (B). The pectoral fin is instrumented<br />

with strain gages along the fin<br />

rays <strong>and</strong> pressure sensors along the body<br />

plate in order to model distributed sensing in<br />

the sunfish. Modified versions of the robotic<br />

pectoral <strong>and</strong> caudal fins,aswellasdorsal<strong>and</strong><br />

anal fins, are implemented on a fish robot (C).<br />

The fish robot can swim freely or be attached<br />

to a rigid mast (shown) so that forces can be<br />

measured. The grooves in the side of the fish<br />

body are used for pressure lines <strong>and</strong> ports.<br />

(Tangorra, Phelan et al., 2011). The<br />

fins are comprised of fin rays, each<br />

with multiple actuated degrees of freedom<br />

(DOF), within a thin, flexible<br />

webbing. The geometries of the fin<br />

rays were defined so that the stiffness<br />

of the robotic fin was proportional to<br />

that of the biological fin across the<br />

fin’s chord <strong>and</strong> span. The architecture<br />

of the robotic fin provides a great degree<br />

of control over the fin’s motions<br />

<strong>and</strong> mechanical properties, which enables<br />

the magnitude <strong>and</strong> direction of<br />

the force produced by the fin tobe<br />

easily modulated (Figure 2). Gross<br />

changes to the profile of the fin’s force<br />

can be made by changing the fin’s gait<br />

66 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Thrust (horizontal) <strong>and</strong> lift (vertical) forces for<br />

pectoral fins executing normal <strong>and</strong> modified<br />

steady swimming gaits. By making relatively<br />

small changes to fin stiffness (A) <strong>and</strong> fin motions<br />

(B), the forces can be moved throughout<br />

the thrust-lift plane. Normal, full-fin steady<br />

swimming for three levels of stiffness (A). The<br />

gait was modified slightly by altering the phase<br />

angle between fin rays by 30° (B, red) or by<br />

using just the upper or lower half of the fin<br />

(B, green). The magnitude of the force can<br />

also be altered simply by changing the frequency<br />

of the fin beat.<br />

pattern, for example, by switching<br />

from a steady swimming gait to the<br />

pattern used by the fish for a turn maneuver.<br />

Smaller changes to the force<br />

profile can be made by changing the<br />

frequency of the fin beat <strong>and</strong>/or by<br />

changing phase relationships between<br />

fin rays (Figure 2A). Considerable<br />

changes to the magnitude <strong>and</strong> direction<br />

of the force can also be made by<br />

adjusting the mechanical properties<br />

of some, or all, of the fin rays. When<br />

the mechanical properties of the fin<br />

rays are under active control, as in<br />

the fish, changes to the force profile<br />

can happen very quickly since the<br />

driven motions of the fin do not have<br />

to be changed.<br />

The designs of the robotic fins were<br />

modified <strong>and</strong> the fins implemented<br />

on a freely swimming biorobotic fish<br />

(Figure 1). Modifications included<br />

placing actuators within the fish body<br />

adjacent to each fin, using a network of<br />

microcontrollers to drive fin motions,<br />

<strong>and</strong> minimizing the number of actuated<br />

DOF for each fin ray.Themotions<br />

<strong>and</strong> orientation of the robotic<br />

fish are controlled by adjusting the<br />

propulsive forces created by five rayfinned<br />

fins. In this first implementation,<br />

the forces are modulated by<br />

switching between several fin gaits<br />

<strong>and</strong> by making predetermined changes<br />

to fin beat frequencies <strong>and</strong> to the phase<br />

relationships between fin rays.<br />

Sensory-Based Control of Fins<br />

What is clearly missing in this robotic<br />

system is the ability to automatically<br />

modulate the kinematics <strong>and</strong><br />

mechanical properties of the fins<br />

based on sensory information about<br />

the fins <strong>and</strong> their interaction with the<br />

water. The motions of the fins are adjusted<br />

based on the forces required to<br />

control the robot’s body, but sensory<br />

information is not being used to exploit<br />

the phenomena that are critical<br />

to the efficient production of force<br />

(e.g., vorticity) nor to adjust behaviors<br />

in response to changes in the flow (e.g.,<br />

speed <strong>and</strong> turbulence). This is due<br />

to the fact that very little is known<br />

about the sensory-based control of<br />

ray-finned fins (Phelan et al., 2010).<br />

The fine level of control that the sunfish<br />

has over fin motions <strong>and</strong> mechanical<br />

properties suggests strongly that<br />

there is closed-loop control of the fins.<br />

However, fundamental questions<br />

about the existence of sensory systems<br />

intrinsic to fins, about the types of<br />

stimuli that elicit responses from fins,<br />

about information in the flow that<br />

is relevant to propulsive forces, <strong>and</strong><br />

about the behavior of the fins in response<br />

to external perturbations have<br />

not yet been answered. This knowledge<br />

is vital for the development of<br />

fin-based propulsors that take advantage<br />

of the phenomena used by fish<br />

to produce forces efficiently <strong>and</strong> that<br />

automatically adjust their behavior in<br />

response to disturbances <strong>and</strong> changes<br />

in operating requirements.<br />

Experimental Methods<br />

<strong>and</strong> Equipment<br />

Experimentation<br />

Experiments were conducted to<br />

evaluate the response of the sunfish’s<br />

pectoral fins to external perturbations<br />

applied to the fin <strong>and</strong>tothefish’s body<br />

during steady swimming. Perturbations<br />

were created using a vortex generator<br />

(Figure 3), which produces a<br />

vortex ring that moves through the<br />

water<strong>and</strong>impartsashortduration<br />

impulse to the fish (Figure 4). The<br />

strength of the vortex was sufficient<br />

to deform the pectoral fin ortodisplace<br />

the fish laterally by several millimeters.<br />

The vortex is not visible, so<br />

it does not elicit a visually mediated response<br />

from the fish. The vortex does,<br />

however, produce a pressure wave that<br />

may be sensed by the fish.<br />

Two bluegill sunfish, with body<br />

lengthsof160±10mm<strong>and</strong>intact<br />

pectoral fins, were used for the experiments.<br />

For the experimental trials,<br />

FIGURE 3<br />

The vortex ring generator <strong>and</strong> vortex (left).<br />

Blue dye was added to the vortex generator’s<br />

cavity to make the vortex visible to the naked<br />

eye. The vortex generator comprises an orifice<br />

plate (1), two cavity plates (2), a latex membrane<br />

(3), <strong>and</strong> a connector plate (4), which<br />

enables the air-line to be connected to the vortex<br />

generator.<br />

July/August 2011 Volume 45 Number 4 67


FIGURE 4<br />

Sunfish in flow tank with vortex generator (A). The sunfish kindly positioned itself in the center<br />

of the test area <strong>and</strong> laser sheet (B). The laser sheet is used with PIV to characterize the vortex as<br />

it travels toward the fish.<br />

asunfish was placed in the working<br />

area (280 × 280 × 800 mm) of a<br />

600-l flow tank <strong>and</strong> was allowed to acclimate<br />

for 2 h. The flow rate was set to<br />

100 mm s −1 , which equates to a steady<br />

swimming speed of approximately<br />

0.6 body lengths s −1 .Atthisspeed,<br />

sunfish generate swimming forces<br />

using primarily their pectoral fins. The<br />

tail <strong>and</strong> the caudal, anal, dorsal, <strong>and</strong><br />

paired pelvic fins are moved very little<br />

but are important for stability. The<br />

vortex generator was positioned approximately<br />

150 mm above the tank<br />

floor <strong>and</strong> placed either perpendicular<br />

to the fish in order to perturb the fish’s<br />

body or at a 45° angle to the fish in<br />

order to perturb the pectoral fin during<br />

its outstroke. A horizontal light sheet<br />

(Figure 4) used for particle image<br />

velocimetry (PIV) was positioned so<br />

that it had the same height as the middle<br />

of the vortex generator. The fish<br />

was directed into the middle of the<br />

test area <strong>and</strong> light sheet by coaxing it<br />

with a wooden dowel. Once the fish<br />

was positioned properly, the vortex<br />

was launched to strike the fish. Vortices<br />

impacted the fish (1) on the body<br />

near the tip of the left pectoral fin<br />

while the fin rested against the body<br />

during the pause between fin beats<br />

<strong>and</strong> (2) at the tip of the left pectoral<br />

fin as the fin completed its outstroke.<br />

High-speed (500 fps), highdefinition<br />

video (1024 × 1024 pixels)<br />

was used to capture the motions of<br />

the fish <strong>and</strong> of the fish’s fins. Two<br />

cameras (Photron 1024 PCI, Photron<br />

USA, Inc., San Diego, CA) were synchronized<br />

<strong>and</strong> positioned so that the<br />

ventral <strong>and</strong> posterior views of the fish<br />

were captured.<br />

Analysis<br />

The linear <strong>and</strong> rotational velocities<br />

of the vortices were analyzed using<br />

DaVis (LaVision GmbH, Göttingen,<br />

Germany).<br />

The motions of the fish <strong>and</strong> of the<br />

fins were analyzed for two fin beats before<br />

<strong>and</strong> two beats after the impact of<br />

the vortex. The coordinates of eight<br />

points along the fish body <strong>and</strong> pectoral<br />

were digitized using Matlab (The<br />

Mathworks Inc., Natick, MA) <strong>and</strong><br />

tracked through time. Deformations<br />

<strong>and</strong> curvatures were calculated for<br />

the pectoral fin during the impact of<br />

thevortexring.Threepointsalongthe<br />

fin were selected to characterize the<br />

shape of the fin <strong>and</strong>todefine the radius<br />

of curvature.<br />

Design of the Vortex<br />

Ring Generator<br />

Vortex rings are commonly generated<br />

using a piston that moves within<br />

a cylindrical cavity <strong>and</strong> pushes a volume<br />

of fluid (the slug) out of the cavity<br />

<strong>and</strong> past an orifice with sharp<br />

edges. The movement of the piston<br />

causes the boundary layer that develops<br />

in the cavity to separate at the orifice’s<br />

edge <strong>and</strong> to roll up into a vortex<br />

ring that has a toroidal shape. The<br />

speed of the piston, the diameter of<br />

the orifice pate, <strong>and</strong> the ratio of cavity<br />

length to cavity diameter influence the<br />

formation of the vortex <strong>and</strong> the speed<br />

at which the vortex travels. Excellent<br />

discussions of vortex generation are<br />

presented in Gharib et al. (1998),<br />

Allen <strong>and</strong> Auvity (2002), Shusser<br />

et al. (2002), <strong>and</strong> Mohseni (2006).<br />

The vortex generator that was developed<br />

for our experiments is similar<br />

to a piston based vortex generator,<br />

but the design was modified so that it<br />

would be more appropriate for the<br />

testing of swimming fish. Two requirements<br />

that influenced the design were<br />

(1) the vortex generator had to be silent,<br />

so that the fish did not hear a<br />

mechanism <strong>and</strong> anticipate the arrival<br />

of the vortex, <strong>and</strong> (2) the system had<br />

to be small, so that it could be placed<br />

at the side of the flow tank without interfering<br />

with the swimming fish. The<br />

vortex generator consists of two acrylic<br />

plates(45×55×12mm)inwhich<br />

a cylindrical cavity is cut (Figure 3).<br />

The plates are covered by a 0.3-mm<br />

thick aluminum plate with either<br />

a 4.0- or 7.5-mm diameter orifice. A<br />

latex membrane is s<strong>and</strong>wiched between<br />

the cavity plates <strong>and</strong> another<br />

acrylic plate in which a cylindrical<br />

well is cut. This plate is connected via<br />

a 6-mm diameter air line (Polyurethane<br />

Tubing, NewWay Air Bearings, Aston,<br />

PA) to a 60-ml syringe (Becton<br />

Dickinson <strong>and</strong> Company, Franklin<br />

Lakes, NJ). A fast push on the syringe<br />

plunger causes the latex membrane to<br />

exp<strong>and</strong> into the cavity <strong>and</strong> to exhaust<br />

the fluid <strong>and</strong> create the vortex. The<br />

68 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


effective length of the cavity can be increased<br />

by drawing the syringe plunger<br />

back. This draws the latex membrane<br />

back into the cylindrical well. Dye<br />

was introduced into the chamber via<br />

a1.6-mmdiameterholedrilledinto<br />

the acrylic plate, radial to the cavity. A<br />

steel tube was inserted into the hole,<br />

<strong>and</strong> was connected via medical tubing<br />

(Scientific Commodities, Inc., Lake<br />

Havasu City, AZ) to a syringe filled<br />

with food-grade dye. The vortex generator<br />

was mounted to an aluminum<br />

arm (80/20 Inc., Columbia City, IN)<br />

so that it could be positioned within<br />

the flow tank.<br />

The force, impulse, <strong>and</strong> linear velocity<br />

of 12 vortices were characterized<br />

to better underst<strong>and</strong> the properties<br />

of the vortex <strong>and</strong> how best to actuate<br />

the plunger. The force generated by<br />

the impact of the vortex was measured<br />

(Figure 5b) by shooting the vortex<br />

against a plate that was connected to<br />

a2.5g force transducer (LSB200, JR<br />

S-Beam Load Cell, Irvine, CA). The<br />

plate was located 100 mm from the orifice<br />

of the vortex generator. The vortex<br />

was imaged using the high-speed<br />

camera as it travelled within the 2-mm<br />

thick light sheet. Mean values for vortices<br />

created using a 5-mm diameter<br />

cavity were: 13 mN force (0.6 mN SE),<br />

0.13 mNs impulse (0.002 mNs SE),<br />

<strong>and</strong> 0.99 m/s velocity (0.01 m/s SE).<br />

A 13-mm diameter cavity produced<br />

a more powerful but slower vortex:<br />

67 mN force (2.3 mN SE), 1.0 mNs<br />

impulse (0.02 mNs SE), <strong>and</strong> 0.85 m/s<br />

velocity (0.01 m/s SE). These values<br />

compare well with estimates we have<br />

made for the peak force <strong>and</strong> impulse<br />

created by a sunfish pectoral fins. At<br />

a swimming speed of 0.5 body length<br />

per second, average fin forces are less<br />

than approximately 10 mN <strong>and</strong> the<br />

impulse over the fin beat is less than<br />

2.5 mNs.<br />

FIGURE 5<br />

Evaluation of vortex ring’s velocity using PIV (A). Force from vortex ring during impact with rigid<br />

plate attached to force transducer (B).<br />

Response to<br />

Vortex Perturbations<br />

Perturbation experiments that<br />

involved hitting the swimming fish<br />

with a vortex ring showed that the<br />

fish did not alter the pectoral fin beat<br />

during the time course of a single fin<br />

stroke but did change the amplitude<br />

<strong>and</strong> timing of the pectoral fin beats<br />

subsequent to a vortex impact that<br />

perturbed the fish’s position.<br />

Response to Vortex Perturbations<br />

Applied to the Body<br />

Vortex perturbations that impacted<br />

the side of the fish displaced the fish<br />

FIGURE 6<br />

laterally by several millimeters (Figure<br />

6), which is significant relative to<br />

the thickness of the fish’s body (maximum<br />

of approximately 25 mm). The<br />

lateral displacement occurred whether<br />

the fish had been drifting toward or<br />

away from the vortex generator prior<br />

to the disturbance <strong>and</strong> was not accompanied<br />

by any obvious change to<br />

the roll or yaw of the fish. An active<br />

response of the fish to the vortex perturbation<br />

was evident in the fishes’<br />

motion after a short delay. The soonest<br />

the active response occurred was<br />

0.05 s, while the longest delay before<br />

a response was evident was 0.20 s. In<br />

the majority of trials, the fishes actively<br />

Distance from orifice plate of the left pectoral fin (purple), the right pectoral fin (blue), <strong>and</strong> the<br />

fish at a point between the pelvic fins (green). The distance between the fish <strong>and</strong> the orifice plate<br />

is amplified relative to the fins <strong>and</strong> is measured at the scale on the right. In this trial, the fish was<br />

moving towards the vortex generator <strong>and</strong> was hit by the vortex at about t = 1.49 (red). The active<br />

response of the fish occurred by t = 1.50 (gray). (Color versions of figures available online at:<br />

http://www.ingentaconnect.com/content/mts/mtsj/2011/00000045/00000004.)<br />

July/August 2011 Volume 45 Number 4 69


moved away from the vortex generator<br />

after being hit by the vortex (Figure 7).<br />

The movement was not particularly<br />

quick, but was always faster than the<br />

fish’s lateral velocity before the perturbation<br />

had occurred. In some cases (e.g.,<br />

Figure 6), the fish actively moved toward<br />

the vortex generator after being<br />

pushed away from the vortex generator<br />

by the impulse. This occurred only<br />

when the fish had been drifting towards<br />

the vortex generator before the perturbation.<br />

In some trials, the fish was startled<br />

by the vortex <strong>and</strong> swam out of the test<br />

area. The startled motions were not<br />

analyzed quantitatively.<br />

The motions of the pectoral fins<br />

during the fin beat subsequent to the<br />

perturbation were significantly differentfromthemotionsofthepectoral<br />

fins prior to the perturbation. However,<br />

the pectoral fins did not seem to<br />

react quickly to the stimulus. In fact,<br />

the initial movement of the fish’s body<br />

in response to the vortex generally<br />

occurred between pectoral fin beats,<br />

while the pectoral fins were against<br />

the fish body (Figures 6 <strong>and</strong> 7).<br />

Thus, the active motion of the fish<br />

was initiated by other fins, which reacted<br />

within as little as 0.05 s. Active<br />

FIGURE 7<br />

Distance from orifice plate of the left pectoral fin (purple), the right pectoral fin (blue), <strong>and</strong> the<br />

fish at a point between the pelvic fins (green). The distance between the fish <strong>and</strong> the orifice plate<br />

is amplified relative to the fins <strong>and</strong> is measured at the scale on the right. In this trial, the fish was<br />

moving away from the vortex generator <strong>and</strong> was hit by the vortex at about t = 1.25 (red). The<br />

active response of the fish occurred by t = 1.35 (gray). The fish continued to move away from<br />

the vortex generator until approximate t = 1.8 s.<br />

movement of the pectoral fins did<br />

not usually resume until 0.10-0.20 s<br />

after the vortex. The frequency of the<br />

pectoral fin beats did not change<br />

consistently after the perturbation. In<br />

three of the eight trials, the frequency<br />

of the pectoral fin beat increased from,<br />

on average, 1.37 Hz (SD = 0.15) to<br />

1.83 Hz (SD = 0.28). In the other<br />

five trials, the frequency of the fin beat<br />

decreased from, on average, 1.69 Hz<br />

(SD = 0.26) to 1.34 Hz (SD = 0.21).<br />

The amplitude of the pectoral fin motions<br />

also changed. This altered the<br />

force balance between the two pectoral<br />

fins <strong>and</strong> contributed to the movement<br />

of the fish body. In the beat after the<br />

vortex stimulus, the amplitude of the<br />

right pectoral fin (opposite the side of<br />

the impact) was consistently smaller<br />

than before the vortex. Its motion decreased<br />

in all eight trials, on average by<br />

18.9% (SD = 10.9%). The amplitude<br />

of the left pectoral fin alsochanged,<br />

but the changes were less consistent.<br />

In four trials, the amplitude decreased<br />

by, on average, 37.4% (SD = 22.8),<br />

while in the other four trials, the<br />

amplitude increased by, on average,<br />

8.9% (SD 7.0%). By the second fin<br />

beat after the perturbation, the motions<br />

of the left <strong>and</strong> right pectoral<br />

fins were much more similar to the<br />

motions before the fin beat, <strong>and</strong> were<br />

similar to each other.<br />

Response to Vortex Perturbations<br />

Applied to the Fin<br />

The pectoral fins were deformed<br />

significantly when struck by the vortex<br />

during the fin beat (Figures 8 <strong>and</strong> 9).<br />

The vortex made contact with the left<br />

pectoral fin near the end of the fin’s<br />

outstroke. The vortex bent the tips<br />

of the fin rays <strong>and</strong> progressively bent<br />

larger portions of the fin as the vortex<br />

travelled towards the fish body. The fin<br />

seemed to bend <strong>and</strong> fold as if it were<br />

made from thin paper <strong>and</strong> exhibited<br />

deformations from the tip to the<br />

base. The maximum measured curvature<br />

of the fin (alongfin ray6)increased<br />

from 0.054 mm −1 near the tip<br />

<strong>and</strong> 0.024 mm −1 near the base during<br />

unperturbed swimming to 0.113 mm −1<br />

near the tip <strong>and</strong> 0.029 mm −1 near<br />

the base when in contact with the vortex.<br />

The vortex remained in contact<br />

with the fin while it travelled towards<br />

the fish body. This resulted in the<br />

pectoral fin being pushed back to the<br />

body faster than during an unperturbed<br />

instroke. Times ranged from one third<br />

to one half of the duration of a normal<br />

instroke <strong>and</strong> were dependent on many<br />

FIGURE 8<br />

Pectoral fin perturbed by vortex during swimming.<br />

The mean fin ray curvature after impact<br />

was 14.4 mm −1 (0.05 mm SE). Reflective<br />

particles are used so that the fluid movement<br />

is visible.<br />

70 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 9<br />

Ventral view of the fish as the left pectoral fin is hit by a vortex (no dye). The left pectoral fin<br />

is hit by the vortex (1). The fin is deformed (2, 3, <strong>and</strong> 4) <strong>and</strong> is pushed to the body by the vortex.<br />

(5 <strong>and</strong> 6) The right pectoral fin continues to beat normally. The body is not deflected by the vortex.<br />

variables, including the speed of the<br />

vortex, how well contact was made<br />

with the fin, <strong>and</strong> the time of impact<br />

within the fin beat.<br />

Despite the severity with which the<br />

vortex changed the shape <strong>and</strong> trajectory<br />

of the perturbed fin, the fish did<br />

not appear to react to the perturbation<br />

or to change its behavior subsequent to<br />

the perturbation. During the perturbation,<br />

the observed motions of the<br />

unperturbed fin <strong>and</strong>ofthefish body<br />

were not visibly different from motions<br />

prior to the perturbation. Subsequent<br />

to the perturbation, the perturbed fin<br />

remained against the fish body until<br />

the unperturbed fin completed its<br />

instroke. Both fins then resumed<br />

what appeared to be a normal fin<br />

beat. Small differences in the pectoral<br />

fin beat <strong>and</strong> the use of other fins likely<br />

occurred to accommodate for differences<br />

in propulsive forces produced<br />

during the perturbation, but these<br />

changes were not visible. Nor were<br />

there changes in the motion of the<br />

fish body, which was not observed to<br />

move laterally or to rotate in yaw.<br />

Discussion<br />

Theobjectiveoftheexperiments<br />

was to determine how sunfish respond<br />

to perturbations applied to the body<br />

<strong>and</strong> fins during steady swimming.<br />

These experiments provided a contextual<br />

underst<strong>and</strong>ing of sensory based<br />

modulation of pectoral fin function.<br />

The experiments produced a mix of<br />

expected <strong>and</strong> surprising results.<br />

As expected, the fishes did alter the<br />

amplitude <strong>and</strong> timing of the pectoral<br />

fin beats subsequent to a perturbation<br />

that disturbed the lateral position of<br />

the fish body. However, the pectoral<br />

fins did not respond quickly to the<br />

disturbance, but remained against the<br />

fish body for durations that were only<br />

slightly different from the pauses<br />

between fin beats prior to the disturbance.<br />

Active movement of the fish’s<br />

body after the disturbance occurred<br />

with a latency of as little as 0.05 s,<br />

which is similar to the 0.08 s latency<br />

measured by Webb (2004) in response<br />

to roll disturbances. The movement of<br />

the fish body is believed to have been<br />

caused by fins other than the pectoral<br />

fins, since the pectoral fins remained<br />

against the body for 0.10–0.20 s after<br />

the perturbation. When the pectoral<br />

fins were moved, the amplitudes of<br />

the fins seemed to have been adjusted<br />

to help equilibrate the movement of<br />

the fish. By the second fin beatafter<br />

the disturbance, the motions of the<br />

two pectoral fins were synchronized<br />

<strong>and</strong> had amplitudes similar to those<br />

before the disturbance.<br />

The delay in the response of the<br />

pectoral fins to the vortex <strong>and</strong> lateral<br />

disturbance is different from the response<br />

of the fins during experiments<br />

where an obstacle was placed in front<br />

of the swimming fish (Gottlieb et al.,<br />

2010). In those experiments, sunfish<br />

altered the motions of the left<br />

<strong>and</strong> right pectoral fins during the outstroke<br />

of a steady swimming beat.<br />

The changes were not subtle, <strong>and</strong> the<br />

fish did not seem to wait for the next<br />

cycle as in the present studies. The pectoral<br />

fin onthesideoftheobstacle<br />

stiffened <strong>and</strong> the fin rays were moved<br />

through trajectories that were very different<br />

from steady swimming. The fin<br />

on the side opposite to the obstacle<br />

nearly stopped <strong>and</strong> served to stabilize<br />

the motion of the fish. The difference<br />

in the pectoral fins’ response to the<br />

obstacle <strong>and</strong> to the vortex <strong>and</strong> lateral<br />

displacement may be related to the<br />

fish’s perception of the stimuli. The<br />

obstacle may have been more threatening<br />

than the vortex, which the fish may<br />

have interpreted as a common fluidic<br />

event. The fish therefore disrupted<br />

the steady swimming gait in order to<br />

produce large lateral forces that turned<br />

the fish away from an unknown obstacle<br />

that may have posed a threat.<br />

In contrast, the disturbance in motions<br />

caused by a fluidic stimulus could be<br />

accommodated simply by adjusting<br />

motions of the fins within their normal<br />

gaits. This would allow the central<br />

pattern generator that drives the motions<br />

of pectoral fins (Westneat et al.,<br />

2004) to continue to produce similar<br />

output characteristic rather than having<br />

to switch between gaits.<br />

Most surprising was the lack of reaction<br />

to the vortex when the vortex<br />

deformed the pectoral fin attheend<br />

July/August 2011 Volume 45 Number 4 71


of the fin’s outstroke <strong>and</strong> throughout<br />

the instroke. Nerves <strong>and</strong> free nerve<br />

endings exist throughout the fin rays<br />

<strong>and</strong> the fin webbing (experimental<br />

findings, M. Hale, University of<br />

Chicago), <strong>and</strong> so it was expected that<br />

at least one of the phenomena that<br />

the vortex created—pressure, impact,<br />

bending—would have elicited a sensory<br />

mediated response. The vortex<br />

was in contact with the fin for over<br />

100ms,<strong>and</strong>sothedurationofthe<br />

stimulus was certainly sufficient for a<br />

sensory-mediated response to occur.<br />

It was also surprising that neither the<br />

motions of the body, nor subsequent<br />

beats of the pectoral fins, were clearly<br />

different from those before the vortex<br />

perturbation. It is highly likely that the<br />

left pectoral fin, while being deformed,<br />

produced forces that were different<br />

from normal. During a normal steady<br />

swimming gait, each pectoral fin will<br />

produce lateral forces that are similar<br />

in magnitude to thrust <strong>and</strong> lift. Since<br />

the fins typically beat synchronously,<br />

the lateral forces from the left <strong>and</strong><br />

right fins balance <strong>and</strong> cancel. This<br />

would not have been the case when<br />

the left pectoral fin was deformed,<br />

<strong>and</strong> the unbalanced forces should<br />

have accelerated the fish body laterally<br />

<strong>and</strong>/or in roll <strong>and</strong> yaw. The lack of obvious<br />

lateral motion <strong>and</strong> adjustment to<br />

the pectoral fin beat may be due simply<br />

to the fish being insensitive to lateral<br />

forces. To move the fish laterally,<br />

forces must accelerate the mass of the<br />

fish <strong>and</strong> also overcome drag forces<br />

<strong>and</strong> the load from the mass of water<br />

against which the side of the fish<br />

pushes. Thus, the loss of lateral force<br />

during a single fin beat can be easily<br />

tolerated because it is difficult for the<br />

fish to move sideways. So although<br />

studies of biorobotic models of the<br />

pectoral fins have shown that the<br />

fin’s kinematics <strong>and</strong> mechanical properties<br />

must be controlled very carefully<br />

to produce forces like the fish<br />

(Tangorra et al., 2007, 2010), the<br />

mechanics of the fish body do not<br />

necessarily require the careful control<br />

of forces at all times in all directions.<br />

Conclusions<br />

Perturbation experiments which<br />

involved hitting the swimming sunfish<br />

with a vortex ring showed that the<br />

fish did not alter the pectoral fin beat<br />

during the time course of a single fin<br />

stroke but did change the amplitude<br />

<strong>and</strong> timing of its motions in beats subsequent<br />

to an impact that disturbed<br />

the fish’s position. Vortices that struck<br />

the pectoral fin during the fin’s outstroke<br />

deformed the fin extensively,<br />

but the perturbations did not cause<br />

the stroke of the unaffected pectoral<br />

fin to change, nor did the perturbation<br />

cause changes in the motions of the<br />

fish body or in the subsequent strokes<br />

of either pectoral fin.<br />

These outcomes suggest that the<br />

kinematics of the pectoral fins is modulated<br />

by sensory information only<br />

when a perturbation results in a disturbance<br />

to the fish body, which is the system<br />

that the fins are working to control.<br />

The pectoral fins did not react quickly<br />

when the vortex displaced the fish’s<br />

body but modulated their motions to<br />

help stabilize the displaced fish after<br />

other fins had already been engaged.<br />

The pectoral fins also did not react reflexively<br />

to vortex perturbations that<br />

deformed the fins’ webbing <strong>and</strong> fin<br />

rays. The fin did not appear to move<br />

away from the vortex or to resist the deformation<br />

by stiffening. The compliant<br />

fin allowed itself to bend <strong>and</strong> perhaps to<br />

shed the load from the vortex, <strong>and</strong> then<br />

altered its motions during the course of<br />

the subsequent fin beat.<br />

Theresultsillustrateabenefit of<br />

compliant mechanisms within a highly<br />

controllable system. The fins of rayfinned<br />

fish have the ability to control<br />

forces precisely by altering the kinematics<br />

<strong>and</strong> mechanical properties of<br />

individual fin rays. Small changes in<br />

either the trajectories or stiffness of<br />

fin rays can significantly alter the<br />

force that is transferred to the fish<br />

(Tangorra et al., 2010). However, it<br />

is not always necessary to regulate the<br />

fins precisely. By maintaining its flexibility<br />

<strong>and</strong> allowing itself to be deformed,<br />

the fin was able to be hit by<br />

the vortex—which had sufficient<br />

forces to displace the fish—without<br />

transferring the full impact of the perturbation<br />

to the fish body. Therefore,<br />

the fin’s passive mechanics made it unnecessary<br />

for the fins to be modulated<br />

in order to restore the fish to equilibrium.<br />

However, when the fish does want to<br />

move the quickly—as in a maneuver<br />

away from the obstacle—the pectoral<br />

fin can be stiffened, the gait changed,<br />

<strong>and</strong> large lateral forces from the fin<br />

can be transferred to the fish body.<br />

Acknowledgments<br />

This work was supported by the Office<br />

of Naval Research grant N00014-<br />

0910352 on fin neuromechanics monitored<br />

by Dr. Thomas McKenna <strong>and</strong><br />

by the National Science Foundation<br />

EFRI 0938043. We are very grateful<br />

to the members of the Lauder <strong>and</strong> the<br />

Tangorra laboratories for many helpful<br />

discussions about fish fins <strong>and</strong> robotic<br />

fins <strong>and</strong> for assistance in executing<br />

experiments <strong>and</strong> analyzing results.<br />

Lead Author:<br />

James Tangorra<br />

3141 Chestnut St.,<br />

R<strong>and</strong>ell 115<br />

Drexel University,<br />

Philadelphia, PA 19104<br />

Email: tangorra@coe.drexel.edu<br />

72 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


References<br />

Akhtar, I., Mittal, R., Lauder, G., & Drucker,<br />

E. 2007. Hydrodynamics of a biologically<br />

inspired t<strong>and</strong>em flapping foil configuration.<br />

Theoret Comput Fluid Dynamics. 21(3):<br />

155-70. doi: 10.1007/s00162-007-0045-2.<br />

Allen, J., & Auvity, B. 2002. Interaction of a<br />

vortex ring with a piston vortex. J Fluid Mech.<br />

465(1):353-78.<br />

Dong, H., Bozkurttas, M., Mittal, R.,<br />

Madden, P., & Lauder, G. 2010. Computational<br />

modelling <strong>and</strong> analysis of the hydrodynamics<br />

of a highly deformable fish pectoral<br />

fin. J Fluid Mech. 645:345-73. doi: 10.1017/<br />

S0022112009992941.<br />

Fish, F.E., Nusbaum, M.K., Beneski, J., &<br />

Ketten, D. 2006. Passive cambering <strong>and</strong><br />

flexible propulsors: cetacean flukes. Bioinspir<br />

Biomim. 1(4):S42-48. doi: 10.1088/<br />

1748-3182/1/4/S06.<br />

Gharib, M., Rambod, E., & Sharriff, K.<br />

1998. A universal time scale for vortex ring<br />

formation. J Fluid Mech. 360(1):121-40.<br />

doi: 10.1017/S0022112097008410.<br />

Gottlieb, J., Tangorra, J., Esposito, C., &<br />

Lauder, G. 2010. A biologically derived<br />

pectoral fin for yaw turn manoeuvres.<br />

Appl Bionics Biomech. 7(1):41-55.<br />

doi: 10.1080/11762320903537782.<br />

Lauder, G.V., & Madden, P.G. 2007. Fish<br />

locomotion: kinematics <strong>and</strong> hydrodynamics<br />

of flexible foil-like fins. Exp Fluids. 43(5):<br />

641-53. doi: 10.1007/s00348-007-0357-4.<br />

Lauder, G.V., Madden, P.G., Mittal, R.,<br />

Dong, H., & Bozkurttas, M. 2006. Locomotion<br />

with flexible propulsors: I. Experimental<br />

analysis of pectoral fin swimming in sunfish.<br />

Bioinspir Biomim. 1:S25. doi: 10.1088/<br />

1748-3182/1/4/S04.<br />

Liu, P., & Bose, N. 1997. Propulsive performance<br />

from oscillating propulsors with<br />

spanwise flexibility. P Roy Soc Lond A Mat.<br />

453(1963):1763.<br />

McHenry, M. 1995. Mechanical control of<br />

swimming speed: stiffness <strong>and</strong> axial wave<br />

form in undulating fish models. J Exp Biol.<br />

198(11):2293-305.<br />

Mittal, R., Dong, H., Bozkurttas, M., Lauder,<br />

G., & Madden, P.G. 2006. Locomotion<br />

with flexible propulsors: II. Computational<br />

modeling of pectoral fin swimming in<br />

sunfish. Bioinspir Biomim. 1(4):S35-41.<br />

doi: 10.1088/1748-3182/1/4/S05.<br />

Mohseni, K. 2006. Pulsatile vortex generators<br />

for low-speed maneuvering of small underwater<br />

vehicles. Ocean Eng. 33(16):2209-23.<br />

doi: 10.1016/j.oceaneng.2005.10.022.<br />

Nicholson, J. W., & Healey, A.J. 2008. The<br />

present state of autonomous underwater<br />

vehicle (AUV) applications <strong>and</strong> technologies.<br />

Mar Technol Soc. 42(1):8.<br />

Office of the Secretary of Defense, USA.2009.<br />

Unmanned Systems Roadmap (2009-2034).<br />

Phelan, C.T., Tangorra, J.L., Lauder, G.V.,<br />

& Hale, M.E. 2010. A biorobotic model of<br />

the sunfish pectoral fin for investigations of fin<br />

sensorimotor control. Bioinspir Biomim. 5.<br />

doi: 10.1088/1748-3182/5/3/035003.<br />

Prempraneerach, P., Hover, F.S., &<br />

Triantafyllou, M.S. 2003. The effect of<br />

chordwise flexibility on the thrust <strong>and</strong><br />

efficiency of a flapping foil. Autonomous<br />

Undersea Systems Institute. In: 13th Int.<br />

Symp. Unmanned Untethered Submersible<br />

Techn. Durham, NH.<br />

Shoele, K., & Zhu, Q. 2009. Fluid–structure<br />

interactions of skeleton-reinforced fins: performance<br />

analysis of a paired fin in lift-based<br />

propulsion. J Exp Biol. 212(16):2679.<br />

doi: 10.1242/jeb.030023.<br />

Shusser, M., Gharib, M., Rosenfeld, M., &<br />

Mohseni, K. 2002. On the effect of pipe<br />

boundary layer growth on the formation of<br />

a laminar vortex ring generated by a piston/<br />

cylinder arrangement. Theoret Comput Fluid<br />

Dynamics. 15(5):303-16. doi: 10.1007/<br />

s001620100051.<br />

Tangorra, J., Lauder, G.V., Hunter, I.W.,<br />

Mittal, R., Madden, P.G., & Bozkurttas, M.<br />

2010. The effect of fin ray flexural rigidity on<br />

the propulsive forces generated by a biorbotic<br />

fish pectoral fin. J Exp Biol. 213:4043-54.<br />

doi: 10.1242/jeb.048017.<br />

Tangorra, J., Phelan, C.T., Esposito, C.J., &<br />

Lauder, G.V. 2011. Use of biorobotic models<br />

of highly deformable fins for studying the<br />

mechanics <strong>and</strong> control of fin forces in fishes.<br />

Integr Comp Biol. in press. doi: 10.1093/<br />

icb/icr036.<br />

Tangorra, J.L., Davidson, S.N., Hunter, I.W.,<br />

Madden, P.G., Lauder, G.V., Dong, H., …<br />

Mittal, R. 2007. The development of a biologically<br />

inspired propulsor for unmanned<br />

underwater vehicles. IEEE J Oceanic Eng.<br />

32(3):533-50. doi: 10.1109/JOE.2007.903362.<br />

Triantafyllou, M., Techet, A., Zhu, Q.,<br />

Beal, D., Hover, F., & Yue, D. 2002.<br />

Vorticity control in fish-like propulsion <strong>and</strong><br />

maneuvering. Integr Comp Biol. 42(5):1026.<br />

doi: 10.1093/icb/42.5.1026.<br />

Triantafyllou, M.S., Hover, F.S., Techet, A.,<br />

& Yue, D. 2005. Review of hydrodynamic<br />

scaling laws in aquatic locomotion <strong>and</strong> fishlike<br />

swimming. Appl Mech Rev. 58(4):226-37.<br />

doi: 10.1115/1.1943433.<br />

Webb, P. 2004. Response latencies to postural<br />

disturbances in three species of teleostean<br />

fishes. J Exp Biol. 207(6):955-61. doi: 10.1242/<br />

jeb.00854.<br />

Westneat, M., Thorsen, D., Walker, J., &<br />

Hale, M. 2004. Structure, function, <strong>and</strong><br />

neural control of pectoral fins in fishes. IEEE J<br />

Oceanic Eng. 29(3):674-83. doi: 10.1109/<br />

JOE.2004.833207.<br />

Zhu, Q., & Shoele, K. 2008. Propulsion<br />

performance of a skeleton-strengthened fin.<br />

J Exp Biol. 211(13):2087. doi: 10.1242/<br />

jeb.016279.<br />

Zhu, Q., Wolfgang, M.J., Yue, D.K.P., &<br />

Triantafyllou, M.S. 2002. Three-dimensional<br />

flow structures <strong>and</strong> vorticity control in fishlike<br />

swimming. J Fluid Mech. 468:1-28.<br />

doi: 10.1017/S002211200200143X.<br />

July/August 2011 Volume 45 Number 4 73


PAPER<br />

Bioinspired Design Process for an Underwater<br />

Flying <strong>and</strong> Hovering Vehicle<br />

AUTHORS<br />

Jason D. Geder<br />

John S. Palmisano<br />

Ravi Ramamurti<br />

Marius Pruessner<br />

Banahalli Ratna<br />

Naval Research Laboratory<br />

William C. S<strong>and</strong>berg<br />

Science Applications<br />

International Corporation<br />

Background<br />

Biologists <strong>and</strong> zoologists have<br />

been studying fish swimming for<br />

many decades, <strong>and</strong> several comprehensive<br />

texts <strong>and</strong> papers exist (Breder,<br />

1926; Lindsey, 1978; Alex<strong>and</strong>er,<br />

1983; Azuma, 1992; Blake, 1983;<br />

Webb, 1975, 1984; Videler, 1993;<br />

Sfakiotakis et al., 1999). The experimental<br />

studies of fish swimming biodynamics<br />

have become increasingly<br />

quantitative as measurement technology<br />

has improved. The use of highspeed<br />

photography shed light upon<br />

the details of fin deformation (Gibb<br />

et al., 1994; Walker & Westneat,<br />

1997). Laser light scattering techniques<br />

enabled observations of not<br />

only the fish body <strong>and</strong> fin dynamics<br />

but also the velocity field about<br />

the fish <strong>and</strong> in the wake (Drucker &<br />

Lauder, 1999, 2002; Gharib et al.,<br />

2002; Bartol et al., 2003). Vehicle designers<br />

are able to draw upon such rich<br />

data sets as they embark upon designs.<br />

But where does one begin<br />

A bioinspired vehicle design should<br />

begin, according to Webb (2004), by<br />

specifying the performance goals of<br />

ABSTRACT<br />

We review here the results obtained during the past several years in a series of<br />

computational <strong>and</strong> experimental investigations aimed at underst<strong>and</strong>ing the origin of<br />

high-force production in the flapping wings of insects <strong>and</strong> the flapping <strong>and</strong> deforming<br />

fins of fish <strong>and</strong> the incorporation of that information into bioinspired vehicle<br />

designs. We summarize the results obtained on pectoral fin force production, flapping<br />

<strong>and</strong> deforming fin design, <strong>and</strong> the emulation of fish pectoral fin swimming in<br />

unmanned vehicles. In particular, we discuss the main results from the computational<br />

investigations of pectoral fin force production for a particular coral reef fish,<br />

the bird wrasse (Gomphosus varius), whose impressive underwater flight <strong>and</strong> hovering<br />

performance matches our vehicle mission requirements. We describe the<br />

tradeoffs made between performance <strong>and</strong> produceability during the bio-inspired<br />

design of an actively controlled curvature pectoral fin <strong>and</strong> the incorporation of it<br />

into two underwater flight vehicles: a two-fin swimming version <strong>and</strong> four-fin swimming<br />

version. We describe the unique computational approach taken throughout<br />

the fin <strong>and</strong> vehicle design process for relating fin deformation time-histories to<br />

specified desired vehicle dynamic behaviors. We describe the development of the<br />

vehicle controller, including hardware implementation, using actuation of the multiple<br />

deforming flapping fins as the only means of propulsion <strong>and</strong> control. Finally,<br />

we review the comparisons made to date between four-fin vehicle experimental<br />

trajectory measurements <strong>and</strong> controller simulation predictions <strong>and</strong> discuss the<br />

incorporation of those comparisons into the controller design.<br />

Keywords: bio-inspired robotics, pectoral fin, unmanned systems, computational<br />

fluid dynamics<br />

the desired mission first <strong>and</strong> then examining<br />

those living creatures whose<br />

performance is relevant. These creatures<br />

have evolved to meet all their<br />

needs, <strong>and</strong> the maneuvering enabled<br />

by these needs may intersect with<br />

the performance requirements driving<br />

a vehicle design. However, since the<br />

living creatures selected for study are<br />

most likely not optimized for the mobility<br />

characteristics that are driving<br />

the design, one should not copy nature<br />

but instead be guided by it.<br />

This point of caution to designers<br />

has been made many times by biologists<br />

(Combes & Daniel, 2001;<br />

Wainwright et al., 2002; Collar et al.,<br />

2008).<br />

The mission selected for the Naval<br />

Research Laboratory (NRL) swimming<br />

vehicle, which we describe<br />

below, requires precise low-speed maneuvering<br />

<strong>and</strong> excellent hovering in a<br />

complex near-shore environment in<br />

addition to excellent position-keeping<br />

in tidal currents. Cost <strong>and</strong> mechanical<br />

simplicity constraints dem<strong>and</strong>ed a<br />

rigid hull. This eliminated undulatory<br />

swimming fish as a primary means of<br />

design inspiration, <strong>and</strong> instead we<br />

were lead to consider fin-based swimming<br />

creatures. Kato <strong>and</strong> Furushima<br />

74 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


(1996) had already proceeded down<br />

the path of paired fin swimming, drawing<br />

inspiration from the black bass to<br />

design rigid pectoral fins <strong>and</strong> incorporate<br />

them into a test-bed vehicle.<br />

He subsequently pursued low-speed<br />

maneuvering using rigid pectoral fins<br />

(Kato, 2000) <strong>and</strong> then went on to develop<br />

a passive flexible fin <strong>and</strong> an active<br />

pneumatic actuator pectoral fin (Kato<br />

et al., 2008). Barrett <strong>and</strong> Triantafyllo<br />

(1995) on the other h<strong>and</strong> had taken<br />

their inspiration from the undulating<br />

body <strong>and</strong> oscillating caudal fin ofthe<br />

tuna to design <strong>and</strong> build their flexible<br />

“Robotuna.” We looked for a fish<br />

which possessed the dynamic performance<br />

characteristics we needed <strong>and</strong><br />

for which experimental measurements<br />

of swimming dynamics, including fin<br />

kinematics, already existed. Experimental<br />

observations of pectoral fin<br />

muscle activity, kinematics, <strong>and</strong> dynamics<br />

in coral reef wrasses (Westneat<br />

& Walker, 1997; Walker & Westneat,<br />

1997) have shown that the body is<br />

essentially held rigid during straightline<br />

motion, thus satisfying our rigid<br />

hull constraint. Very rapid (∼10 body<br />

lengths per second) translational<br />

motions were observed (Walker &<br />

Westneat, 2000) to give comparable<br />

swimming performance to that seen<br />

in body-caudal fin swimmers of comparable<br />

size even though there was no<br />

contribution from body undulation<br />

or caudal fin oscillation in the wrasses.<br />

This solution offered the potential of<br />

fast forward swimming (or positionkeeping<br />

in strong currents) while<br />

avoiding the complexity of a flexible<br />

hull. In addition, Walker <strong>and</strong><br />

Westneat (1997) had observed high<br />

maneuverability by these swimmers<br />

in their complex reef habitats. These<br />

habitats are reasonable representations<br />

of our vehicle’s projected operating<br />

environment, hence we saw the possibility<br />

of obtaining both high forward<br />

speed <strong>and</strong> excellent low-speed maneuvering<br />

<strong>and</strong> hovering performance as<br />

well.<br />

The fish we selected to inspire our<br />

design was one of the coral reef pectoral<br />

fin swimmers, the bird wrasse<br />

(Gomphosus varius), shown in Figure 1.<br />

FIGURE 1<br />

Bird wrasse (Gomphosus varius).<br />

Controlled Deformation<br />

Fin <strong>Technology</strong><br />

Development<br />

Living creatures, such as insects,<br />

birds, <strong>and</strong> pectoral fin swimmers,<br />

generate lift <strong>and</strong> thrust by executing<br />

large-amplitude wing/fin flapping,<br />

often with substantial shape deformation<br />

from root to tip <strong>and</strong> leading edge<br />

to trailing edge. The flow for these<br />

motions is three-dimensional <strong>and</strong> unsteady,<br />

<strong>and</strong> conventional steady-state<br />

aerodynamics is unable to correctly<br />

compute the corresponding time history<br />

of flapping-force generation.<br />

Comprehensive reports on the research<br />

carried out to study the fluid dynamics<br />

of flapping fins <strong>and</strong> wings<br />

(Rozhdestvensky & Ryzhov, 2003)<br />

<strong>and</strong> of biomimetic fins for underwater<br />

vehicles (Triantafyllo et al., 2004)<br />

exist, in addition to an extensive number<br />

of studies, too great to list here, on<br />

unsteady lift production by flapping<br />

insect wings. Three-dimensional unsteady<br />

computations are necessary to<br />

correctly predict the lift <strong>and</strong> thrust variation<br />

throughout the flapping stroke<br />

cycle. Such computations, for creaturesorvehicleswithmoving<strong>and</strong><br />

deforming surfaces, provide the timevarying<br />

pressure distribution on all<br />

surfaces, which in turn can provide insights<br />

into how the flapping forces <strong>and</strong><br />

maneuvering moments are being generated.<br />

This information can be coupled<br />

with computational visualization<br />

of the time-varying flow about the<br />

fish to analyze the origin of body <strong>and</strong><br />

fin vorticity, its growth, <strong>and</strong> eventual<br />

shedding into the wake. This is the<br />

approach we developed for tuna caudal<br />

fin force production analysis<br />

(Ramamurti et al., 1996, 1999), subsequently<br />

validated against wrasse experimental<br />

data (Ramamurti et al.,<br />

2002), <strong>and</strong> which we also followed<br />

throughout our fin <strong>and</strong> vehicle development<br />

efforts described below.<br />

There are 13 multiply bifurcating<br />

fin rays in the bird wrasse pectoral fin<br />

shown in Figure 2, each contributing<br />

to the fin curvature time-variation<br />

throughout the stroke cycle. For ease<br />

of design, manufacture, actuation,<br />

<strong>and</strong> control, it is ideal to have the fewest<br />

possible number of rays (which we<br />

refer to as ribs) <strong>and</strong> have each of them<br />

FIGURE 2<br />

Bird wrasse fin structure (from Walker &<br />

Westneat, 1997).<br />

July/August 2011 Volume 45 Number 4 75


e as simple in shape <strong>and</strong> structure<br />

as possible. But for more effective fin<br />

propulsion, it is ideal to maximize the<br />

number of ribs since more control<br />

points result in a smoother fit to<br />

desired fin curvature time-histories.<br />

Our computational investigations<br />

(Ramamurti et al., 2004) have shown<br />

that the loss of force magnitude over<br />

the stroke cycle is very small if we<br />

considerably reduce the number of<br />

ribs, as long as we maintain the ability<br />

to properly modify the surface curvature.<br />

The computations showed that<br />

when the fin was made rigid by specifying<br />

the motion with just the leading<br />

edge of the fin tip, the thrust produced<br />

during the upstroke was less than half<br />

of the peak thrust produced by the<br />

flexible fin computations. During the<br />

downstroke, the computations for<br />

the rigid <strong>and</strong> nearly rigid fin produced<br />

no positive thrust, while the<br />

partially <strong>and</strong> fully flexible cases produced<br />

substantial thrust. In the case<br />

of the rigid fin, there was also a substantial<br />

penalty in lift during the<br />

FIGURE 3<br />

Effect of fin flexibility on the time variation of thrust forces (from Ramamurti et al., 2004).<br />

(Color versions of figures available online at: http://www.ingentaconnect.com/content/mts/<br />

mtsj/2011/00000045/00000004.)<br />

FIGURE 4<br />

Representative rib cross-sectional geometry<br />

<strong>and</strong> bending analysis showing a 20° rib<br />

deflection.<br />

upstroke. An example from these<br />

computational investigations is shown<br />

below in Figure 3.<br />

Assessment of these findings led us<br />

to reduce the number of ribs from 13<br />

to 5. Five was selected since it enabled<br />

reduced fin complexity, thus substantially<br />

reducing fin size <strong>and</strong> weight,<br />

while maintaining the critically important<br />

flexural capability. Each of the five<br />

fin rays were individually designed<br />

<strong>and</strong> constructed from compliant ABS<br />

plastic material using a 3-D printer to<br />

achieve the desired tip deflection with<br />

an achievable linear actuation force applied<br />

to each individual rib at the root<br />

(Trease et al., 2003), as illustrated in<br />

Figure 4. The pushing <strong>and</strong> pulling of<br />

the ribs at the root of the fin is similar<br />

to how fish bend their ribs using muscle<br />

actuation.<br />

The root section of the fin was selected<br />

to be of rectangular cross section<br />

with rounded leading edges <strong>and</strong> a tapered<br />

trailing edge in order to accommodate<br />

the actuators at the base of the<br />

ribs as shown in Figure 5a. A translucent<br />

silicone rubber membrane skin,<br />

optimized in thickness to approximately<br />

0.5 mm via finite-element analysis<br />

(Palmisano et al., 2007), provided<br />

the continuous surface covering for<br />

the rays as shown in Figure 5b.<br />

After several parametric 3-D unsteady<br />

computational fluid dynamics<br />

(CFD) studies had been carried out<br />

(Ramamurti & S<strong>and</strong>berg, 2006), various<br />

fin parameters were chosen that<br />

optimized thrust performance, given<br />

the mechanical constraints of the fin.<br />

An example of these computations<br />

showing the variation of lift <strong>and</strong><br />

thrust as a function of fin flexibility,<br />

stroke amplitude, <strong>and</strong> fin strokebias<br />

is shown in Figure 6.<br />

The angle of attack of the root section<br />

of the fin was chosen to be 20°, the<br />

amplitude of the oscillation to be 114°,<br />

the flapping frequency to be 1 Hz, <strong>and</strong><br />

the rib spacing to be the minimum<br />

possible value of 1.2 cm dictated by<br />

the size of the actuators. These specific<br />

values, selected for maximizing fin<br />

forceproduction,areforasolitary<br />

flapping <strong>and</strong> deforming fin. Incorporation<br />

of the fin into a vehicle design<br />

presents challenges to retain fin performance<br />

while maintaining vehicle<br />

simplicity.<br />

Two-Fin Test-Bed Vehicle<br />

A simple vehicle was designed <strong>and</strong><br />

built to serve as a test-bed for evaluating<br />

the performance of all aspects of<br />

76 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 5<br />

Mechanical fin (a) CAD image without skin <strong>and</strong> (b) actual with skin.<br />

the controlled curvature fin technology,<br />

including its capability for vehicle<br />

propulsion, low-speed maneuvering,<br />

<strong>and</strong> hovering. It was, therefore, intentionally,<br />

a minimalist design that<br />

served to house the fins, actuators,<br />

<strong>and</strong> a battery. We described above<br />

how the parameters that govern the<br />

force production by the fin were chosen.<br />

However, due to mechanical<br />

constraints <strong>and</strong> a desire to easily<br />

manufacture a vehicle prototype, additional<br />

3-D unsteady CFD studies of<br />

the flapping fins incorporated in the<br />

FIGURE 6<br />

test-bed vehicle led us to modify the<br />

fin-alone values. The angle of attack<br />

of the root section of the fin was chosentobe0°,<strong>and</strong>theribspacingwas<br />

reduced to 0.8 cm. It is during construction<br />

tradeoffs of this type that<br />

one relies upon what has been learned<br />

from the studies of nature to meet operational<br />

performance goals while balancing<br />

that with the desire to create a<br />

vehicle that is producible at a reasonable<br />

cost. Fixing the fin root at a specific<br />

angle deviates from nature, but<br />

the construction simplicity <strong>and</strong> cost<br />

Mean fin generated (a) thrust <strong>and</strong> (b) lift as functions of non-dimensionalized stroke amplitude<br />

<strong>and</strong> fin flex, or curvature. The surfaces indicate a bias in the fin stroke angle of 0° (red), 20°<br />

(green), <strong>and</strong> 40° (blue).<br />

savings are so substantial that some<br />

performance penalty was accepted.<br />

The flapping stroke amplitude <strong>and</strong><br />

frequency, as well as the phasing of<br />

the individual fin tipdeflection time<br />

histories, were retained as controllable<br />

parameters, <strong>and</strong> we have performed<br />

CFD analyses of force production<br />

with this finconfiguration (Ramamurti<br />

et al., 2010).<br />

In keeping with the review nature<br />

of this paper, we are emphasizing the<br />

process carried out for our specific<br />

bio-inspired vehicle design. The details<br />

of the controlled curvature fin design,<br />

the linear actuator design <strong>and</strong> construction,<br />

the isolated fin construction<br />

<strong>and</strong> testing, the vehicle design, the vehicle<br />

construction, the experimental<br />

testing, <strong>and</strong> the validation of computations<br />

were previously reported<br />

(Palmisano et al., 2007, 2008; S<strong>and</strong>berg<br />

& Ramamurti, 2008). The fin technology<br />

test-bed demonstration vehicle<br />

incorporated two actively controlled deformation<br />

fins. The two-fin vehicle is<br />

shown in Figure 7.<br />

Four-Fin Test-Bed Vehicle<br />

The test results for the two-fin vehicle<br />

demonstrated that the controlled<br />

deformation fin force production was<br />

capable of meeting our vehicle propulsion<br />

(position-keeping in a current)<br />

requirements (Geder et al., 2008).<br />

However, by design, this test-bed fin<br />

technology demonstration vehicle<br />

had restricted options for sensor payload<br />

<strong>and</strong> did not have the fore-aft<br />

force production capability needed<br />

for heave-pitch control. Hence, the<br />

design of a larger 41-cm long four-fin<br />

vehicle was initiated (Figure 8). The<br />

fore-aft symmetry of the four-fin design<br />

enables hover <strong>and</strong> higher precision<br />

positioning capabilities by decoupling<br />

vehicle pitch <strong>and</strong> heave control.<br />

July/August 2011 Volume 45 Number 4 77


FIGURE 7<br />

Two-fin technology test-bed demonstration vehicle: (a) exterior view <strong>and</strong> (b) interior view.<br />

The current four-fin vehicle design shown here employs a water-tight cylinder<br />

for housing the power source <strong>and</strong> electronics with a flooded space in the nose <strong>and</strong><br />

tail for buoyancy trimming <strong>and</strong> supplemental sensors. Hardware control <strong>and</strong> all<br />

computations are performed by a 16-MHz ATmega2560 microcontroller. Computations<br />

(Ramamurti et al., 2010) <strong>and</strong> experimental tests carried out to date<br />

(Geder et al., 2011) characterized how changes in fin stroke amplitude, frequency,<br />

bias angle, <strong>and</strong> curvature affect the thrust <strong>and</strong> lift forces for zero free stream flow<br />

speed. Further testing <strong>and</strong> computations are ongoing to fully characterize the fin<br />

forces <strong>and</strong> vehicle dynamics. However, current models have the necessary fidelity<br />

to accurately predict vehicle performance as outlined in the following sections.<br />

Four-Fin Vehicle Control<br />

With the vehicle state variables defined as in Figure 8, the vehicle dynamics can<br />

be written as,<br />

Mv →̇ þ Cv ð<br />

→ Þv → þ Dv ð →<br />

Þv → þ g →<br />

ðη →<br />

Þ ¼ τ → ;<br />

where M is a matrix of rigid body mass <strong>and</strong> inertial terms, C is a matrix of centripetal<br />

<strong>and</strong> Coriolis terms, D is a matrix of hydrodynamic lift <strong>and</strong> drag terms, g is a<br />

vector of hydrostatic terms, v =[uvwpqr] T , η =[xyzϕθψ] T is the position <strong>and</strong><br />

orientation vector in the earth-fixed frame where ϕ, θ, <strong>and</strong> ψ are roll, yaw, <strong>and</strong><br />

pitch angles, <strong>and</strong> τ is a vector of all forces <strong>and</strong> moments external to the rigid body.<br />

The portion of the vector, τ, that is effected by the fins is represented as,<br />

→<br />

τfins ¼<br />

2<br />

6<br />

4<br />

3<br />

f T ;LF þ f T ;LB þ f T ;RF þ f T ;RB<br />

0<br />

f L;LF f L;LB f L;RF f L;RB<br />

<br />

<br />

y L f L;LF þ f L;LB y R f L;RF þ f L;RB ; ð2Þ<br />

7<br />

x F f L;LF þ f L;RF þ xB f L;LB þ f L;RB<br />

5<br />

<br />

f T ;LF þ f T ;LB f T ;RF þ f T ;RB<br />

y L<br />

y R<br />

ð1Þ<br />

where f T is fin thrust <strong>and</strong> f L is fin lift.<br />

Subscripts ‘LF’, ‘LB’, ‘RF’, <strong>and</strong>‘RB’<br />

identify the left front, left back, right<br />

front, <strong>and</strong> right back fins, respectively.<br />

The x-position of the center of pressure<br />

on the fins is denoted by x F for the<br />

front fins <strong>and</strong> x B for the back fins.<br />

The y-position of the center of pressure<br />

on the fins is denoted by y L for the left<br />

fins <strong>and</strong> y R for the right fins.<br />

Mathematical models representing<br />

thedynamicperformanceofthefins<br />

have been developed to include the effects<br />

on force production of inflow velocities<br />

to the leading edge (or trailing<br />

edge for reverse motion) <strong>and</strong> to include<br />

the effects of fin interactions with each<br />

other, namely the effects of the trailing<br />

vortices off the front fins on the inflow<br />

to the back fins (Geder et al., 2011).<br />

Other fin dynamic representations<br />

modeled controllable parameters including<br />

fin curvature <strong>and</strong> stroke amplitude(Ramamurtietal.,2010).In<br />

the earlier 3-D unsteady CFD studies,<br />

these two key parameters were found<br />

to have a direct relationship with thrust<br />

generation—increasing stroke amplitude<br />

or fin curvature increased thrust<br />

(Ramamurti & S<strong>and</strong>berg, 2006).<br />

However, since both stroke amplitude<br />

<strong>and</strong> flapping frequency are limited mechanically<br />

in the vehicle, optimal combinations<br />

of amplitude <strong>and</strong> frequency<br />

were experimentally found for high<br />

thrust <strong>and</strong> lift fin gaits. The best mix<br />

of these parameters for our vehicle<br />

was determined to be 100° for stroke<br />

amplitude <strong>and</strong> 1.8 Hz flapping frequency,<br />

which yielded not only high<br />

force output but also relatively low<br />

power consumption (Palmisano et al.,<br />

2007). These findings allowed us to fix<br />

stroke amplitude <strong>and</strong> frequency as<br />

constants <strong>and</strong> to focus on fin curvature<br />

as the primary thrust control parameter.<br />

Further, biasing the fin stroke up<br />

or down, as in Figure 9, affects fin lift<br />

78 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 8<br />

Four-fin technology test-bed vehicle, (a) exterior view, (b) interior view.<br />

generation while maintaining constant<br />

thrust (Geder et al., 2008).<br />

In previous work we evaluated the<br />

benefits of two vehicle control methods<br />

(Geder et al., 2008). The first<br />

method, called weighted gait combination,<br />

used combinations of thrustgenerating<br />

<strong>and</strong> lift-generating fin<br />

gaits to produce vectored propulsive<br />

forces. The second method, called<br />

mean bulk angle bias (MBAB), used<br />

weighted forward-reverse gait control<br />

with stroke bias angle control. Between<br />

these two control methods,<br />

our results showed that MBAB better<br />

decoupled control over body-fixed<br />

thrust <strong>and</strong> lift forces <strong>and</strong> yielded better<br />

vehicle response characteristics in simulation.<br />

As such, the MBAB method<br />

is used to control the four-fin vehicle.<br />

The vehicle controller comm<strong>and</strong>s<br />

changes to the fins to effect changes<br />

in the forces <strong>and</strong> moments imparted<br />

on the vehicle, as shown in equation 2.<br />

These fin comm<strong>and</strong>sarebasedon<br />

errors in the vehicle dynamic states,<br />

computed as the comm<strong>and</strong>ed values<br />

minus the computed values. States<br />

are computed onboard the vehicle<br />

using a suite of sensors (three axes<br />

of accelerometers, three axes of rate<br />

gyros, magnetic compass, <strong>and</strong> pressure<br />

sensor) <strong>and</strong> sensor fusion <strong>and</strong> filtering<br />

schemes (Geder et al., 2009). Errors in<br />

surge motion (x-axis translation) dictate<br />

comm<strong>and</strong>s for fin thrust changes<br />

to all fins. Errors in heave motion<br />

(z-axis translation) dictate comm<strong>and</strong>s<br />

for fin lift changes to all fins. Errors<br />

in roll motion (x-axis rotation) dictate<br />

comm<strong>and</strong>s for differential lift changes<br />

between left <strong>and</strong> right fins. Errors in<br />

FIGURE 10<br />

Four-fin vehicle operating in a Naval Research Laboratory test facility.<br />

pitch motion (y-axis rotation) dictate<br />

comm<strong>and</strong>s for differential lift changes<br />

in forward <strong>and</strong> back fins. Errors in yaw<br />

motion (z-axis rotation) dictate comm<strong>and</strong>s<br />

for differential thrust changes<br />

in left <strong>and</strong> right fins. The vehicle has<br />

no direct control over sway motion<br />

( y-axis translation), <strong>and</strong> instead comm<strong>and</strong>s<br />

yaw motion changes to move<br />

in this direction. The direction of<br />

yaw motion depends on the sway<br />

error. The output of a proportionalintegral-derivative<br />

(PID) controller<br />

for each vehicle state is used to determine<br />

forward <strong>and</strong> reverse fin gait percentage<br />

<strong>and</strong> bulk angle comm<strong>and</strong>s.<br />

Four-Fin Vehicle<br />

Performance<br />

Initial measurements of the dynamic<br />

performance of the four-fin vehicle<br />

have been conducted in two test<br />

facilities, one a 6 × 2.5 × 2 foot water<br />

tank <strong>and</strong> the other a 50-foot diameter<br />

by 50-foot deep water tank (Figure 10).<br />

The experimental measurements have<br />

served to validate the vehicle dynamic<br />

model <strong>and</strong> to begin the assessment of<br />

vehicle performance.<br />

FIGURE 9<br />

Vehicle images showing all four fins with<br />

strokes (a) biased down to produce positive<br />

lift <strong>and</strong> (b) biased up to produce negative lift.<br />

July/August 2011 Volume 45 Number 4 79


FIGURE 11<br />

FIGURE 12<br />

Comparison of experimental <strong>and</strong> simulated open-loop heading angle responses (from Geder et al.,<br />

2011).<br />

Comparison of experimental <strong>and</strong> simulated closed-loop heading angle responses with (a) proportional<br />

control <strong>and</strong> (b) PID control (from Geder et al., 2011). The solid curve represents the actual<br />

simulated heading response of the vehicle based on the modeled dynamics. The dashed curve<br />

represents the actual experimental heading response observed in vehicle testing. The square<br />

data represent the measured heading response of the vehicle based on the output of sensor models.<br />

The circle data represent the measured experimental vehicle heading from the output of onboard<br />

sensors.<br />

An open loop test was conducted to<br />

characterize vehicle heading angle response<br />

<strong>and</strong> to compare model simulation<br />

performance with experimental<br />

performance (Geder et al., 2011).<br />

The right finsweresetatfullreversekinematics,<br />

<strong>and</strong> the left fins were set to<br />

closely match the opposite thrust of<br />

the right fins. At t =11s,thegaitweighting<br />

inputs were reversed. The simulated<br />

<strong>and</strong> experimental responses show very<br />

good agreement (Figure 11) with a maximum<br />

difference between the two responses<br />

at any given time of 5°. Both<br />

simulated <strong>and</strong> experimental results exhibit<br />

a 30°/s maximum turning rate,<br />

4 s time from zero to maximum speed,<br />

<strong>and</strong> braking angle of 35°—the amount<br />

of residual turning distance after fin<br />

kinematics are reversed.<br />

After validating the four-fin vehicle<br />

dynamic model in yaw motion <strong>and</strong> implementing<br />

state feedback control,<br />

initial closed-loop experiments were<br />

done to test heading angle control<br />

(Geder et al., 2011). In Figure 12, a<br />

comparison of experimental <strong>and</strong> simulated<br />

results is given. For a simple proportional<br />

control algorithm (Figure 12a),<br />

we see the results match well with a<br />

30-40° amplitude <strong>and</strong> 6.5-s period. Differences<br />

between measured experimental<br />

heading from onboard sensors<br />

<strong>and</strong> simulated heading can be attributed<br />

to the noise <strong>and</strong> sensitivity characteristics<br />

of the sensors (Geder et al.,<br />

2009). External magnetic disturbances<br />

in our test facility cause errors up to<br />

10° in heading measurements, also factoring<br />

into the variation between experimental<br />

<strong>and</strong> simulated heading<br />

responses, as our calibration of the onboard<br />

compass was not perfect. Adding<br />

in a derivative gain to damp the<br />

heading response <strong>and</strong> a small integral<br />

gain to eliminate any steady-state error,<br />

we see the response to a 180° step<br />

comm<strong>and</strong> in heading in Figure 12b.<br />

With PID control over heading angle,<br />

the response is nearly critically damped<br />

with a rise time of 7 s. We also see close<br />

agreement between measured <strong>and</strong> actual<br />

angles, <strong>and</strong> again differences in<br />

the responses can be attributed to sensor<br />

noise <strong>and</strong> sensitivity, as well as compass<br />

calibration errors.<br />

80 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Conclusions<br />

We have reviewed the history<br />

of our bioinspired vehicle design process.<br />

The process began by collaborating<br />

with biologists in order to<br />

underst<strong>and</strong> the dynamics of flapping<br />

<strong>and</strong> deforming fins in pectoral fin<br />

swimmers. The detailed kinematics<br />

they measured in fish swimming experiments<br />

provided the time-varying<br />

fin surface curvature data necessary<br />

for computing the 3-D unsteady flow<br />

about the swimming fish. Examination<br />

of the computed flow variations<br />

about the flapping <strong>and</strong> deforming<br />

fins <strong>and</strong> the fish body provided insights<br />

into the relationship between the fin<br />

flows <strong>and</strong> the fin force time-histories<br />

throughout the stroke cycle. Parametric<br />

variations of key stroke parameters in<br />

3-D unsteady flow computations<br />

yielded the sensitivity of the timevarying<br />

fin forces, which in turn provided<br />

the information needed to modify<br />

our fin designfromthatofthebird<br />

wrasse. It is during such computations<br />

that insights from nature can be<br />

blended with design constraints to<br />

yield a range of possible bio-inspired<br />

designs. A controlled curvature fin utilizing<br />

individually designed fin ribs,<br />

each actuated at the rib root by a linear<br />

actuator was built <strong>and</strong> tested. The successful<br />

fin tests were followed by the<br />

design <strong>and</strong> construction of a two-fin<br />

test-bed vehicle to demonstrate the<br />

mobility potential of the concept. Incorporation<br />

of the fins into a vehicle<br />

necessitated further compromises<br />

where we balanced biomimetic performance<br />

with produceability <strong>and</strong> cost.<br />

These test-bed vehicle tests were followed<br />

by the design <strong>and</strong> construction<br />

of a more capable four-fin vehicle, incorporating<br />

the same fins. A series of<br />

controllers were developed <strong>and</strong> built<br />

to enable assessment of vehicle propulsion<br />

<strong>and</strong> maneuvering performance as<br />

a function of the varying kinematics of<br />

each fin. Deforming fin force time histories,<br />

including fin-fin unsteady interactions,<br />

have been incorporated into<br />

the vehicle dynamic model used for<br />

controller development. The preliminary<br />

four-fin vehicle dynamic performance<br />

measurements indicate very<br />

good agreement with computed performance.<br />

These initial results are<br />

very encouraging <strong>and</strong> indicate that<br />

our efforts to emulate the positionkeeping,<br />

low-speed maneuvering, <strong>and</strong><br />

hovering performance of the bird<br />

wrasse into a producible <strong>and</strong> low cost<br />

vehicle are bearing fruit.<br />

Lead Author:<br />

Jason D. Geder<br />

Laboratory for Computational<br />

Physics <strong>and</strong> Fluid Dynamics<br />

Naval Research Laboratory<br />

Overlook Avenue, SW,<br />

Washington, DC<br />

Email: jgeder@lcp.nrl.navy.mil<br />

References<br />

Alex<strong>and</strong>er, R.M. 1983. The history of fish<br />

biomechanics. In: Fish Biomechanics, eds.<br />

Webb, P.W., & Weihs, D., 1-35.<br />

New York: Praeger.<br />

Azuma, A. 1992. The Biokinetics of Flying<br />

<strong>and</strong> Swimming. Tokyo: Springer-Verlag.<br />

265 pp.<br />

Barrett, D.S., & Triantafyllo, M.S. 1995.<br />

The design of a flexible hull undersea vehicle<br />

propelled by an oscillating foil. In: Proc.<br />

9th Int. Symp. on Unmanned Untethered<br />

Submersible <strong>Technology</strong>. Durham, NH:<br />

Autonomous Undersea Systems Institute<br />

(AUSI).<br />

Bartol, I.K., Gharib, M., Weihs, D., Webb,<br />

P.W., Hove, J.R., & Gordon, M.S. 2003.<br />

Hydrodynamic stability of swimming in<br />

ostraciid fishes: role of the carapace in the<br />

smooth trunkfish Lactophrys triqueter<br />

(Teleostei: Ostraciidae). J Exp Biol. 206:<br />

725-44. doi: 10.1242/jeb.00137.<br />

Blake, R.W. 1983. Fish Locomotion.<br />

Cambridge, UK: Cambridge University<br />

Press. 228 pp.<br />

Breder, C.M. 1926. The locomotion of fishes.<br />

Zoologica. 4:159-297.<br />

Collar, D., Wainwright, P., & Alfaro, M.<br />

2008. Integrated diversification of locomotion<br />

<strong>and</strong> feeding in labrid fishes. Biol Lett. 4:84-6.<br />

doi: 10.1098/rsbl.2007.0509.<br />

Combes, S., & Daniel, T. 2001. Shape,<br />

flapping, <strong>and</strong> flexion: wing <strong>and</strong> fin design<br />

for forward flight. J Exp Biol. 204:2073-85.<br />

Drucker, E.G., & Lauder, G.V. 1999.<br />

Locomotor forces on a swimming fish:<br />

Three-dimensional vortex wake dynamics<br />

quantified using digital particle image<br />

velocimetry. J Exp Biol. 203:2393-412.<br />

Drucker, E.G., & Lauder, G.V. 2002.<br />

Experimental hydrodynamics of fish<br />

locomotion: Functional insights from wake<br />

visualization. Integr Comp Biol. 42:243-57.<br />

doi: 10.1093/icb/42.2.243.<br />

Geder, J.D., Palmisano, J., Ramamurti, R.,<br />

S<strong>and</strong>berg, W.C., & Ratna, B. 2008. Fuzzy<br />

logic PID-based control design <strong>and</strong> performance<br />

for pectoral fin-propelled unmanned<br />

underwater vehicle. In: Proc. Int. Conf on<br />

Control, Automation <strong>and</strong> Systems. Seoul,<br />

Korea: IEEE. doi: 10.1109/ICCAS.2008.<br />

4694526.<br />

Geder, J.D., Ramamurti, R., Palmisano, J.,<br />

Pruessner, M., Ratna, B., & S<strong>and</strong>berg, W.C.<br />

2009. Sensor data fusion <strong>and</strong> submerged<br />

test results of a pectoral fin propelled UUV.<br />

In: Proc. 16th Intl Symp on Unmanned<br />

Untethered Submersible <strong>Technology</strong>. Durham,<br />

NH: Autonomous Undersea Systems Institute<br />

(AUSI).<br />

Geder, J.D., Ramamurti, R., Palmisano, J.,<br />

Pruessner, M., S<strong>and</strong>berg, W.C., & Ratna, B.<br />

2011. Four-fin bio-inspired UUV: modeling<br />

July/August 2011 Volume 45 Number 4 81


<strong>and</strong> control solutions. ASME Intl Mech Eng<br />

Congress Exposition. IMECE. 2011-64005.<br />

Gharib, M., Pereira, F., Dabiri, D., Hove,<br />

J.R., & Modarress, D. 2002. Quantitative<br />

flow visualization: Toward a comprehensive<br />

flow diagnostic tool. Integr Comp Biol.<br />

42:964-70. doi: 10.1093/icb/42.5.964.<br />

Gibb, A.C., Jayne, B.C., & Lauder, G.V.<br />

1994. Kinematics of pectoral fin locomotion<br />

in the bluegill sunfish Lepomis Macrochirus.<br />

J Exp Biol. 189:133-61.<br />

Kato, N. 2000. Control performance in<br />

the horizontal plane of a fish robot with<br />

mechanical pectoral fins. IEEE J Oceanic Eng.<br />

25:125-9. doi: 10.1109/48.820744.<br />

Kato, N., Ando, Y., Tomokazu, A., Suzuki, H.,<br />

Suzumori, K., K<strong>and</strong>a, T., & Endo, S. 2008.<br />

Elastic pectoral fin actuators for biomimetic<br />

underwater vehicles. In: Bio-mechanisms<br />

of Swimming <strong>and</strong> Flying, eds. Kato, N.,<br />

Kamimura, S., 271-82. Tokyo: Springer.<br />

Kato, N., & Furushima, M. 1996. Pectoral<br />

fin model for maneuver of underwater vehicles.<br />

In: Proc. Symp. on Autonomous<br />

Underwater Vehicle <strong>Technology</strong>. Monterrey,<br />

CA: IEEE. doi: 10.1109/AUV.1996.532400.<br />

Lindsey, C.C. 1978. Form, function, <strong>and</strong><br />

locomotory habits in fish. In: Fish Physiology<br />

Vol. II, eds. Hoar, W.S., & R<strong>and</strong>all, D.J.,<br />

1-100. New York: Academic Press.<br />

Palmisano, J., Geder, J.G., Ramamurti, R.,<br />

Liu, K.J., Cohen, J.J., Mengesha, T., …<br />

Ratna, B. 2008. Design, development <strong>and</strong><br />

testing of flapping fin with actively-controlled<br />

curvature for an unmanned underwater vehicle.<br />

In: Bio-mechanisms of Swimming <strong>and</strong><br />

Flying, eds. Kato, N., & Kamimura, S.,<br />

283-94. Tokyo: Springer. doi: 10.1007/<br />

978-4-431-73380-5_23.<br />

Palmisano, J., Ramamurti, R., Liu, K.J.,<br />

Cohen, J., S<strong>and</strong>berg, W.C., & Ratna, B.<br />

2007. Design of a bio-mimetic controlledcurvature<br />

robotic pectoral fin. In: Proc.<br />

IEEE Intl. Conf. on Robotics <strong>and</strong> Automation.<br />

Rome, Italy: IEEE. doi: 10.1109/ROBOT.<br />

2007.363110.<br />

Ramamurti, R., Geder, J.D., Palmisano, J.,<br />

Ratna, B., & S<strong>and</strong>berg, W.C. 2010. Computations<br />

of flapping flow propulsion for unmanned<br />

underwater vehicle design. AIAA J.<br />

48(1):188-201. doi: 10.2514/1.43389.<br />

Ramamurti, R., Lohner, R., & S<strong>and</strong>berg,<br />

W.C. 1996. Computation of the unsteady<br />

flow past a tuna with caudal fin oscillation.<br />

In: Advances in Fluid Mechanics, eds. Rahman,<br />

M.,&Brebbia,C.,vol.9,169-78.Southampton,<br />

UK: Computational Mechanics Publishing.<br />

Ramamurti, R., Lohner, R., & S<strong>and</strong>berg,<br />

W.C. 1999. Computation of the 3-D<br />

unsteady flow past deforming geometries.<br />

Int J Comput Fluid D. 13:83-99.<br />

doi: 10.1080/10618569908940891.<br />

Ramamurti, R., & S<strong>and</strong>berg, W.C. 2006.<br />

Computational fluid dynamics study for<br />

optimization of a fin design. In: Proc. 24th<br />

AIAA Applied Aerodynamics Conference.<br />

AIAA. AIAA-2006-3658.<br />

Ramamurti, R., S<strong>and</strong>berg, W.C., & Lohner, R.<br />

2004. The influence of fin rigidity <strong>and</strong><br />

gusts on the force production in fish <strong>and</strong><br />

insects: a computational study. In: Proc. 42nd<br />

AIAA Aerospace Sciences Meeting <strong>and</strong> Exhibit.<br />

Am Inst Aeronautics <strong>and</strong> Astronautics.<br />

AIAA-2004-0404.<br />

Ramamurti, R., S<strong>and</strong>berg, W.C., Lohner, R.,<br />

Walker, J.A., & Westneat, M.W. 2002. Fluid<br />

dynamics of flapping aquatic flight in the<br />

bird wrasse: 3-D unsteady computations with<br />

fin deformation. J Exp Biol. 205:2997-3008.<br />

Rozhdestvensky, K.V., & Ryzhov, V.A.<br />

2003. Aerohydrodynamics of flapping-wing<br />

propulsors. Prog Aerosp Sci. 39:585-633.<br />

doi: 10.1016/S0376-0421(03)00077-0.<br />

S<strong>and</strong>berg, W.C., & Ramamurti, R. 2008.<br />

3-D Unsteady computations of flapping flight<br />

in insects, fish, <strong>and</strong> unmanned vehicles.<br />

In: Bio-mechanisms of Swimming <strong>and</strong><br />

Flying, eds. Kato, N., & Kamimura, S.,<br />

205-217. Tokyo: Springer. doi: 10.1007/<br />

978-4-431-73380-5_17.<br />

Sfakiotakis, M., Lane, D.M., & Davies, J.B.C.<br />

1999. Review of fish swimming modes for<br />

aquatic locomotion. J Exp Biol. 24:241-52.<br />

Trease, B.P., Liu, K.J., & Kota, S. 2003.<br />

Biomimetic compliant system for actuatordriven<br />

aquatic propulsor: preliminary results.<br />

In: ASME Intl. Mech. Eng. Congress <strong>and</strong><br />

Exposition. IMECE 2003-41446.<br />

Triantafyllo, M.S., Techet, A.H., & Hover,<br />

F.S. 2004. Review of experimental work in<br />

biomimetic foils. IEEE J Oceanic Eng.<br />

29:585-94. doi: 10.1109/JOE.2004.833216.<br />

Videler, J.J. 1993. Fish Swimming. London,<br />

UK: Chapman <strong>and</strong> Hall. 260 pp.<br />

Wainwright, P., Bellwood, D., & Westneat,<br />

M.W. 2002. Ecomorphology of locomotion<br />

in labrid fishes. Environ Biol Fish. 65:47-62.<br />

doi: 10.1023/A:1019671131001.<br />

Walker, J.A., & Westneat, M.W. 1997.<br />

Labriform propulsion in fishes: Kinematics<br />

of flapping aquatic flight in the bird wrasse<br />

Gomphosus varius (Labridae). J Exp Biol.<br />

200:1549-69.<br />

Walker, J.A., & Westneat, M.W. 2000.<br />

Mechanical performance of aquatic rowing<br />

<strong>and</strong> flying. P Roy Soc Lond B Bio. 267:<br />

1875-81. doi: 10.1098/rspb.2000.1224.<br />

Webb, P.W. 1975. Hydrodynamics <strong>and</strong><br />

energetics of fish propulsion. Bull Fish Res Bd<br />

Can. 190:1-158.<br />

Webb, P.W. 1984. Form <strong>and</strong> function in fish<br />

swimming. Sci Amer. 251:58-68.<br />

Webb, P.W. 2004. Maneuverability—<br />

General issues. IEEE J Oceanic Eng.<br />

29:547-53. doi: 10.1109/JOE.2004.833220.<br />

Westneat, M.W., & Walker, J.A. 1997.<br />

Motor patterns of underwater flight: an<br />

electromyographic study of the pectoral<br />

muscles in the bird wrasse, Gomphosus varius.<br />

J Exp Biol. 200:1881-93.<br />

82 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

A Twistable Ionic Polymer-Metal Composite<br />

Artificial Muscle for <strong>Marine</strong> Applications<br />

AUTHORS<br />

Kwang J. Kim<br />

David Pugal<br />

Active Materials <strong>and</strong> Processing<br />

Laboratory, Department of<br />

Mechanical Engineering,<br />

University of Nevada-Reno<br />

Kam K. Leang<br />

Electroactive Systems <strong>and</strong> Controls<br />

Laboratory, Department of<br />

Mechanical Engineering,<br />

University of Nevada-Reno<br />

Introduction<br />

I<br />

onic polymer-metal composite<br />

(IPMC) material is one of the most<br />

promising active (smart) materials<br />

for developing novel soft biomimetic<br />

actuators <strong>and</strong> sensors, preferably for<br />

underwater applications (Shahinpoor<br />

et al., 1998; Shahinpoor & Kim, 2001;<br />

Kim & Shahinpoor, 2003; Shahinpoor<br />

& Kim, 2005). The advantages of<br />

the IPMC include low driving voltage<br />

(


FIGURE 1<br />

(a) Scanning electron microscope image of the cross-section of a Nafion-based IPMC. (b) Illustrative<br />

movement of cations <strong>and</strong> water molecules inside of an IPMC.<br />

saturated in a polar solvent (such as<br />

water) <strong>and</strong> then an electric field is applied<br />

across the electrodes, the composite<br />

bends. The bending is caused<br />

by induced swelling on the cathode<br />

side of the composite <strong>and</strong> shrinking<br />

on the anode side [see Figure 1(b)]<br />

due to a sudden flux of cations <strong>and</strong><br />

polar solvent (such as water). An oppositely<br />

applied voltage causes bending in<br />

the opposite direction. Conversely,<br />

when an IPMC is mechanically deformed,<br />

charges develop on the electrodes<br />

<strong>and</strong> thus IPMCs can function<br />

as current or voltage sensor (Alici<br />

et al., 2008; Pugal et al., 2010a).<br />

The electromechanical behavior of<br />

IPMCs have many noteworthy applications.<br />

Due to their biocompatibility,<br />

IPMC actuators show great promise in<br />

biomedical devices such as active endoscopes<br />

(Yoon et al., 2007) <strong>and</strong> smart<br />

catheters (Fang et al., 2007). Strips of<br />

IPMCs can be used as sensors in<br />

h<strong>and</strong> prostheses (Biddiss & Chau,<br />

2006). But one of the most promising<br />

applications is innovative propulsion<br />

systems for underwater autonomous<br />

systems (Fish et al., 2008; Kamamichi<br />

et al., 2006; Kim et al., 2005,<br />

2007; Lauder, 2007). Specifically,<br />

IPMCs can replace or enhance the<br />

design of propulsors for underwater<br />

walking <strong>and</strong> swimming machines that<br />

are currently based on traditional actuators<br />

such as DC motors (Ayers &<br />

Witting, 2007; Bozkurttas et al., 2008;<br />

Buchholz et al., 2008; Kato, 1998;<br />

Krieg & Mohseni, 2008; Tangorra<br />

et al., 2007), pneumatic actuators<br />

(Cai et al., 2010), <strong>and</strong> magnetic actuators<br />

(Tortora et al., 2010). In fact,<br />

strips of IPMCs have been used to construct<br />

artificial tentacles for a jellyfishlike<br />

robot (Guo et al., 2007). The<br />

walking speed of the jellyfish robot<br />

was controlled through the frequency<br />

of the input voltage applied to the<br />

IPMC-based legs. Likewise, an artificial<br />

caudal fin to propel a robotic fish<br />

was created from an IPMC actuator<br />

(Chen et al., 2010; Guo et al., 2006;<br />

Aureli et al., 2010a). The achievable<br />

peak swimming speed of the robotic<br />

fish in (Chen et al., 2010) was reported<br />

at 22 mm/s. In terms of performance,<br />

the maximum (stall) torque to weight<br />

ratio between a prototype twistable<br />

IPMC AM with dimensions 50 mm ×<br />

25 mm × 1 mm is comparable to a<br />

small ungeared DC motor as illustrated<br />

in the comparison shown in Figure 2.<br />

The comparable performance of<br />

IPMCs <strong>and</strong> traditional actuators suggests<br />

that IPMCs can play a critical<br />

role in the development of highly<br />

maneuverable <strong>and</strong> efficient marine<br />

systems, e.g., the system described in<br />

Menozzi et al. (2008).<br />

Due to the nature of the material<br />

deformation caused by cation <strong>and</strong><br />

solvent flux, bending motion is the<br />

most commonly studied <strong>and</strong> applied<br />

for IPMCs (Kim et al., 2007). Particularly,<br />

when an IPMC strip is mounted<br />

in the cantilever configuration, with<br />

one end fixed <strong>and</strong> the other free, an<br />

applied electric field to the IPMC as<br />

illustrated causes the actuator to bend<br />

as illustrated in Figure 3(a). This single<br />

degree-of-freedom bending motion<br />

has wide applications, such as a singlelink<br />

(Aureli et al., 2010a) <strong>and</strong> multilink<br />

(Yim et al., 2007) oscillatory<br />

propulsor. However, IPMCs with<br />

multiple degrees of freedom are highly<br />

desirable to create control surfaces,<br />

which can undergo complex motion<br />

<strong>and</strong> deformation, for both station<br />

keeping (B<strong>and</strong>yopadhyay et al.,<br />

2008) as well as propulsion <strong>and</strong> maneuvering.<br />

It has been observed that<br />

the propulsion <strong>and</strong> maneuvering<br />

capabilities of the Bluegill (Lepomis<br />

marcrochirus) sunfish is primarily due<br />

to its highly deformable pectoral fin<br />

(Bozkurttas et al., 2008). Thus, multiple<br />

degrees-of-freedom IPMC actuator<br />

technology offers many possibilities for<br />

mimicking such behavior to create<br />

more efficient <strong>and</strong> maneuverable<br />

underwater systems (Chen & Tan,<br />

2010). As depicted in Figure 3(b),<br />

twisting motion can be achieved by<br />

patterning sectored electrodes combined<br />

with independent control each<br />

isolated region. By carefully creating<br />

electrodes on the surface of the<br />

84 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Comparison of torque-to-weight ratios for traditional motors to twistable IPMC AM fin. IPMC AM fin<br />

was fixed on one end, then 5-V DC input was applied to each electrode, with opposing polarity, <strong>and</strong><br />

the blocking torque of the free end was measured with an ATI Nano17 force/torque sensor.<br />

polymer with proper electrical isolation<br />

between adjacent units, sections<br />

of the AM fin can be independently<br />

controlled to achieve complex shapes<br />

FIGURE 3<br />

IPMC AM fin motion: (a) bending <strong>and</strong> (b) twisting.<br />

<strong>and</strong> deformations. Additionally, isolated<br />

electrodes can be patterned for<br />

sensing motion <strong>and</strong> deformation of<br />

the AM fin (Kruusamae et al., 2009).<br />

IPMC Manufacturing<br />

Basic Structure of IPMC <strong>and</strong><br />

Nafion Membrane Fabrication<br />

A basic IPMC consists of an ion exchange<br />

polymer, such perfluorinated<br />

alkenes or styrene/divinylbenzenebased<br />

polymers, s<strong>and</strong>wiched between<br />

two noble metallic electrodes as<br />

shown in Figure 1(a). The conducting<br />

media can be palladium, silver, gold,<br />

carbon, graphite, <strong>and</strong> even nanotubes;<br />

however, platinum is the most commonly<br />

used. The metal electrodes are<br />

often chemically deposited on the<br />

polymer’s surface through a reduction<br />

process (Kim & Shahinpoor, 2003).<br />

Conductive paint can also be applied<br />

to the surface of the membrane to<br />

serve as an electrode; however, this approach<br />

is not as robust <strong>and</strong> effective as<br />

electrochemical plating.<br />

The commonly used ion exchange<br />

membrane Nafion (Dupont) for manufacturing<br />

IPMCs is easily available<br />

from distributors. In the case of<br />

Nafion, the typical chemical structure<br />

is shown in Figure 4, <strong>and</strong> it consists of<br />

fluorocarbons, oxygen, sulfonate<br />

groups, <strong>and</strong> a mobile cation, which<br />

are typically either hydrogen, sodium,<br />

or lithium. Commercially available<br />

Nafion membrane such as Nafion<br />

115, Nafion 117, <strong>and</strong> Nafion 1110<br />

for fabricating IPMCs have nominal<br />

dry thicknesses of 127, 178, <strong>and</strong><br />

254 μm, respectively (Aureli et al.,<br />

2010a). Methods to enhance the performance<br />

of IPMCs include boosting<br />

the capacitance of the composite<br />

(Akle et al., 2005; Aureli et al.,<br />

2009), where this is motivated by studies<br />

that correlate actuation <strong>and</strong> sensing<br />

performancewithcapacitance(Akle<br />

et al., 2005). Enhancements in performance<strong>and</strong>blockingforcehavealso<br />

been made by incorporating nanoparticulates<br />

into the polymer matrix<br />

July/August 2011 Volume 45 Number 4 85


FIGURE 4<br />

The chemical structure of Nafion, which includes fluorocarbons, oxygen, sulfonate groups, <strong>and</strong><br />

a mobile cation X + that can be hydrogen, sodium, or lithium. The K is usually 5–11, <strong>and</strong> the L is<br />

usually 1 (Shahinpoor & Kim, 2001).<br />

(Nam et al., 2003; Nguyen et al., 2007).<br />

In these studies, the nanocompositebased<br />

IPMCs were observed to have<br />

higher water uptake <strong>and</strong> slower water<br />

loss, thus leading to larger bending<br />

displacement <strong>and</strong> blocking force. It<br />

was also found that by using a dispersing<br />

agent in the reduction process<br />

to form fine platinum polycrystals<br />

whichsubsequentlyleadtodeeper<br />

penetration of the platinum layer, the<br />

blocking force was increased significantly<br />

(Shahinpoor & Kim, 2001).<br />

More recent work to improve output<br />

force is by increasing the thickness of<br />

the Nafion membrane (Kim &<br />

Shahinpoor, 2002). Since relatively<br />

thick Nafion membrane is not readily<br />

available, researchers have explored the<br />

solution casting process (Kim &<br />

Shahinpoor, 2002; Kim et al., 2003;<br />

Pak et al., 2004; Shan & Leang,<br />

2009) <strong>and</strong> the hot pressing technique<br />

(Lee et al., 2006). The output force<br />

enhancement for thicker IPMCs is evident<br />

by considering two IPMC strip<br />

actuators, both having the same length<br />

<strong>and</strong> width, but each have a different<br />

thicknesses, such as t <strong>and</strong> 2t (twice as<br />

thick). Assuming that for both actuators<br />

the same tip displacement is required,<br />

then the required strain for<br />

the thick actuator is 2ɛ. As the stress<br />

tensor for linear beam with thickness<br />

of t, width b, <strong>and</strong> length L can be expressed<br />

as (Kim & Shahinpoor, 2002)<br />

σ t ¼ 6F tL<br />

bt 2 ; ð1Þ<br />

the ratio of the stresses is<br />

σ 2t<br />

σ t<br />

≈ 2 ¼ F 2t<br />

2 2 F t<br />

; ð2Þ<br />

FIGURE 5<br />

hence F 2t =8F t .Therefore,athicker<br />

IPMC will produce a larger blocking<br />

force, motivating the need to create<br />

thicker membranes for fabricating<br />

IPMCs. At the same time, the measurements<br />

show that actuation speed<br />

reduces with the increase of the thickness.<br />

Therefore, further study is necessary<br />

to find an optimal thickness of the<br />

membrane for marine applications.<br />

Pretreatment <strong>and</strong> Platinum<br />

Plating Process<br />

The manufacturing of IPMCs begins<br />

with the pretreatment of the ion<br />

exchange membrane (Nafion) <strong>and</strong><br />

the platinum plating process as<br />

outlined in the flowchart shown in<br />

Figure 5. First, the surface of the membrane<br />

is either mechanically roughened<br />

or chemically etched (Yoon et al.,<br />

2007) to either enhance the capacitance<br />

or to improve adhesion of the<br />

metal electrode to the surface. Then,<br />

organic <strong>and</strong> metallic impurities on<br />

thebareNafion membrane are removed<br />

through a pretreatment process<br />

by initially chemically cleaning the<br />

Nafion membrane in 3% hydrogen<br />

peroxide (H 2 O 2 ). Next, the cleaned<br />

membrane is rinsed in 0.5 M sulfuric<br />

acid (H 2 SO 4 )at80°C.Afterwards,<br />

the pretreated <strong>and</strong> cleaned Nafion<br />

membrane is immersed in an appropriate<br />

metal salt solution such as<br />

tetraamineplatinum (II) chloridemonohydrate<br />

[(Pt(NH 3 ) 4 )Cl 2 H 2 O] for<br />

2 h, followed by several washings in<br />

de-ionized water. Platinum particles<br />

are metalized on the surface of the<br />

Nafion membrane by reducing the<br />

Left: The process flow for pretreating the Nafion membrane <strong>and</strong> applying platinum electrodes.<br />

Right: fabricated IPMCs from commercially available Nafion membrane.<br />

86 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


membrane in a sodium borohydride<br />

(NaBH 4 ) or lithium borohydride<br />

(LiBH 4 ) solution for 3 h. Finally, the<br />

platinum-plating process is repeated to<br />

achieve at least three layers of platinum<br />

on the surface of the Nafion membrane<br />

to enhance surface conductivity<br />

<strong>and</strong> overall performance. The platinum<br />

particulate layer is often buried<br />

1-20 μm withintheIPMCsurface<br />

<strong>and</strong> is highly dispersed.<br />

Electrode Patterning Process<br />

The sectored electrodes for the<br />

twistable IPMC AM fin can be created<br />

by masking, surface machining, or ablating<br />

the electrode using a high-power<br />

laser. The first approach involves the<br />

use of a mask to cover areas of the<br />

membrane that should not be exposed<br />

to the platinum-plating process,<br />

subsequently creating an isolation<br />

region between adjacent electrodes.<br />

The second approach uses a precision<br />

computer-controlled milling machine<br />

to mechanically remove the platinum<br />

electrodes from the surface of an IPMC<br />

membrane to create the isolation region.<br />

Masking Technique: The masking<br />

technique to create IPMCs with<br />

sectored electrodes involves the use<br />

of UHMW (Ultra-high-molecularweight)<br />

polyethylene tape (3M). The<br />

tape is used to cover specific regions<br />

of the bare Nafion membrane to inhibit<br />

the plating of platinum on the surface<br />

of the membrane. For example,<br />

the tape is applied to the bare Nafion<br />

membrane, then the taped Nafion<br />

membrane is processed using the pretreatment<br />

<strong>and</strong> plating process described<br />

above <strong>and</strong> outlined in Figure 5. Sample<br />

IPMCs with sectored electrodes are<br />

shown in Figure 6(a), <strong>and</strong> an outline<br />

of the process for the masking technique<br />

is illustrated in Figure 7(a).<br />

Machining Technique: The<br />

surface machining method utilizes a<br />

computer-controlled milling machine,<br />

such as an automated circuit board<br />

router (e.g., ProtoMat S42, LPKF) to<br />

mechanically remove the plated platinum<br />

metal on the surface of the Nafion<br />

membrane. The surface machining<br />

process is outlined in Figure 7(b) <strong>and</strong><br />

described as follows:<br />

1. Create the machining path to create<br />

the electrode pattern using CAD<br />

software (e.g., Solidworks) or a circuit<br />

board layout program (e.g.,<br />

Eagle). The result of this step is a<br />

CAD/CAM file, which is ran by<br />

the milling machine.<br />

2. Attach an IPMC sample (with platinum<br />

electrodes) to the working<br />

surface of the milling machine<br />

using an adhesive layer, for example<br />

double-sided tape (see Figure 7(b))<br />

or a vacuum system. Air bubbles<br />

trapped underneath the IPMC<br />

sample should be removed. Additionally,<br />

the locations of the corners<br />

of the IPMC on the working surface<br />

are marked for aligning the<br />

sample.<br />

3. Load the CAD/CAM file onto the<br />

milling machine <strong>and</strong> start the milling<br />

process.<br />

4. Remove the machined IPMC sample,<br />

then flip it over <strong>and</strong> attach the<br />

sample to the working surface making<br />

sure the corners are aligned with<br />

the markings created in Step 2.<br />

Repeat Steps 2 <strong>and</strong> 3.<br />

5. Remove the machined IPMC sample<br />

<strong>and</strong> trim away excess material.<br />

Sample IPMCs with sectored<br />

electrodes created by the surface<br />

machining process are shown in Figures<br />

6(b1)-6(b4). During the machining<br />

process, care is taken to<br />

avoid removing too much material<br />

for thin IPMC membranes. In general,<br />

the machine should be set to remove<br />

only the platinum material (approximately<br />

25-50 μm deep). Comparing<br />

the masking <strong>and</strong> surface-machining results,<br />

the machining process allows<br />

better control of the shape <strong>and</strong> pattern<br />

of the electrode. One of the major<br />

challenges of the masking process is<br />

ensuring that the tape adheres to the<br />

IPMC membrane’s surface all through<br />

the plating process. During machining,<br />

if the depth of cut is not well-controlled,<br />

removal of excess material affects the<br />

mechanical properties, for example,<br />

stiffness, of the IPMC actuator.<br />

Modeling<br />

Electromechanical<br />

Transduction<br />

Two approaches are discussed to<br />

model the electromechanical bending<br />

<strong>and</strong> twisting response of the sectoredelectrode<br />

IPMCs: (1) a simplified<br />

finite-element analysis (FEA) model<br />

derived from the piezoelectric effect<br />

<strong>and</strong> (2) a more comprehensive physicsbased<br />

model. The former model has<br />

the advantage of being easy to implement<br />

due to fact that software packages<br />

are available <strong>and</strong> specifically tailored to<br />

solve the problem. Additionally, the<br />

established algorithms are computationally<br />

efficient. However, the first<br />

method lacks consideration of the<br />

physical processes that governs the behavior<br />

of the IPMC. The latter model,<br />

on the other h<strong>and</strong>, considers the underlying<br />

physics that includes electrostatic<br />

forces, osmotic pressure, charge<br />

imbalance <strong>and</strong> the effects of local strains<br />

(Kim & Shahinpoor, 2003; Nemat-<br />

Nasser & Jiang, 2000; Tadokoro<br />

et al., 2000; Chen & Tan, 2008; Leo<br />

et al., 2005). This type of model offers<br />

valuable insight on the physical behavior<br />

of the composite material <strong>and</strong><br />

the model can be used to help guide<br />

the development of the material on<br />

many levels. Despite being more<br />

realistic—<strong>and</strong> sometimes more<br />

accurate—the physics-based model<br />

July/August 2011 Volume 45 Number 4 87


FIGURE 6<br />

Fabricated IPMCs with sectored electrodes: samples created using (a) the masking technique <strong>and</strong> (b1)-(b4) the surface-machining approach.<br />

Dimensions are in millimeters (mm).<br />

is more computationally dem<strong>and</strong>ing<br />

to solve. Simulation results comparison<br />

with experimental data are<br />

presented in Performance Characterization<br />

<strong>and</strong> Discussion<br />

Simplified FEA Model for<br />

Electromechanical Response<br />

Asimplified finite-element model<br />

is created to predict <strong>and</strong> gain insight<br />

on the bending <strong>and</strong> twisting capabilities<br />

of the sectored-electrode IPMC<br />

AM fin. Such a model can also be<br />

applied to optimize the design of an<br />

AM fin for specific applications. The<br />

key feature of this simple model is<br />

the deformation is estimated using an<br />

equivalent bimorph beam model analogous<br />

to a piezoelectric actuator (Kim<br />

& Tadokoro, 2007). As a result the<br />

model only captures the basic electromechanical<br />

behavior <strong>and</strong> is thus easy to<br />

implement <strong>and</strong> computationally efficient.<br />

Figure 8 shows the boundary conditions<br />

for the finite-element model,<br />

where the ‘clamped area’ (25 mm ×<br />

5 mm) is considered fixed. The bimorph<br />

finite-element model consists of two<br />

electro-mechanical actuation layers,<br />

layers (i) <strong>and</strong> (iii), as shown in Figure 8.<br />

88 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 7<br />

Electrode patterning process: (a) the masking <strong>and</strong> (b) the surface-machining technique.<br />

The material separating patterned areas<br />

of the material are modeled by layer (ii).<br />

The IPMC material is treated as a<br />

homogenous material (uniform stiffness<br />

<strong>and</strong> density). The st<strong>and</strong>ard stiffness<br />

matrix k, piezoelectric strain matrix d,<br />

<strong>and</strong> permittivity matrix ɛ are used in<br />

the modeling (Moheimani & Fleming,<br />

2006). The material properties are as<br />

follows: density is 2930 kg/m 3 ;<br />

Poisson’s ratio is 0.49; E=1.16GPa;<br />

d 31 = d 32 =4.11×10 -6 m/V (bending);<br />

d 31 = d 32 =2.67×10 −7 m/V (twisting).<br />

The finite-element model is created<br />

using ANSYS software (Canonsburg,<br />

PA). A SOLID98 tetrahedral coupled<br />

field solid element is chosen for the<br />

electro-mechanical material, while a<br />

SOLID187 tetrahedral structural solid<br />

element is chosen for the electrode separation<br />

material. Both material properties<br />

are assigned a tetrahedral solid to compensate<br />

for the irregular mess interface<br />

caused by large aspect ratio between the<br />

IPMCs thickness <strong>and</strong> length. By selecting<br />

these elements, a uniform tetrahedral<br />

mesh throughout the entire model is<br />

achieved.<br />

Physics-Based Model for<br />

Electromechanical Transduction<br />

When a voltage is applied to the<br />

electrodes of an IPMC, the freely movable<br />

cations inside the polymer start<br />

migrating due to the imposed electric<br />

FIGURE 8<br />

Finite-element model structure. Layers (i) <strong>and</strong> (iii) are the bimorph (electro-mechanical) layers, <strong>and</strong> layer (ii) is assumed to be a nonconducting<br />

electrode separation layer (absent of electro-mechanical properties).<br />

July/August 2011 Volume 45 Number 4 89


field. However, the attached anions do not migrate nor diffuse. In case of waterbased<br />

IPMCs, migrating cations drag the water molecules along, causing<br />

osmotic pressure changes <strong>and</strong> therefore swelling near the cathode <strong>and</strong> contraction<br />

of the polymer near the anode. This in turn results in bending of the material<br />

towards the anode side. Furthermore, the migrated cations cause charge imbalance<br />

near the electrodes <strong>and</strong> this possibly results in the electrostatic force contribution<br />

to the local strain. In the following, the basic underlying equations<br />

are presented <strong>and</strong> how to apply the calculated charge imbalance near the<br />

electrodes in modeling the IPMC actuation in a 3-D domain is discussed<br />

(see Figure 9).<br />

First, the ionic migration <strong>and</strong> diffusion are described in the polymer domain by<br />

Nernst-Planck <strong>and</strong> Poisson equations,<br />

∂C<br />

∂t<br />

þ ∇· ð D∇C zμFC∇ϕÞ ¼0;<br />

∇ 2 ϕ ¼ F ð C C 0Þ<br />

; ð4Þ<br />

ɛ<br />

where C is the cation concentration, μ is the mobility of counter ions, D is the diffusion<br />

constant, F is the Faraday constant, z is the charge number, ϕ is the electric<br />

potential, <strong>and</strong> C 0 is the anion concentration with initial value of C 0 =1200mol/m 3 .<br />

For the electrode, Ohm’s law for the current density is<br />

σ∇V ¼ → j; ð5Þ<br />

where σ is electric conductivity, V is the voltage, <strong>and</strong> → j is the current density in the<br />

electrode. It must be noted that the electric potential ϕ inside the polymer <strong>and</strong> the<br />

electric potential V in the electrode are different. To relate the variables of<br />

the electrodes V <strong>and</strong> → j to the ionic flux inside the polymer, the Ramo-Shockley<br />

FIGURE 9<br />

Conceptual physics-based model. Calculations are done in three domains—the polymer domain<br />

<strong>and</strong> two electrode domains.<br />

ð3Þ<br />

theorem is used (Ramo, 1939; Shockley,<br />

1938), i.e.,<br />

I ¼ 1 φ ∑ →<br />

n q nW ð<br />

→<br />

r n Þ· →v n ;<br />

ð6Þ<br />

where → r n ; → v n ,<strong>and</strong> → q n are the position vector,<br />

instantaneous velocity, <strong>and</strong> charge<br />

of cation n, respectively. The electric<br />

field that would be produced by 1 V applied<br />

potential without any charges,<br />

neither mobile nor fixed, is denoted by<br />

→<br />

W . The current in the external circuit is<br />

represented by I, <strong>and</strong>ϕ is a constant<br />

with value of 1 V. Equation (6) can be<br />

simplified for a 2-D domain with parallel<br />

electrodes<br />

j ¼ 1 h ∫ f dl;<br />

ð7Þ<br />

where f is an ionic flux in the electrode<br />

direction <strong>and</strong> has a unit of Cm 2 s <strong>and</strong> j is<br />

local current density at the inner boundary<br />

of an electrode. The term dl is the<br />

integration element along the path<br />

where the particle moves, which in this<br />

case is assumed to be from one electrode<br />

to other.<br />

The Equations (3)-(7) are used to<br />

calculate the time dependent cation<br />

concentration C <strong>and</strong> corresponding<br />

charge density ρ = C − C 0 in the polymer<br />

of IPMC in response to an applied<br />

voltage. To relate the charge density<br />

ρ to the physical bending, the force<br />

coupling similar to the one shown in<br />

Pugal et al. (2008) is used. A set of<br />

continuum mechanics equations were<br />

implemented for the polymer domain.<br />

Normal <strong>and</strong> shear strain are by<br />

definition<br />

ɛ i ¼ ∂u i<br />

; ɛ ij ¼ 1 <br />

∂u i<br />

þ ∂u <br />

j<br />

; ð8Þ<br />

∂x i 2 ∂x j ∂x i<br />

where u is the displacement vector,<br />

x denotes a coordinate <strong>and</strong> indices i<br />

90 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


̂<br />

<strong>and</strong> j are in the range of 1-3 <strong>and</strong> denote<br />

components correspondingly to x, y,or<br />

z direction. The stress-strain relationship<br />

is<br />

σ f ¼ Dɛ;<br />

ð9Þ<br />

where D is a 6 × 6 elasticity matrix,<br />

consisting of components of Young’s<br />

modulus <strong>and</strong> Poisson’s ratio. The system<br />

is in equilibrium, if the relation<br />

∇· σ f ¼ → F ð10Þ<br />

is satisfied. This is the Navier’s equation<br />

for displacement (Heinbockel,<br />

2001). The body force <strong>and</strong> charge coupling<br />

is defined as<br />

→<br />

F ¼ Aρx; ̂<br />

ð11Þ<br />

the range of hundreds of thous<strong>and</strong>s<br />

of degrees of freedom. The challenges<br />

are described in more detail in Pugal<br />

et al. (2010b). To reduce the problem<br />

size without significant loss in the<br />

calculation precision, the following<br />

modeling approach for 3-D actuation<br />

of IPMC is developed.<br />

Firstly, the cation concentration<br />

C (from which the charge density ρ<br />

can be directly calculated) <strong>and</strong> voltage<br />

ϕ are calculated in a 2-D domain.<br />

As the 2-D domain is scaled in the longitudinal<br />

(x) direction of an IPMC, the<br />

electrode conductivity value is also<br />

linearly scaled. The electrode currents<br />

<strong>and</strong> corresponding voltage gradient in<br />

the electrodes are taken into account in<br />

the 2-D model to obtain more precise<br />

cation concentration. See, for instance,<br />

Figure 10—there the molar flux f<br />

<strong>and</strong> electric current j are depicted. As<br />

a result of the calculations, spatial<br />

<strong>and</strong> temporal C <strong>and</strong> ϕ in 2-D are<br />

found <strong>and</strong> stored. This is done for<br />

each patterned electrode that is subjected<br />

to a different voltage. Due to<br />

thesmallDebyescreeninglengthof<br />

the charges near the electrodes (Porfiri,<br />

2009), the 2-D model calculations are<br />

done on a mesh that is extremely fine<br />

near the boundaries.<br />

Secondly, the calculated C <strong>and</strong> ϕ are<br />

extruded onto slightly coarser mesh in<br />

the 3-D domain as illustrated in<br />

Figure 11. For instance, Figure 12<br />

shows the extruded voltage ϕ values<br />

on the 3-D domain boundary. It<br />

should be noted that in the current<br />

model, it is not necessary to extrude<br />

the values of ϕ, but it was done for<br />

illustration purposes. Finally, the<br />

where A is a parametrically determined<br />

constant <strong>and</strong> x is IPMC’s longitudinal<br />

direction. This approach allows calculating<br />

deformation that is in the typical<br />

IPMC actuation range. When a very<br />

large deformation is expected, for instance,<br />

in case of a very thin membrane,<br />

geometric nonlinearity <strong>and</strong><br />

more precise force coupling must be<br />

used in the model. The conceptual<br />

model with the variables is illustrated<br />

in Figure 9.<br />

The finite element method is used<br />

to solve Equations (3), (4), <strong>and</strong> (5)<br />

with Equation (7) as the boundary<br />

condition between the electrode <strong>and</strong><br />

polymer domain <strong>and</strong> Equation (10).<br />

The equations are implemented in<br />

Comsol Multiphysics software package<br />

(Multiphysics, 2011). The high aspect<br />

ratio of the domain representing<br />

an IPMC <strong>and</strong> the nonlinear nature<br />

of the problem make it difficult to<br />

directly solve the equations in a<br />

full-scale 3-D domain. Namely, the<br />

problem size would be very large, in<br />

FIGURE 10<br />

Electric current streamlines in the electrodes <strong>and</strong> total ionic flux streamlines in the polymer<br />

A<br />

domain. The color depicts the total current density m2<br />

in the electrodes. (Color versions of<br />

figures available online at: http://www.ingentaconnect.com/content/mts/mtsj/2011/00000045/<br />

00000004.)<br />

July/August 2011 Volume 45 Number 4 91


FIGURE 11<br />

Extruding the cation concentration C that has been calculated for each<br />

time step in 2-D into a 3-D domain.<br />

FIGURE 12<br />

Extruded data from 2-D domain. The surface depicts the extruded voltage<br />

values on the electrodes. Notice the voltage gradient in the longitudinal<br />

x direction.<br />

coupling Equations (11) <strong>and</strong> (10) are<br />

carried out to calculate the timedependent<br />

bending of an IPMC in 3-D.<br />

Performance<br />

Characterization<br />

<strong>and</strong> Discussion<br />

The bending <strong>and</strong> twisting performance<br />

of the fabricated IPMC AM<br />

fins are characterized <strong>and</strong> then compared<br />

to the electromechanical models.<br />

The experimental setup consists of<br />

control electronics connected to the<br />

IPMC AM fin <strong>and</strong> a measurement system<br />

(laser displacement sensors <strong>and</strong><br />

data acquisition system) for collecting<br />

the output response. Each set of electrodes<br />

are independently controlled<br />

by a separate custom-designed voltage<br />

amplifier as depicted in Figure 13. A<br />

complete description of the experimental<br />

setup is described in Riddle<br />

et al. (2010). All measurements are<br />

taken while the subject IPMC is actuated<br />

in de-ionized water.<br />

Measured Bending <strong>and</strong><br />

Twisting Performance<br />

Themeasuredresponseforaselected<br />

masked <strong>and</strong> machined electrode<br />

IPMC, Figure 6(b2), are shown in Figure<br />

14. The results in Figure 14 show<br />

that both types of actuators provided<br />

approximately the same degree of<br />

FIGURE 13<br />

twist. Furthermore, the actuators also<br />

exhibited non-smooth twisting<br />

motion, which may have been caused<br />

by the effects of the fabrication process.<br />

A peak twist angle of 7.3° is measured<br />

for the machined IPMC [see<br />

Figures 14(c) <strong>and</strong> 14(d)]. The frequency<br />

of the actuation of 1 Hz is<br />

Experimental setup for measuring bending <strong>and</strong> twisting response using two non-contact laser<br />

sensors (Micro-Epsilon, optoNCDT 1402). Voltage amplifier gain is A.<br />

92 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


sufficient for underwater propulsion<br />

(Chen et al., 2010). The bending<br />

angle could be enhanced in various<br />

ways, depending on the application<br />

of interest. For instance, rigid extensions<br />

with different shapes can be<br />

used (Anton, 2008).<br />

FIGURE 14<br />

Measured IPMC twisting response for 5 V sinusoidal input at 1 Hz: (a) input voltage applied to<br />

electrodes; (b) twisting response for masked electrode IPMC; (c) twisting response for machined<br />

electrode IPMC; (d) sequence images of machined electrode IPMC showing twisting behavior.<br />

Simulated Responses<br />

The bending <strong>and</strong> twisting response<br />

of the IPMC, Figure 6(b2), produced<br />

by the simple FEA model for an<br />

input voltage of 2 V are shown in<br />

Figures 15(a) <strong>and</strong> 15(b). For bending,<br />

the tip displacement is predicted at<br />

approximately 2.5 mm. For twisting,<br />

the angle of twist is estimated using<br />

θ ¼ tan 1 a b<br />

, where a <strong>and</strong> b are<br />

shown in Figure 15(c). The predicted<br />

maximum angle of twist is approximately<br />

0.7° for a 2-V input. In Figures<br />

15(d) <strong>and</strong> 15(e), the FEA results<br />

<strong>and</strong> the response for the bending <strong>and</strong><br />

twisting motion measured along the<br />

length of the IPMC actuator are<br />

compared. For bending, the predicted<br />

<strong>and</strong> measured results agreed well,<br />

where the maximum root-meansquared<br />

(RMS) error is approximately<br />

0.2 mm. For bending, however, the<br />

RMS error increased significantly for<br />

3 V <strong>and</strong> beyond. It is noted that the<br />

finite-element approach assumes the<br />

material is isotropic <strong>and</strong> operating<br />

within the linear-elastic range. Due to<br />

the complex nature of the IPMC<br />

material, the simplified FEA may be<br />

limited in its ability to accurately<br />

predict the performance for large<br />

electric fields. For example, as shown<br />

in Figure 14, the measured twist<br />

angle is approximately 7° for an input<br />

voltage with magnitude of 5 V. This<br />

result indicates that the IPMC’s response<br />

can be highly nonlinear for<br />

large input voltages. Therefore, one<br />

of the challenges is developing detailed<br />

enough models to aid in designing the electrode patterns to meet a specific<br />

application.<br />

To adequately model the twisting deformation with the physics-based electromechanical<br />

model presented in Physics-Based Model for Electromechanical<br />

Transduction, the sinusoidal voltages<br />

u 1 ðÞ¼5 t ½V<br />

Šsinð2πtÞ; u 2 ðÞ¼ t<br />

5½V<br />

Šsinð2πtÞ; ð12Þ<br />

are set as time-dependent boundary conditions to the electrodes 1, 2 <strong>and</strong> 3, 4,<br />

respectively (see Figure 13). The calculated displacement fields at two different<br />

times are shown in Figure 16. The model is validated against measured time response<br />

[Figure 14(c)]. The model estimation versus measurements are shown in<br />

Figure 17. It can be seen that the model predicts the twisting deformation well.<br />

Slight discrepancies in the peak values can be attributed to the fact that the model<br />

July/August 2011 Volume 45 Number 4 93


FIGURE 15<br />

Finite-element modeling results for (a) bending <strong>and</strong> (b) twisting for 2-V input. (c) Definition of twist angle θ. Comparison of finite-element model<br />

results <strong>and</strong> measured response: (d) bending <strong>and</strong> (e) twisting response for different applied voltages.<br />

does not take into account electrochemical<br />

currents <strong>and</strong> therefore does<br />

not provide correct voltage gradient on<br />

the electrodes for higher applied voltages.<br />

It must be noted that the hydrodynamic<br />

effects are not considered in<br />

the calculations due to the low frequency<br />

<strong>and</strong> rather small twisting amplitude.<br />

Conclusions<br />

IPMC AMs are suited for creating<br />

artificial fish-like propulsors that can<br />

mimic the undulation, flapping, <strong>and</strong><br />

complex motions of fish fins. A newly<br />

developed IPMC AM fin withpatterned<br />

electrodes was introduced for<br />

realizing multiple degrees-of-freedom<br />

motion, such as bending <strong>and</strong> twisting.<br />

These twistable AM fins are suited for<br />

creating artificial fish-like propulsors<br />

that can mimic the undulatory, flapping,<br />

<strong>and</strong> complex motions of fish<br />

fins as well as novel control surfaces<br />

for applications in a wide spectrum of<br />

micro-autonomous robots <strong>and</strong> marine<br />

systems. The masking <strong>and</strong> surface machining<br />

fabrication process are viable<br />

approaches to create a twistable AM<br />

fin. Experimental characterization<br />

showed that peak twisting for the<br />

sectored-electrode IPMC exceeded 7°.<br />

A finite-element bimorph beam<br />

model was used to predict the bending<br />

<strong>and</strong> twisting behavior of a selected<br />

IPMC actuator, where good agreement<br />

between the measured response <strong>and</strong><br />

model output was found for low<br />

electric fields (2 V). A full-scale 3-D<br />

94 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 16<br />

Calculated twisting of IPMC at t = 0.25 s (left) <strong>and</strong> at t = 0.75 s (right). The color shows y-directional displacement.<br />

physics-based model to simulate electromechanical<br />

twisting transduction<br />

of IPMC was developed. Comparison<br />

between the experimental results <strong>and</strong><br />

model output for the electromechanical<br />

transduction indicated that the<br />

FIGURE 17<br />

Comparison of experimental (solid line) <strong>and</strong> simulated twisting angle (dash line).<br />

model predicts the twisting deformation<br />

well. The future work will explore<br />

complex electrode patterns, integrated<br />

sensing electrodes, an alternative<br />

approach to create a twistable fin structure<br />

using a soft boot, <strong>and</strong> the development<br />

of a prototype autonomous<br />

marine system powered by the twistable<br />

IPMC AM fin.<br />

Acknowledgments<br />

The authors gratefully thank<br />

the financial support from the U.S.<br />

Office of Naval Research (grant<br />

N0001409102183) <strong>and</strong> Dr. Tom<br />

McKenna. The authors also thank<br />

S.M. Kim, Y.S. Jung, S. Song,<br />

R. Riddle, <strong>and</strong> Y. Shan for their help<br />

with the experiments. KJK expresses<br />

his special thanks to Dr. Promode<br />

B<strong>and</strong>yopadhyay of the Naval Undersea<br />

Warfare Center (NUWC) in<br />

Newport, RI, for his thoughtful<br />

encouragement.<br />

Lead Authors:<br />

Kwang J. Kim<br />

Active Materials <strong>and</strong><br />

Processing Laboratory<br />

Department of Mechanical<br />

Engineering, University<br />

of Nevada-Reno<br />

Email: kwangkim@unr.edu<br />

July/August 2011 Volume 45 Number 4 95


Kam K. Leang<br />

Electroactive Systems <strong>and</strong><br />

Controls Laboratory<br />

Department of Mechanical<br />

Engineering, University<br />

of Nevada-Reno<br />

Email: kam@unr.edu<br />

References<br />

Anton, M. 2008. Mechanical modeling<br />

of IPMC actuators at large deformations.<br />

Ph.D. Thesis, University of Tartu.<br />

Akle, B.J., Leo, D.J., Hickner, M. A., &<br />

McGrath, J.E. 2005. Correlation of capacitance<br />

<strong>and</strong> actuation in ionomeric polymer<br />

transducers. J Mater Sci. 40:3715-24.<br />

doi: 10.1007/s10853-005-3312-x.<br />

Alici, G., Spinks, G.M., Madden, J.D.,<br />

Wu, Y., & Wallace, G.G. 2008. Response<br />

characterization of electroactive polymers as<br />

mechanical sensors. IEEE/ASME T Mech.<br />

13(2):187-96.<br />

Aureli, M., Kopman, V., & Porfiri, M.<br />

2010a. Free-locomotion of underwater<br />

vehicles actuated by ionic polymer metal<br />

composites. IEEE/ASME T Mech.<br />

15(4):603-14.<br />

Aureli, M., Lin, W., & Porfiri, M. 2009.<br />

On the capacitance-boost of ionic polymer<br />

metal composites due to electroless plating:<br />

Theory <strong>and</strong> experiments. J Appl Phys.<br />

105:104911. doi: 10.1063/1.3129503.<br />

Aureli, M., Prince, C., Porfiri, M., &<br />

Peterson, S.D. 2010b. Energy harvesting from<br />

base excitation of ionic polymer metal<br />

composites in fluid environments. Smart<br />

Mater Struct. 19:015003. doi: 10.1088/<br />

0964-1726/19/1/015003.<br />

Ayers, J., & Witting, J. 2007. Biomimetic<br />

approaches to the control of underwater<br />

walking machines. Philos T R Soc A.<br />

365:273-95. doi: 10.1098/rsta.2006.1910.<br />

B<strong>and</strong>yopadhyay, P.R., Beal, D.N., &<br />

Menozzi, A. 2008. Biorobotic insights into<br />

how animals swim. J Exp Biol. 211:206-14.<br />

doi: 10.1242/jeb.012161.<br />

Bhat, N.D., & Kim, W.-J. 2004. Precision<br />

force <strong>and</strong> position control of ionic polymer-metal<br />

composite. J Syst Contr Eng. 218(6):421-32.<br />

Biddiss, E., & Chau, T. 2006. Electroactive<br />

polymeric sensors in h<strong>and</strong> prostheses: Bending<br />

response of an ionic polymer metal composite.<br />

Med Eng Phys. 28(6):568-78. doi: 10.1016/<br />

j.medengphy.2005.09.009.<br />

Bozkurttas, M., Tangorra, J., Lauder, G., &<br />

Mittal, R. 2008. Underst<strong>and</strong>ing the hydrodynamics<br />

of swimming: From fish fins to<br />

flexible propulsors for autonomous underwater<br />

vehicles. Adv Sci Tech. 58:193-202.<br />

doi: 10.4028/www.scientific.net/AST.58.193.<br />

Brufau-Penella, J., Puig-Vidal, M., Giannone,<br />

P., Graziani, S., & Strazzeri, S. 2008. Characterization<br />

of the harvesting capabilities of<br />

an ionic polymer metal composite device.<br />

Smart Mater Struct. 17(1):015009.<br />

doi: 10.1088/0964-1726/17/01/015009.<br />

Buchholz, J.H.J., Clark, R.P., & Smits, A.J.<br />

2008. Thrust performance of unsteady propulsors<br />

using a novel measurement system,<br />

<strong>and</strong> corresponding wake patterns. Exp<br />

Fluids. 45:461-72. doi: 10.1007/s00348-<br />

008-0489-1.<br />

Cai, Y., Bi, S., & Zheng, L. 2010. Design<br />

<strong>and</strong> experiments of a robotic fish imitating<br />

wow-nosed ray. J Bionic Eng. 7:120-126.<br />

doi: 10.1016/S1672-6529(09)60204-3.<br />

Chen, Z., Shatara, S., & Tan, X. 2010.<br />

Modeling of biomimetic robotic fish propelled<br />

by an ionic polymer-metal composite caudal<br />

fin. IEEE/ASME T Mech. 15:448-59.<br />

Chen, Z., & Tan, X. 2008. A controloriented<br />

<strong>and</strong> physics-based model for ionic<br />

polymer-metal composite actuators. IEEE/<br />

ASME T Mech. 13(5):519-29.<br />

Chen, Z., & Tan, X. 2010. Monolithic<br />

fabrication of ionic polymer-metal composite<br />

actuators capable of complex deformation.<br />

Sensor Actuat A-Phys. 157:246-57.<br />

doi: 10.1016/j.sna.2009.11.024.<br />

Chen, Z., Um, T., & Bart-Smith, H. 2011.<br />

A novel fabrication of ionic polymer-metal<br />

composite actuator capable of 3-dimensional<br />

kinematic motions. Sensor Actuat A-Phys.<br />

168(1):131-9. doi: 10.1016/j.<br />

sna.2011.02.034.<br />

Fang, B.-K., Ju, M.-S., & Lin, C.-C. K. 2007.<br />

A new approach to develop ionic polymermetal<br />

composites (IPMC) actuator: Fabrication<br />

<strong>and</strong> control for active catheter systems. Sensor<br />

Actuat A-Phys. 137(2):321-9. doi: 10.1016/<br />

j.sna.2007.03.024.<br />

Fish, F.E., Howle, L.E., & Murray, M.M.<br />

2008. Hydrodynamic flow control in marine<br />

mammals. Integr Comp Biol. 48:1-13.<br />

doi: 10.1093/icb/icn029.<br />

Guo, S., Ge, Y., Li, L., & Liu, S. 2006.<br />

Underwater swimming micro robot using<br />

IPMC actuator. In: Proceedings of the<br />

2006 IEEE International Conference on<br />

Mechatronics <strong>and</strong> Automation, June 25-28.<br />

pp. 249-54. Piscataway, NJ: IEEE.<br />

doi: 10.1109/ICMA.2006.257505.<br />

Guo, S., Shi, L., Ye, X., & Li, L. 2007. A<br />

new jellyfish type of underwater microrobot.<br />

In: Proc. IEEE International Conference on<br />

Mechatronics <strong>and</strong> Automation. pp. 509-14.<br />

Piscataway, NJ: IEEE.<br />

Heinbockel, J. 2001. Introduction to<br />

Tensor Calculus <strong>and</strong> Continuum Mechanics.<br />

Bloomington, IN: Trafford Publishing.<br />

Jeon, J.H., & Oh, I.K. 2009. Selective growth<br />

of platinum electrodes for MDOF IPMC<br />

actuators. Thin Solid Films. 517:5288-92.<br />

doi: 10.1016/j.tsf.2009.03.111.<br />

Jeon, J.H., Oh, I.K., Han, J.H., & Lee,<br />

S.W. 2007. Development of bio-mimetic<br />

patterned IPMC actuators with multiple<br />

electrodes. Key Eng Mat. 334:1005-8.<br />

doi: 10.4028/www.scientific.net/KEM.<br />

334-335.1005.<br />

Kamamichi, N., Yamakita, M., Asaka, K.,<br />

& Luo, Z.-W. 2006. A snake-like swimming<br />

robot using IPMC actuator/sensor.<br />

In: IEEE International Conference on<br />

Robotics <strong>and</strong> Automation. pp. 1812-7.<br />

Piscataway, NJ: IEEE.<br />

Kang, S., Shin, J., Kim, S.J., Kim, H.J., &<br />

Kim, Y.H. 2007. Robust control of ionic<br />

polymer-metal composites. Smart Mater<br />

96 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Struct. 16(6):2457-63. doi: 10.1088/<br />

0964-1726/16/6/049.<br />

Kato, N. 1998. Locomotion by mechanical<br />

pectoral fins. J Mar Sci Technol. 3:113-21.<br />

doi: 10.1007/BF02492918.<br />

Kim, B., Kim, B.M., Ryu, J., Oh, I.-H.,<br />

Lee, S.-K., Cha, S.-E., & Pak, J. 2003.<br />

Analysis of mechanical characteristics of the<br />

ionic polymer metal composite (IPMC)<br />

actuator using cast ion-exchange film. In:<br />

Proceeding of the SPIE Smart Structures<br />

<strong>and</strong> Materials: Electroactive Polymer Actuators<br />

<strong>and</strong> Devices (EAPAD). pp. 486-95.<br />

San Diego, CA: SPIE.<br />

Kim, B., Kim, D.-H., Jung, J.J., & Park,<br />

J.-O. 2005. A biomimetic undulatory tadpole<br />

robot using ionic polymer-metal composite<br />

actuators. Smart Mater Struct. 14:1579-85.<br />

Kim, K.J., & Shahinpoor, M. 2002. A novel<br />

method of manufacturing three-dimensional<br />

ionic polymer-metal composites (ipmcs)<br />

biomimetic sensors, actuators <strong>and</strong> artificial<br />

muscles. Polymer. 43(3):797-802.<br />

doi: 10.1016/S0032-3861(01)00648-6.<br />

Kim, K.J., & Shahinpoor, M. 2003. Ionic<br />

polymer-metal composites: II. Manufacturing<br />

techniques. Smart Mater Struct. 12(1):65-79.<br />

doi: 10.1088/0964-1726/12/1/308.<br />

Kim, K.J., & Tadokoro, S. 2007. Electroactive<br />

Polymers robotics Applications: Artificial<br />

Muscles <strong>and</strong> Sensors. London: Springer-<br />

Verlag.<br />

Kim, K.J., Yim, W., Paquette, J.W., & Kim, D.<br />

2007. Ionic polymer-metal composites<br />

for underwater operation. J Intel Mat Syst<br />

Str. 18(2):123-31. doi: 10.1177/<br />

1045389X06063468.<br />

Krieg, M., & Mohseni, K. 2008. Thrust<br />

characterization of a bioinspired vortex ring<br />

thruster for locomotion of underwater<br />

robots. IEEE J Oceanic Eng. 33(2):123-32.<br />

doi: 10.1109/JOE.2008.920171.<br />

Krishen, K. 2009. Space applications for ionic<br />

polymer-metal composite sensors, actuators,<br />

<strong>and</strong> artificial muscles. Acta Astronaut.<br />

64(11-12):1160-6.<br />

Kruusamae, K., Brunetto, P., Graziani, S.,<br />

Punning, A., Di Pasquale, G., & Aabloo, A.<br />

2009. Self-sensing ionic polymer-metal<br />

composite actuating device with patterned<br />

surface electrodes. Polym Int. 59(3):300-4.<br />

doi: 10.1002/pi.2752.<br />

Lauder, G.V. 2007. How fish swim: Flexible<br />

fin thrusters as an EAP platform. In: Proceedings<br />

of the SPIE Smart Structures <strong>and</strong><br />

Materials: Electroactive Polymer Actuators<br />

<strong>and</strong> Devices (EAPAD). pp. 652402. San<br />

Diego, CA: SPIE.<br />

Lavu, B.C., Schoen, M.P., & Mahajan, A.<br />

2005. Adaptive intelligent control of ionic<br />

polymer metal composites. Smart Mater<br />

Struct. 14(4):466-74. doi: 10.1088/<br />

0964-1726/14/4/002.<br />

Lee, S.J., Han, M.J., Kim, S.J., Jho, J.Y., Lee,<br />

H.Y., & Kim, Y.H. 2006. A new fabrication<br />

method for ipmc actuators <strong>and</strong> application to<br />

artificial fingers. Smart Mater Struct. 15:<br />

1217-24. doi: 10.1088/0964-1726/15/5/008.<br />

Leo, D.J., Farinholt, K., & Wallmersperger, T.<br />

2005. Computational models of ionic<br />

transport <strong>and</strong> electromechanical transduction<br />

in ionomeric polymer transducers. In: Proceedings<br />

of the SPIE Smart Structures <strong>and</strong><br />

Materials: Electroactive Polymer Actuators<br />

<strong>and</strong> Devices (EAPAD). pp. 170-81.<br />

Menozzi, A., Leinhos, H.A., Beal, D.N., &<br />

B<strong>and</strong>yopadhyay, P.R. 2008. Open-loop control<br />

of a multifin biorobotic rigid underwater<br />

vehicle. IEEE J Oceanic Eng. 33(2):59-68.<br />

doi: 10.1109/JOE.2008.918687.<br />

Moheimani, S.O.R., & Fleming, A.J. 2006.<br />

Piezoelectric Transducers for Vibration Control<br />

<strong>and</strong> Damping (Advances in Industrial<br />

Control). London: Springer.<br />

Multiphysics, C. 2011. User’s Guide,<br />

version 4.1 a. COMSOL AB. Los Angeles, CA.<br />

Nam, J.D., Choi, H.R., Tak, Y.S., & Kim,<br />

K.J. 2003. Novel electroactive, silicate<br />

nanocomposites prepared to be used as<br />

actuators <strong>and</strong> artificial muscles. Sensor<br />

Actuat A-Phys. 105:83-90.<br />

Nemat-Nasser, S., & Jiang, Y.L. 2000.<br />

Electromechanical response of ionic<br />

polymer-metal composites. J Appl Phys.<br />

87(7):3321-31. doi: 10.1063/1.372343.<br />

Nguyen, V.K., Lee, J.W., & Yoo, Y. 2007.<br />

Characteristics <strong>and</strong> performance of ionic<br />

polymer-metal composite actuators based<br />

on nafion/layered silicate <strong>and</strong> nafion/silica<br />

nanocomposites. Sensor Actuat B. 120:<br />

529-37. doi: 10.1016/j.snb.2006.03.015.<br />

Pak, J.J., Kim, J., Oh, S.W., Son, J.H.,<br />

Cho, S.H., Lee, S.-K., … Kim, B. 2004.<br />

Fabrication of ionic-polymer-metal-composite<br />

(IPMC) micropump using a commercial<br />

nafion. In: Proceedings of the SPIE Smart<br />

Structures <strong>and</strong> Materials: Electroactive Polymer<br />

Actuators <strong>and</strong> Devices (EAPAD). pp. 272-80.<br />

San Diego, CA: SPIE.<br />

Porfiri, M. 2009. Influence of electrode<br />

surface roughness <strong>and</strong> steric effects on the<br />

nonlinear electromechanical behavior of<br />

ionic polymer metal composites. Phys Rev E.<br />

79(4):041503. doi: 10.1103/<br />

PhysRevE.79.041503.<br />

Pugal, D., Jung, K., Aabloo, A., & Kim,<br />

K.J. 2010a. Ionic polymer metal composite<br />

mechanoelectrical transduction: Review <strong>and</strong><br />

perspectives. Polym Int. 59(3):279-89.<br />

doi: 10.1002/pi.2759.<br />

Pugal, D., Kim, K.J., & Aabloo, A. 2010b.<br />

Modeling the transduction of IPMC in<br />

3D configurations. In: Proceedings of the<br />

SPIE Smart Structures <strong>and</strong> Materials <strong>and</strong><br />

Nondestructive Evaluation <strong>and</strong> Health<br />

Monitoring. pp. 76441T. San Diego, CA:<br />

SPIE.<br />

Pugal, D., Kim, K.J., Punning, A., Kasemagi,<br />

H., Kruusmaa, M., & Aabloo, A. 2008.<br />

A self-oscillating ionic polymer-metal composite<br />

bending actuator. J Appl Phys.<br />

103(8):084908-6. doi: 10.1063/1.2903478.<br />

Ramo, S. 1939. Currents induced by<br />

electron motion. Proc IRE. 27(9):584-5.<br />

doi: 10.1109/JRPROC.1939.228757.<br />

Richardson, R.C., Levesley, M.C., Brown,<br />

M.D., Hawkes, J.A., Watterson, K., &<br />

Walker, P.G. 2003. Control of ionic polymer<br />

metal composites. IEEE/ASME T Mech.<br />

8(2):245-3.<br />

July/August 2011 Volume 45 Number 4 97


Riddle, R.O., Jung, Y., Kim, S.-M., Song, S.,<br />

Stolpman, B., Kim, K.J., & Leang, K.K.<br />

2010. Sectored-electrode IPMC actuator for<br />

bending <strong>and</strong> twisting motion. In: Proceedings<br />

of the SPIE Smart Structures <strong>and</strong> Materials<br />

<strong>and</strong> Nondestructive Evaluation <strong>and</strong> Health<br />

Monitoring. San Diego, CA: SPIE.<br />

Shahinpoor, M., Bar-Cohen, Y., Simpson,<br />

J.O., & Smith, J. 1998. Ionic polymer-metal<br />

composites (IPMCs) as biomimetic sensors,<br />

actuators <strong>and</strong> artificial muscles—a review.<br />

Smart Mater Struct. 7(6):R15-30.<br />

doi: 10.1088/0964-1726/7/6/001.<br />

Shahinpoor, M., & Kim, K.J. 2001. Ionic<br />

polymer-metal composites: I. Fundamentals.<br />

Smart Mater Struct. 10:819-33.<br />

doi: 10.1088/0964-1726/10/4/327.<br />

Shahinpoor, M., & Kim, K.J. 2005. Ionic<br />

polymer-metal composites: IV. Industrial <strong>and</strong><br />

medical applications. Smart Mater Struct.<br />

14(1):197-214. doi: 10.1088/0964-<br />

1726/14/1/020.<br />

Shan, Y., & Leang, K.K. 2009. Frequencyweighted<br />

feedforward control for dynamic<br />

compensation in ionic polymer-metal<br />

composite actuators. Smart Mater Struct.<br />

18(12):125016. doi: 10.1088/0964-1726/18/<br />

12/125016.<br />

unmanned underwater vehicles. IEEE<br />

J Oceanic Eng. 32(3):533-50.<br />

doi: 10.1109/JOE.2007.903362.<br />

Tiwari, R., Kim, K.J., & Kim, S. 2008. Ionic<br />

polymer-metal composite as energy harvesters.<br />

Smart Struct Syst. 4(5):549-63.<br />

Tortora, G., Caccavaro, S., Valdastri, P.,<br />

Menciassi, A., & Dario, P. 2010. Design of an<br />

autonomous swimming miniature robot based<br />

on a novel concept of magnetic actuation.<br />

In: IEEE International Conference on<br />

Robotics <strong>and</strong> Automation. pp. 1592-7.<br />

Piscataway, NJ: IEEE. doi: 10.1109/<br />

ROBOT.2010.5509599.<br />

Yim, W., Lee, J., & Kim, K.J. 2007. An<br />

artificial muscle actuator for biomimetic<br />

underwater propulsors. Bioinspir Biomim.<br />

2:S31-41. doi: 10.1088/1748-3182/2/2/S04.<br />

Yoon, W.J., Reinhall, P.G., & Seibel, E.J.<br />

2007. Analysis of electro-active polymer<br />

bending: A component in a low cost<br />

ultrathin scanning endoscope. Sensor Actuat<br />

A-Phys. 133(2):506-17. doi: 10.1016/<br />

j.sna.2006.04.037.<br />

Shockley, W. 1938. Currents to conductors<br />

induced by a moving point charge. J Appl<br />

Phys. 9(10):635-6. doi: 10.1063/1.1710367.<br />

Smela, E., Kallenbach, M., & Holdenried, J.<br />

1999. Electrochemically driven polypyrrole<br />

bilayers for moving <strong>and</strong> positioning bulk<br />

micromachined silicon plates. IEEE/ASME<br />

T Microelectromech Syst. 8(4):373-83. doi:<br />

10.1109/84.809051.<br />

Tadokoro, S., Yamagami, S., & Takamori, T.<br />

2000. An actuator model of ICPF for robotic<br />

applications on the basis of physicochemical<br />

hypotheses. In: IEEE International Conference<br />

on Robotics <strong>and</strong> Automation. pp. 1340-6.<br />

Piscataway, NJ: IEEE.<br />

Tangorra, J.L., Davidson, S.N., Hunter,<br />

I.W., Madden, P.G.A., Lauder, G.V., Dong,<br />

H., … Mittal, R. 2007. The development<br />

of a biologically inspired propulsor for<br />

98 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Batoid Fishes: Inspiration for the Next<br />

Generation of Underwater Robots<br />

AUTHORS<br />

Keith W. Moored<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering Department,<br />

Princeton University<br />

Frank E. Fish<br />

Department of Biology,<br />

University of West Chester<br />

Trevor H. Kemp<br />

Hilary Bart-Smith<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering Department,<br />

University of Virginia<br />

Introduction<br />

ABSTRACT<br />

For millions of years, aquatic species have utilized the principles of unsteady<br />

hydrodynamics for propulsion <strong>and</strong> maneuvering. They have evolved high-endurance<br />

swimming that can outperform current underwater vehicle technology in the areas<br />

of stealth, maneuverability <strong>and</strong> control authority. Batoid fishes, including the manta<br />

ray, Manta birostris, the cownose ray, Rhinoptera bonasus, <strong>and</strong> the Atlantic stingray,<br />

Dasyatis sabina, have been identified as a high-performing species due to their<br />

ability to migrate long distances, maneuver in spaces the size of their tip-to-tip<br />

wing span, produce enough thrust to leap out of the water, populate many underwater<br />

regions, <strong>and</strong> attain sustained swimming speeds of 2.8 m/s with low flapping/<br />

undulating frequencies. These characteristics make batoid fishes an ideal platform<br />

to emulate in the design of a bio-inspired autonomous underwater vehicle. The<br />

enlarged pectoral fins of each ray undergoes complex motions that couple spanwise<br />

curvature with a chordwise traveling wave to produce thrust <strong>and</strong> to maneuver. Researchers<br />

are investigating these amazing species to underst<strong>and</strong> the biological principles<br />

for locomotion. The continuum of swimming motions—from undulatory to<br />

oscillatory—demonstrates the range of capabilities, environments, <strong>and</strong> behaviors<br />

exhibited by these fishes. Direct comparisons between observed swimming motions<br />

<strong>and</strong> the underlying cartilage structure of the pectoral fin have been made. A<br />

simple yet powerful analytical model to describe the swimming motions of batoid<br />

fishes has been developed <strong>and</strong> is being used to quantify their hydrodynamic performance.<br />

This model is also being used as the design target for artificial pectoral<br />

fin design. Various strategies have been employed to replicate pectoral fin motion.<br />

Active tensegrity structures, electro-active polymers, <strong>and</strong> fluid muscles are three<br />

structure/actuator approaches that have successfully demonstrated pectoral-finlike<br />

motions. This paper explores these recent studies to underst<strong>and</strong> the relationship<br />

between form <strong>and</strong> swimming function of batoid fishes <strong>and</strong> describes attempts<br />

to emulate their abilities in the next generation of bio-inspired underwater vehicles.<br />

Keywords: biomimicry, bioinspired, autonomous underwater vehicle, manta ray,<br />

tensegrity structures<br />

There has been an explosion of activity<br />

in the area of biomimicry <strong>and</strong><br />

bioinspired engineering research. Biomimicry<br />

directly emulates the form<br />

<strong>and</strong> function of species to illuminate<br />

the physical principles behind nature’s<br />

designs. Bioinspired engineering takes<br />

advantage of the knowledge gained<br />

through biomimetic studies to judiciously<br />

apply novel physical principles<br />

to develop solutions with added functionality<br />

over conventional engineering<br />

approaches. Biology, biomimetics<br />

<strong>and</strong> bioinspired engineering are intimately<br />

linked. Thus it comes as no surprise<br />

that biologists <strong>and</strong> engineers are<br />

collaborating by developing biorobotic<br />

devices to: (1) elucidate key insights<br />

into biological form <strong>and</strong> function <strong>and</strong><br />

(2) develop bioinspired autonomous<br />

underwater vehicles (BAUVs) to improve<br />

functionality of autonomous<br />

underwater vehicles (AUVs).<br />

Aquatic species outperform conventional<br />

AUVs in the areas of maneuverability<br />

<strong>and</strong> control authority<br />

(B<strong>and</strong>yopadhyay, 2005), while having<br />

a low-noise signature that blends into<br />

the background <strong>and</strong> high swimming<br />

efficiencies. Batoid rays excel in all<br />

of these areas, giving them an abundance<br />

of recent attention. The focus<br />

of this paper is to present the growing<br />

body of work being done to underst<strong>and</strong><br />

<strong>and</strong> quantify the swimming performance<br />

of batoid fishes (i.e., skates,<br />

sting rays, manta rays) <strong>and</strong> the stateof-the-art<br />

in robotic mimicry. Of particular<br />

interest are the mechanisms<br />

associated with the swimming of<br />

these fishes, which employ flattened<br />

pectoral fins to propel <strong>and</strong> maneuver<br />

in the ocean <strong>and</strong> in rivers.<br />

Underst<strong>and</strong>ing Biological Foundation<br />

will discuss our current biological<br />

July/August 2011 Volume 45 Number 4 99


underst<strong>and</strong>ing of batoid rays. Rationale<br />

for Mimicking Batoid Rays will<br />

present compelling reasons for scientists<br />

<strong>and</strong> engineers to study batoids<br />

rays, highlighting their swimming<br />

characteristics that would be desired<br />

in an underwater vehicle. Bioinspired<br />

Robotics delves into the exp<strong>and</strong>ing<br />

worldofbio-inspiredunderwatervehicles,<br />

with particular emphasis on<br />

ray-like platforms. Concluding Comments<br />

<strong>and</strong> Future Directions concludes<br />

with a discussion on critical areas that<br />

need to be addressed in order for the<br />

next generation of underwater vehicles<br />

to be truly bioinspired.<br />

Underst<strong>and</strong>ing Biological<br />

Foundation<br />

Fish swim by imparting momentum<br />

to water from the movements<br />

of a variety of propulsors, which can<br />

include the body, median fins, <strong>and</strong><br />

paired fins (Sfakiotakis et al., 1999).<br />

Although primitive batoid fishes use<br />

the body <strong>and</strong> caudal fin toswim,<br />

more advanced batoids have become<br />

specialized to swim with enlarged pectoral<br />

fins. It is emphasized that pectoral<br />

fin locomotion can have significant<br />

advantages in maneuvering <strong>and</strong> stationkeeping<br />

(Sfakiotakis et al., 1999). In<br />

recent years, more attention is being<br />

paid to pectoral fin hydrodynamics as<br />

their importance is being realized in not<br />

only steady-state locomotion but also in<br />

transient maneuvers (B<strong>and</strong>yopadhyay,<br />

FIGURE 1<br />

2005; Lauder et al., 2002; Combes &<br />

Daniel, 2001; Palmisano et al., 2007).<br />

Batoid rays take pectoral fin locomotion<br />

to an evolutionary extreme<br />

(Figure 1). Rays have a dorso-ventrally<br />

flattened body with enlarged pectoral<br />

fins that are seamlessly merged with<br />

their body to form a biological blended<br />

wing-body configuration. Propulsive<br />

waves are passed through the fins by<br />

serial contraction of the appendicular<br />

musculature. The waves have their<br />

greatest amplitude toward the periphery<br />

of the fin.<br />

Even though among rays there is<br />

similar morphology, their locomotor<br />

strategies can be very different. Undulatory<br />

motion, defined as having<br />

greater than one or more waves present<br />

on a fin (Rosenberger, 2001), is one<br />

extreme of kinematic motion <strong>and</strong> was<br />

termed ‘rajiform’ by Breder (1926).<br />

These fishes swim just over the ocean<br />

floor. The other extreme is oscillatory<br />

motion, defined as having less than<br />

half of a wave present (flapping) on a<br />

fin, <strong>and</strong> was coined ‘mobuliform’ by<br />

Webb (1994). The mobuliform swimming<br />

mode appears as a wing-like flapping<br />

motion <strong>and</strong> is associated with rays<br />

that have a more pelagic existence. The<br />

various species of batoids exhibit a<br />

continuum of kinematic motions between<br />

the two extremes of undulation<br />

<strong>and</strong> oscillation (Rosenberger, 2001).<br />

Myliobatoids, i.e., the mid-water<br />

rays, including manta, eagle, bat, <strong>and</strong><br />

(a) Image of a manta ray. (b) Image of a cownose ray. Both rays are part of the batoid family <strong>and</strong><br />

swim via an oscillatory motion.<br />

cownose rays, nearly exclusively utilize<br />

oscillatory motion (Klausewitz, 1964;<br />

Sasko et al., 2006; Heine, 1992).<br />

Some research has been done to characterize<br />

the biology <strong>and</strong> behavior of<br />

myliobatoids (Schaefer & Summers,<br />

2005; Summers, 2000). Heine (1992)<br />

studied the kinematics of the cownose<br />

ray by videotaping live rays swimming<br />

in a flow tank. Rosenberger (2001)<br />

compared the kinematics of many<br />

batoid rays spanning the undulationoscillation<br />

continuum <strong>and</strong> suggested<br />

that oscillatory rays have evolved to<br />

have efficient locomotion. Klausewitz<br />

(1964) describes the kinematic motions<br />

of the manta ray while Moored<br />

et al. (Moored, 2010; Moored et al.,<br />

2011b) developed a simple yet powerful<br />

analytical model to quantify the kinematics<br />

of different species of batoid<br />

rays (such as the manta ray, Atlantic<br />

stingray, <strong>and</strong> the cownose ray). This<br />

model is used as a target deformation<br />

field for a bio-inspired fin (Moored<br />

et al., 2011b) <strong>and</strong> to calculate the<br />

swimming performance of different<br />

batoid ray species (Moored et al.,<br />

2011b; Pederzani et al., 2011).<br />

Morphometrics<br />

The greatly enlarged pectoral fins<br />

form wide lateral extensions of the<br />

body that range in morphology from<br />

a circular disc to triangular, wing-like<br />

planforms. Species of batoids show<br />

overa90-foldrangeinsizewiththe<br />

largest being the manta (Manta birostris).<br />

Rays that swim by undulations<br />

of the fin intherajiformmodehave<br />

fin shapes with relatively low aspect<br />

ratios (the ratio of span to chord). Oscillatory<br />

swimmers, using the mobuliform<br />

mode, possess higher aspect ratio<br />

fins with longer spans.<br />

The cross-sectional geometry of<br />

batoid rays has a streamlined appearance.<br />

Rajiform swimmers have a<br />

100 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


ody <strong>and</strong> pectoral fins with a flattened<br />

ventral side <strong>and</strong> low vaulted dorsum,<br />

giving a design similar to a cambered<br />

wing. Although the central portion of<br />

the body shows a slight asymmetry<br />

with a flattened ventral surface <strong>and</strong><br />

convex dorsal surface, the pectoral<br />

fins of mobuliform swimmers display<br />

symmetrical cross-sectional profiles<br />

reminiscent of engineered foils<br />

(Abbott & von Doenhoff, 1949).<br />

The internal skeleton of the pectoral<br />

fins of batoids are composed of numerous<br />

short, cylindrical cartilaginous<br />

elements (Heine, 1992; Schaefer &<br />

Summers, 2005). These cartilaginous<br />

elements are the supportive radials of<br />

the fin. The radials are stacked end to<br />

end. The radial cartilages are mineralized<br />

to varying degree depending on<br />

the species of batoid, where the mineralization<br />

is found on the exterior<br />

of the cartilaginous element with the<br />

core being unmineralized (Schaefer &<br />

Summers, 2005). Rajiform swimming<br />

rays display joint staggering with little<br />

calcification of the joints, whereas<br />

the skeleton of oscillatory swimmers<br />

shows cross-bracing <strong>and</strong> calcification.<br />

The skeleton is moved by long thin<br />

muscles that run from the exp<strong>and</strong>ed<br />

pectoral girdle along each fin rayto<br />

every radial. The range of motion of<br />

the articulated radials is small (∼15°),<br />

but the large number of components<br />

in the pectoral fins permits sufficient<br />

spanwise <strong>and</strong> chordwise flexibility<br />

for propulsion <strong>and</strong> maneuvering<br />

(Rosenberger, 2001; Klausewitz,<br />

1964; Heine, 1992; Schaefer &<br />

Summers, 2005).<br />

Rationale for Mimicking<br />

Batoid Rays<br />

With respect to pursuing bioinspired<br />

engineering, a key question<br />

to be answered is to explain/justify<br />

why a particular species is a good c<strong>and</strong>idate<br />

to emulate. These reasons can<br />

be very diverse <strong>and</strong> are motivated by<br />

the particular application envisioned.<br />

In the case of AUVs, compelling reasons<br />

to consider biology as a starting<br />

point for the development of the next<br />

generation vehicle are (1) a stealthy<br />

signature, (2) high efficiency <strong>and</strong> economy,<br />

(3) exp<strong>and</strong>ed working environment,<br />

<strong>and</strong> (4) scalability/payload<br />

capacity. Additionally, a key justification<br />

for this approach is that there are<br />

tangible improvements that can be<br />

made over current AUV technologies.<br />

The pool of species to emulate is vast.<br />

However, recent studies of batoid rays<br />

have demonstrated significant swimming<br />

abilities that would be desirable<br />

in an underwater vehicle.<br />

Stealth means either quiet operation<br />

or the ability to blend into the<br />

background noise. A biomimetic approach<br />

naturally fulfills these requirements<br />

by creating a vehicle whose<br />

minimal noise signature blends in<br />

with the environment. The noise signature<br />

of a fish is very different to<br />

that of a propeller (B<strong>and</strong>yopadhyay,<br />

2005). Even biomimetic sensor arrays<br />

such as an artificial lateral line (Yang<br />

et al., 2006) or artificial seal whiskers<br />

(Stocking et al., 2010) that are sensitive<br />

to hydrodynamic wake signatures<br />

would presumably delineate a flapping<br />

fin wake as an animal <strong>and</strong> a propeller<br />

wake as a man-made device. Given recent<br />

advances in underwater imaging<br />

technology including LIDAR systems<br />

(Jaffe et al., 2001), synthetic aperture<br />

sonar (Kocak & Caimi, 2005) <strong>and</strong><br />

biomimetic sonar systems (Dobbins,<br />

2007), the shape <strong>and</strong> movements of<br />

an AUV are becoming increasingly important.<br />

By mimicking the body form<br />

of an aquatic species, the identification<br />

of such a stealthy vehicle as a manmade<br />

device becomes difficult.<br />

Batoid rays offer an intriguing design<br />

solution for a high-endurance vehicle.<br />

Pelagic rays, such as the manta<br />

ray or cownose ray, migrate thous<strong>and</strong>s<br />

of miles a year. This suggests that these<br />

species have evolved to become highendurance<br />

swimmers. As discussed<br />

previously, myliobatoid rays have a<br />

dorso-ventrally flattened body with<br />

enlarged pectoral fins, forming a natural<br />

gliding morphology. In terms of<br />

avehicle,abatoid-inspiredUVisan<br />

advance on current underwater glider<br />

technology. This bioinspired platform<br />

wouldenableavehicletohavehighendurance<br />

capabilities like current underwater<br />

gliders, such as the Slocum<br />

AUV (Webb & Simonetti, 1999). Additionally,<br />

this system has the potential<br />

to transition to a faster, more maneuverable<br />

vehicle that can operate in dynamic<br />

environments such as the<br />

littoral zone or areas with large currents<br />

<strong>and</strong> high wave action. Observations<br />

of various rays show them to be<br />

highly maneuverable <strong>and</strong> adaptable<br />

to local conditions—for example, to<br />

station keep <strong>and</strong> even swim backwards.<br />

Their ability to control their<br />

stability via the pectoral fins, especially<br />

when compensating for challenging<br />

environments, must also be considered<br />

as a desirable characteristic to emulate<br />

in an underwater vehicle.<br />

Scalability is an attractive feature<br />

in any artificial system. With respect<br />

to bioinspired underwater vehicles, batoids<br />

display an extraordinary range of<br />

dimensions, growing in excess of<br />

9 m tip-to-tip in the case of manta<br />

rays.Thus,thesize<strong>and</strong>speedthat<br />

batoids perform at are equivalent to<br />

the operation range of marine vehicles.<br />

The size of the vehicle will very much<br />

depend on the mission requirements,<br />

but by using the batoid as the foundation,<br />

it is feasible to produce a variety<br />

of sized vehicles that can explore <strong>and</strong><br />

July/August 2011 Volume 45 Number 4 101


traverse a wide range of ocean space<br />

while performing a wide range of tasks.<br />

Moreover, a batoid-inspired vehicle<br />

would have a large planform surface<br />

area, making this platform an excellent<br />

c<strong>and</strong>idate for flexible solar cells to extend<br />

its range (Dennler et al., 2008),<br />

similar to the solar powered SAUV II<br />

vehicle (Jalbert et al., 2003). In addition,<br />

the rigid body of batoids permits<br />

space for control systems, sensory<br />

devices <strong>and</strong> increased payload.<br />

Bioinspired Robotics<br />

There has been growing use of<br />

AUVs in recent years with over<br />

240 different AUV platforms developed<br />

<strong>and</strong> used in the field (B<strong>and</strong>yopadhyay,<br />

2005). These AUVs typically are built<br />

for reconnaissance/surveying <strong>and</strong> were<br />

originally designed for endurance<br />

(Blidberg, 2001). This gave rise to<br />

the design of underwater gliders<br />

using conventional design principles<br />

(i.e., steady-state hydrodynamics)<br />

that have high endurance but little maneuverability<br />

(Webb & Simonetti,<br />

1999). From another perspective,<br />

biology has created thous<strong>and</strong>s of swimming<br />

platforms that can outmaneuver<br />

the best AUVs while still<br />

having highly efficient, high-speed<br />

<strong>and</strong> high-endurance performance.<br />

Moreover, many of these biological<br />

systems can also hover in place with no<br />

forward locomotion, generate large<br />

enough forces to hold station under<br />

adverse environmental conditions,<br />

burst with incredible acceleration <strong>and</strong><br />

have a significantly reduced noise signature<br />

compared to man-made AUVs<br />

(Fish & Lauder, 2006). In an attempt<br />

to bridge the performance gap between<br />

conventional AUVs <strong>and</strong> biological<br />

systems, engineers have been<br />

shifting focus to BAUVs, which is a<br />

highly multi-disciplinary research area<br />

(B<strong>and</strong>yopadhyay, 2005; Colgate &<br />

Lynch, 2004). To reach some of these<br />

goals, there is a spectrum of first generation<br />

BAUVs that have been developed.<br />

Some form an exotic collection<br />

mimicking lamprey (Ayers et al.,<br />

2000; Crespi et al., 2004), tuna<br />

(Barrett et al., 1996; Yu et al., 2004;<br />

Anderson & Chhabra, 2002), <strong>and</strong> dolphins<br />

(Yu et al., 2007), while others<br />

aremoreconventionalstyleAUV<br />

designs outfitted with bioinspired flapping<br />

propulsors (Fish et al., 2003; Low<br />

&Willy,2006;Listaketal.,2005;<br />

Borgen et al., 2003; Mojarrad, 2000;<br />

Licht et al., 2004). These different<br />

BAUV designs were made possible<br />

partly from advances in our underst<strong>and</strong>ing<br />

of unsteady hydrodynamics<br />

<strong>and</strong> the biology of nektonic (swimming)<br />

organisms. However, this BAUV<br />

technology still has a long way to go<br />

before the performance gap is bridged.<br />

Recently, researchers have turned<br />

to pectoral fin locomotion for inspiration.<br />

Pectoral fin motions utilized by<br />

sunfish, perch, bass <strong>and</strong> bird wrasse<br />

for low-speed swimming <strong>and</strong> maneuvering<br />

have been studied (Gibb et al.,<br />

1994; Drucker & Jensen, 1997;<br />

Lauder & Jayne, 1996; Walker &<br />

Westneat, 1997). To underst<strong>and</strong> the<br />

forces <strong>and</strong> moments produced by<br />

these pectoral fins, biorobotic solutions<br />

began with paddle-like fins that<br />

mimic the bulk kinematics of labriform<br />

swimming (Kato, 1998; Kato &<br />

FIGURE 2<br />

A biomimetic sunfish pectoral fin (Tangorra et al., 2008a, 2008b).<br />

Furushima, 2002). In recent years, devices<br />

have been constructed to more<br />

accurately replicate the kinematics<br />

utilized by the fish with the advent of<br />

actively flexible fins that can produce<br />

chordwise undulations as well as<br />

spanwise curvature (Yan et al., 2010;<br />

Palmisano et al., 2008; Kato et al.,<br />

2008; Tangorra et al., 2007). Actively<br />

flexible fins deform due to the presence<br />

of actuators instead of undergoing<br />

rigid body motions, like the heave<br />

<strong>and</strong> pitch of oscillating airfoils. The<br />

motion of actively flexible fins is fully<br />

prescribed. In contrast, passively flexible<br />

fins deform under fluid loading<br />

such that their motion is not fully prescribed,<br />

but a function of the forces applied<br />

to the fin. Tangorra et al. (2008a,<br />

2008b) advanced their artificial pectoral<br />

fin (Figure 2) by not only matching<br />

the kinematics of the sunfish but also<br />

replicating the internal fin structure<br />

<strong>and</strong> material properties, which allowed<br />

the fin to have a greater degree of passive<br />

flexibility. This fin wasusedto<br />

fully characterize how sunfish produce<br />

<strong>and</strong> manipulate fluid forces to propel<br />

themselves <strong>and</strong> maneuver. With<br />

equivalent passive flexibility as the<br />

fins of the sunfish, the artificial fin<br />

was able to produce thrust on both<br />

the outstroke <strong>and</strong> instroke of its fin<br />

beat, as observed of the animal.<br />

An excellent example demonstrating<br />

the link between biology, biomimetics,<br />

<strong>and</strong> bioinspired engineering<br />

102 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


is in the development of an artificial<br />

ghost knifefish (Curet et al., 2010).<br />

Through observation of biology, a<br />

biorobotic device was developed <strong>and</strong><br />

used to underst<strong>and</strong> the locomotion<br />

strategies of this fish. Particle image<br />

velocimetry, in conjunction with computational<br />

fluid dynamics, were employed<br />

to explore the propulsive <strong>and</strong><br />

station-keeping characteristics of this<br />

fascinating fish.<br />

Batoid-Inspired Devices<br />

There have also been attempts by<br />

researchers to develop batoid-inspired<br />

fins <strong>and</strong> AUVs. These devices mimic<br />

both the undulatory swimming seen<br />

inbenthicrays(similartothelocomotion<br />

of the ghost knifefish) as well<br />

as the oscillatory swimming seen in pelagic<br />

rays, such as the manta. Many of<br />

these batoid-inspired devices are used<br />

as a platform for exploring actuation<br />

technologies.<br />

Motors <strong>and</strong> servomotors are used<br />

in ray-like devices due to their simple<br />

controllability, high-speed operation<br />

<strong>and</strong> repeatability. Some motor-driven<br />

devices mimic undulatory rays (Low<br />

& Willy, 2006; V. y Alvarado et al.,<br />

2010), while others mimic oscillatory<br />

rays (Yang et al., 2009; Zhou &<br />

Low, 2010; Gao et al., 2007). Researchers<br />

have developed oscillatory<br />

ray-like vehicles based on pneumatic<br />

pectoral fins (Brower, 2006; Cai<br />

et al., 2010; Suzumori et al., 2007).<br />

Sfakiotakis et al. (2005) also used<br />

pneumatically driven “fin rays” to produce<br />

an undulatory ray-like device.<br />

Festo (2008) has built a BAUV called<br />

AquaRay, utilizing fluidic muscles.<br />

This robot uses an oscillatory flapping<br />

motion to swim, but no quantitative<br />

data on the performance is given.<br />

Takagi et al. (2007) utilized ionic<br />

polymer-metal composites (IPMCs)<br />

actuators to develop a stingray-like<br />

device that could achieve a swimming<br />

speed of 0.24 BL/s. Chen et al. (2011)<br />

developed a novel fabrication method<br />

to produce IPMCs that can deform<br />

with complex three-dimensional kinematics.<br />

This fabrication technology<br />

was used to produce a manta ray-like<br />

device. Shape memory alloys have<br />

also been employed in the design of artificial<br />

pectoral fins (Yong-hua et al.,<br />

2007; Wang et al., 2008). Wang<br />

et al. (2008) presented a robotic squid<br />

utilizing a rajiform mode of swimming<br />

to achieve 0.24 BL/s swimming<br />

speed. The best swimming speed performance<br />

of these actuator platforms<br />

was 1.4 BL/s achieved by the servomotor<br />

driven devices; however, the<br />

associated power cost is not given.<br />

These studies showcase the plethora<br />

of actuator technologies that can<br />

be utilized to produce deformations<br />

similar to that of rays. One concern<br />

with this approach is that the actuator<br />

choice is directly coupled with the fin<br />

technology. An alternative approach is<br />

to start with a fin design that is actuator<br />

independent <strong>and</strong> so the choice of<br />

actuator is dependent on the application<br />

of the vehicle. For instance, if actuator<br />

efficiency is not a concern but<br />

noiseless operation is of prime importance,<br />

an SMA actuator could be chosen.<br />

Also, this approach opens up the<br />

possibility of replacing current actuators<br />

with new technologies that may<br />

be superior. Solutions like this are discussed<br />

next.<br />

In a study to increase our underst<strong>and</strong>ing<br />

of the hydrodynamics of<br />

batoid locomotion, Clark <strong>and</strong> Smits<br />

(2006) designed <strong>and</strong> built an active<br />

artificial oscillating fin that was independent<br />

of the actuator. They quantified<br />

the performance of the fin by<br />

measuring the efficiency <strong>and</strong> thrust<br />

production, as well as determining an<br />

optimal traveling wave wavelength.<br />

Furthermore, by using dye flow visualization,<br />

they characterized the wake<br />

structure as a series of interacting trailing<br />

edge vortices forming a threedimensional<br />

reverse von Kármán<br />

vortex street. In free swimming tests<br />

(Moored et al., 2011a), a swimming<br />

speed of 2 BL/s <strong>and</strong> a swimming economy,<br />

ζ (ζ = U = P f ,whereU is the<br />

swimming speed <strong>and</strong> P f is the power<br />

consumption), of 0:132 BL/J was<br />

reached for an actively flexible single<br />

fin. When some passive flexibility<br />

was introduced, the swimming speed<br />

dropped to 1:7 BL/s while the economy<br />

rose to 0:18 BL/J at the same flapping<br />

frequency of 2 Hz. This work<br />

has also highlighted the prime importance<br />

of the traveling wave in ray-like<br />

propulsion.<br />

Another actuator-independent approach<br />

has been developed using active<br />

tensegrity structures. Tensegrity structures<br />

are truss-like structures where<br />

some of the rigid elements have been<br />

replaced by cable elements (Figure 3).<br />

The cable elements must be in a<br />

state of tension for the structure to<br />

have integrity, giving rise to the contraction<br />

of “tensional-integrity” to<br />

tensegrity. Tensegrity structures act<br />

as a “skeleton-tendon” foundation<br />

that can use any actuator type to support<br />

the generation of large loads,<br />

match the kinematics of batoid rays,<br />

<strong>and</strong> perform with minimal actuation<br />

energy (Moored et al., 2011b; Moored<br />

& Bart-Smith, 2009).<br />

Various tensegrity actuation strategies<br />

are explored that are capable of<br />

matching the key kinematic features<br />

of batoid-propulsion: a chordwise<br />

traveling wave coupled with a large<br />

amplitude curved spanwise deformation.<br />

The strategies involve either<br />

embedding the actuators into the tensegrity<br />

structure (embedded actuation)<br />

or migrating the actuators outside of<br />

July/August 2011 Volume 45 Number 4 103


FIGURE 3<br />

(a) Three-dimensional tensegrity structures (three, four, <strong>and</strong> six strut prismatic structures). (b) Tensegrity-based fin concept. The fin can deform<br />

with coupled curved spanwise motion <strong>and</strong> chordwise undulation to mimic the kinematics of the manta ray <strong>and</strong> the batoid family in general. The<br />

tensegrity deforms when active elements contract or exp<strong>and</strong>.<br />

the structure (remote actuation). With<br />

respect to embedded actuation, optimal<br />

solutions have been calculated<br />

that give the location <strong>and</strong> actuation<br />

strain necessary to match a target displacement<br />

field (Moored & Bart-<br />

Smith, 2007). However, embedded<br />

actuation is problematic, as it requires<br />

many actuators to match the complex<br />

ray kinematics, adds mass to the<br />

active structure (thereby requiring<br />

more power to flap) <strong>and</strong> limits the<br />

scalability of solutions to the size of<br />

the actuator. Remote actuation overcomes<br />

these limitations by placing<br />

the actuators outside of the active region<br />

<strong>and</strong> connecting to the structure<br />

via a routed cable. A general numerical<br />

model––applicable to any topology<br />

<strong>and</strong> any actuation strategy––has<br />

been derived (Moored & Bart-Smith,<br />

2009).<br />

Moored et al. (2011) derive analytical<br />

solutions for active planar tensegrity<br />

beam structures. These solutions coupled<br />

with the numerical solution are<br />

utilized to identify optimal stiffnessto-mass<br />

<strong>and</strong> strength-to-mass strategies.<br />

Structural performance metrics<br />

were calculated showing that the fin<br />

structure can closely match the kinematics<br />

of the manta ray, under external<br />

loading, using open-loop actuation of<br />

four actuators remotely located outside<br />

of the active structure (Moored et al.,<br />

2011b). In an attempt to simplify the<br />

experimental design of an artificial<br />

fin, actuated via remote actuation, a<br />

single tensegrity beam was built<br />

<strong>and</strong>placedwithinanelastomerfin.<br />

Figure 4a shows images of a single<br />

tensegrity beam as it is actuated. The<br />

beam enables leading-edge actuation<br />

of the artificial fin (Figure4b).This<br />

fin was then tested in a flow tank to<br />

observe the influence of frequency on<br />

the wake topology. Figures 4c <strong>and</strong> 4d<br />

show the actuating fin inwaterfrom<br />

the side <strong>and</strong> below. The black lines<br />

superimposed on these images represent<br />

the kinematical model for the kinematics<br />

of a cownose ray—note the<br />

excellent agreement between experiment<br />

<strong>and</strong> theory, especially with the<br />

passive response of the elastomer fin.<br />

This approach costs minimal power<br />

consumption <strong>and</strong> shows the simple design<br />

of a high-performance tensegritybased<br />

artificial pectoral fin.<br />

Concluding Comments<br />

<strong>and</strong> Future Directions<br />

The idea to look to nature for inspiration<br />

is not new, <strong>and</strong> this rich arena<br />

104 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 4<br />

Example of a tensegrity-based actuating fin. (a) Photos of a tensegrity beams employing remote<br />

actuation. (b) Dorsal view of elastomer fin with leading edge actuation via tensegrity beam.<br />

(c) Posterior view of actuating fin shown in (b). (d) Lateral view of actuating fin. Note the lines<br />

in (c) <strong>and</strong> (d) represent the mathematical model derived to describe kinematic motions of a manta<br />

ray pectoral fin.<br />

continues to be a source for engineers<br />

to aid in solving challenging problems.<br />

From Leonardo Da Vinci’s flying machine,<br />

Helical Air Screw, to leading<br />

edge whale-like tubercles on wind turbine<br />

blades to improve efficiencies<br />

(Fish et al., 2011). The opportunities<br />

to learn from nature <strong>and</strong> emulate its<br />

unique approach to overcoming challenges<br />

seem endless. In this paper, we<br />

have touched upon the challenge of<br />

creating efficient, economic, <strong>and</strong> maneuverable<br />

underwater swimming<br />

platforms. We have focused on batoid<br />

fishes for inspiration in the design of<br />

the next generation of bioinspired underwater<br />

vehicles, as emulation of its<br />

swimming characteristics—efficiency,<br />

maneuverability, stealth, working<br />

environments, scalability—has the<br />

potential to significantly improve<br />

upon current state-of-the-art in AUV<br />

technologies.<br />

The development of a batoidinspired<br />

underwater vehicle can be<br />

classified in terms of the approach<br />

taken to achieve ray-like swimming.<br />

The first is developing a batoidinspired<br />

AUV that performs as a platform<br />

to test the capabilities of a variety<br />

of traditional <strong>and</strong> novel actuating technologies.<br />

The motivation here is to<br />

demonstrate the capabilities of such<br />

devices—usually in terms of force<br />

<strong>and</strong> stroke—<strong>and</strong> is not necessarily a<br />

desire to truly replicate the biological<br />

system. These actuators include<br />

electroactive polymers, fluidic muscles,<br />

shape memory alloys, motors<br />

<strong>and</strong> servomotors. Of particular interest<br />

in testing these actuators has been the<br />

challenge of quantifying the swimming<br />

performance of the particular<br />

vehicle, many of which do not necessarily<br />

mimic the kinematics of batoid<br />

fishes. As biology has limitations due<br />

to the materials available to construct<br />

abody<strong>and</strong>theevolutionaryprocess<br />

that produces an organism, possible<br />

improvements to the basic body plan<br />

can perhaps be engineered to enhance<br />

performance beyond the capabilities of<br />

nature.<br />

The second approach considers fundamental<br />

questions associated with<br />

biology’s solutions to propulsion,<br />

maneuverability, stability, <strong>and</strong> stealth.<br />

<strong>Technology</strong> is used in this case to replicate<br />

the biology to help answer these<br />

questions. Using the underlying biology<br />

as the basis for inquiry, the<br />

mechanisms that dictate batoid swimming<br />

performance are explored. This is<br />

being done through the design <strong>and</strong> development<br />

of artificial systems—either<br />

real or virtual—that can achieve near<br />

identical kinematic motions of the rays<br />

being studied. Specifically, researchers<br />

are working to elucidate the dominant<br />

mechanisms in batoid swimming that<br />

dictate efficiency <strong>and</strong> maneuverability.<br />

A key outcome of this work is to fully<br />

explore nature’s design space <strong>and</strong><br />

beyond. By mimicking biology, we<br />

attempt to elucidate the key features<br />

that control <strong>and</strong> optimize function.<br />

Nature evolves solutions that satisfy<br />

multiple constraints; engineers <strong>and</strong><br />

scientists can design for a single desired<br />

outcome. By identifying <strong>and</strong> quantifying<br />

the key features/characteristics that<br />

dictate optimality, we can judiciously<br />

choose to build these into an artificial<br />

system, depending on the required<br />

functionality of the device. For example,<br />

one may desire a vehicle that<br />

can swim for as long as possible or<br />

as fast as possible—two different solutions<br />

may be necessary for these two<br />

requirements.<br />

As mentioned in Underst<strong>and</strong>ing<br />

Biological Foundation, there has<br />

been an extensive study of the underlying<br />

cartilage structure of various<br />

July/August 2011 Volume 45 Number 4 105


FIGURE 5<br />

batoid rays (Schaefer & Summers,<br />

2005). Biomechanical studies of the<br />

cartilage arrangement have been carried<br />

out to examine the relationship<br />

between the form of the underlying<br />

structure <strong>and</strong> its impact on the function<br />

(Russo et al., 2011). In this<br />

work, Russo et al. have taken the<br />

cartilage architecture <strong>and</strong> developed<br />

a numerical model to study the kinematic<br />

function of this form. This<br />

initial study is beginning to explore<br />

the relationship between form <strong>and</strong><br />

function.<br />

One of the most exciting developments<br />

in the creation of these bioinspired<br />

devices <strong>and</strong> vehicles is in the<br />

development of rapid prototyping fabrication<br />

(Figure 5). This has opened<br />

the possibility to build a cartilage<br />

structure that uses the same design<br />

principles observed in nature, as described<br />

by Schaefer & Summers<br />

(2005). In this physical model, cartilage<br />

elements are connected in the<br />

spanwise direction with cross-bracing<br />

in the chordwise direction to mimic<br />

the architecture of the Atlantic stingray<br />

(www.bartsmithlabs.com). This<br />

technology enables the design to be<br />

quickly <strong>and</strong> easily varied so as to answer<br />

questions regarding the influence<br />

of the architecture on kinematic<br />

performance. The images in Figure 5<br />

are compelling, as they demonstrate<br />

kinematic capabilities <strong>and</strong> possibilities<br />

of such a structure. By mimicking<br />

the underlying structure of biology,<br />

we can explore the capabilities of<br />

these species <strong>and</strong> potentially exp<strong>and</strong><br />

upon them.<br />

Significant progress has been made,<br />

but there is still much to be done. Information<br />

on the kinematics of swimming<br />

is being generated, but there is<br />

still much to be learned, especially<br />

with respect to some of the more fine<br />

motor skills observed. Also, not much<br />

Example of artificial cartilage structure design using rapid-prototyping technology. The individual<br />

elements represent the cartilage elements found in the pectoral fins of batoid rays. The arrangement<br />

<strong>and</strong> connectivity are similar to portions found in the Atlantic stingray.<br />

is known about the hydroacoustic<br />

properties of the biology. Material<br />

properties of the constituent parts of<br />

the pectoral fin are needed to improve<br />

the fidelity biomechanical models that<br />

describe form <strong>and</strong> function. Lastly,<br />

more investigation of the sensing <strong>and</strong><br />

control strategies of batoids is needed.<br />

This improved underst<strong>and</strong>ing will<br />

providevaluableinsightwhenmore<br />

sophisticated vehicles are developed.<br />

With regard to engineering a batoidinspired<br />

AUV, there are huge opportunities<br />

in actuator design <strong>and</strong> development.<br />

Structural <strong>and</strong> material design <strong>and</strong> selection<br />

also are areas that need to be<br />

addressed. For example, how do we design<br />

a skin that can accommodate both<br />

the out-of-place hydrodynamic forces<br />

<strong>and</strong> the potential in-plane stretching<br />

experienced during actuation. How<br />

do the properties of the artificial system<br />

scale with the biological properties<br />

Actuation technology is an area<br />

that has the potential to revolutionize<br />

the field of biomimicry <strong>and</strong> bioinspired<br />

engineering.<br />

In this paper, we have focused on<br />

a small subset of bio-inspired underwater<br />

vehicles. We have presented a<br />

review of work related to the development<br />

of a bio-inspired underwater<br />

robot—actuation technology integrated<br />

into a batoid-like vehicle <strong>and</strong><br />

underst<strong>and</strong>ing the biological foundation<br />

to explore the full design space.<br />

It is clear though that these two categories<br />

are very closely related. Without<br />

actuator development, it may not be<br />

possible to achieve anything close to<br />

what biology achieves. But a clear picture<br />

of biology function is needed<br />

so that actuator requirements can be<br />

quantified. Synergy between biology,<br />

biomimicry, <strong>and</strong> bio-inspired engineering<br />

is essential if we want to<br />

develop the next generation of underwater<br />

vehicles.<br />

106 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Acknowledgments<br />

The authors would like to acknowledge<br />

funding from the Office of Naval<br />

Research through the MURI program<br />

on Biologically Inspired Autonomous<br />

Sea Vehicles (Program Manager: Dr.<br />

R. Brizzolara, Contract No. N00014-<br />

08-1-0642) <strong>and</strong> the David <strong>and</strong> Lucille<br />

Packard Foundation.<br />

References<br />

Abbott, I.H., & von Doenhoff, A.E. 1949.<br />

Theory of Wing Sections: Including a<br />

Summary of Airfoil Data. New York:<br />

McGraw-Hill.<br />

Ayers, J., Wilbur, C., & Olcott, C. 2000.<br />

Lamprey robots. In: Proceedings of 1st International<br />

Symposium on Aqua Bio-mechanisms,<br />

eds. Wu, T., & Kato, N. Honolulu, Hawaii:<br />

Tokai University Pacific Center.<br />

Anderson, J.M., & Chhabra, N.K. 2002.<br />

Maneuvering <strong>and</strong> stability performance of<br />

a robotic tuna. Integr Comp Biol. 42(1):<br />

118-26. doi: 10.1093/42.1.118.<br />

B<strong>and</strong>yopadhyay, P.R. 2005. Trends in biorobotic<br />

autonomous undersea vehicles.<br />

IEEE J Oceanic Eng. 30(1):109-139.<br />

doi: 10.1109/JOE.2005.843748.<br />

Barrett, D., Grosenbaugh, M., & Triantafyllou,<br />

M. 1996. The optimal control of a flexible<br />

hull robotic undersea vehicle propelled by<br />

an oscillating foil. In: Proceedings of the Symposium<br />

on Autonomous Underwater Vehicle<br />

<strong>Technology</strong>, AUV’96. 1-9. Monterey, CA:<br />

IEEE. doi: 10.1109/AUV.1996.532833.<br />

Blidberg, D.R. 2001. The development of<br />

autonomous underwater vehicles (AUVs):<br />

A brief summary. In: IEEE ICRA Conference,<br />

Seoul Korea, May 21-25, 2001.<br />

Borgen, M.G., Washington, G.N., & Kinzel,<br />

G.L. 2003. Design <strong>and</strong> evolution of a piezoelectrically<br />

actuated miniature swimming<br />

vehicle. IEEE-ASME T Mech. 8(1):66-76.<br />

doi: 10.1109/TMECH.2003.809131.<br />

Breder, C.M., Jr. 1926. The locomotion of<br />

fishes. Zoologica. 4:159-297.<br />

Brower, T.P.L. 2006. Design of a manta<br />

ray inspired underwater propulsive mechanism<br />

for long range, low power operation. Master’s<br />

thesis, Tufts University.<br />

Cai, Y., Bi, S., & Zheng, L. 2010. Design<br />

<strong>and</strong> experiments of a robotic fish imitating<br />

cow-nosed ray. J Bionic Eng. 7(2):120-6.<br />

doi: 10.1016/S1672-6529(09)60204-3.<br />

Chen, Z., Um, T.I., & Bart-Smith, H. 2011.<br />

A novel fabrication of ionic polymer-metal<br />

composite membrane actuator capable of<br />

3-dimensional kinematic motions. Sensor<br />

Actuat A-Phys. 168(1):131-9. doi: 10.1016/<br />

j.sna.2011.02.034.<br />

Clark, R.P., & Smits, A.J. 2006. Thrust<br />

production <strong>and</strong> wake structure of a batoidinspired<br />

oscillating fin. J Fluid Mech. 562:<br />

415-29. doi: 10.1017/S0022112006001297.<br />

Colgate, J.E., & Lynch, K.M. 2004.<br />

Mechanics <strong>and</strong> control of swimming: A<br />

review. IEEE J Oceanic Eng. 29(3):660-73.<br />

Combes, S.A., & Daniel, T.L. 2001. Shape,<br />

apping <strong>and</strong> exion: wing <strong>and</strong> fin design for<br />

forward flight. J Exp Biol. 204(12):2073-85.<br />

Crespi, A., Badertscher, A., Guignard, A., &<br />

Ijspeert, A.J. 2004. An amphibious robot<br />

capable of snake <strong>and</strong> lamprey-like locomotion.<br />

In: Proceedings of the 35th International<br />

Symposium on Robotics ISR 2004. Paris,<br />

France: ISR publisher.<br />

Curet, O.M., Patankar, N.A., Lauder, G.V.,<br />

& MacIver, M.A. 2010. Aquatic manoeuvering<br />

with counter-propagating waves: a novel<br />

locomotive strategy. J R Soc Interface.<br />

8(60):1041-50.<br />

Dennler, G., Scharber, M., Ameri, T.,<br />

Denk, P., Forberich, K., Waldauf, C., &<br />

Brabec, C. 2008. Design rules for donors<br />

in bulk-heterojunction t<strong>and</strong>em solar cells<br />

Towards 15% Energy-Conversion Eficiency.<br />

Adv Mater. 20(3):579-83. doi: 10.1002/<br />

adma.200702337.<br />

Dobbins, P. 2007. Dolphin sonar: Modelling<br />

a new receiver concept. Bioinspir Biomim. 2:19.<br />

doi: 10.1088/1748-3182/2/1/003.<br />

Drucker, E.G., & Jensen, J.S. 1997. Kinematic<br />

<strong>and</strong> electromyographic analysis of steady<br />

pectoral fin swimming in the surfperches.<br />

J Exp Biol. 200:1709-23.<br />

Festo, M. 2008. Fluidic Muscle Brochure.<br />

http://www/festo.com/cms/en-us_us/5030.htm.<br />

Fish, F.E., & Lauder, G.V. 2006. Passive <strong>and</strong><br />

active ow control by swimming fishes <strong>and</strong><br />

mammals. Annu Rev Fluid Mech. 38:193-224.<br />

doi: 10.1146/annurev.fluid.38.050304.092201.<br />

Fish, F.E., Lauder, G.V., Mittal, R., Techet,<br />

A.H., Triantafyllou, M.S., Walker, J.A., &<br />

Webb, P.W. 2003. Conceptual design for the<br />

construction of a biorobotic AUV based on<br />

biological hydrodynamics. In: Proceedings of<br />

the 13th International Symposium on Unmanned<br />

Untethered Submersible <strong>Technology</strong>.<br />

Durham, NH: AUSI (Autonomous Undersea<br />

Systems Institute).<br />

Fish, F.E., Weber, P.W., Murray, M.M., &<br />

Howle, L.E. 2011. The humpback whale’s<br />

ipper: Application of bio-inspired tubercle<br />

technology. Integr Comp Biol. 211:1857-67.<br />

Gao, J., Bi, S., Xu, Y., & Liu, C. 2007.<br />

Development <strong>and</strong> design of a robotic manta<br />

ray featuring flexible pectoral fins. In: IEEE<br />

International Conference on Robotics <strong>and</strong><br />

<strong>Biomimetics</strong>, December 15-18, 2007. ROBIO<br />

2007, 519-23. Yalong Bay of Sanya, China:<br />

IEEE.<br />

Gibb, A., Jayne, B., & Lauder, G. 1994.<br />

Kinematics of pectoral fin locomotion in the<br />

bluegill sunfish Lepomis machrochirus.<br />

J Exp Biol. 189:133-61.<br />

Heine, C. 1992. Mechanics of apping fin<br />

locomotion in the cownose ray, Rhinoptera<br />

bonasus (Elasmobranchii, Myliobatidae).<br />

Ph.D. thesis, Duke University.<br />

Jaffe, J., Moore, K., McLean, J., & Str<strong>and</strong>,<br />

M. 2001. Underwater optical imaging: status<br />

<strong>and</strong> prospects. Oceanography. 14(3):66-76.<br />

Jalbert, J., Baker, J., Duchesney, J., Pietryka,<br />

P., Dalton, W., Blidberg, D., … Holappa, K.<br />

2003. Solarpowered autonomous underwater<br />

vehicle development. In: Proceedings of the<br />

Thirteenth International Symposium on Unmanned<br />

Untethered Submersible <strong>Technology</strong>.<br />

Durham, NH: Autonomous Undersea Systems<br />

Institute.<br />

July/August 2011 Volume 45 Number 4 107


Kato, N. 1998. Locomotion by mechanical<br />

pectoral fins. J Mar Sci Technol. 3(3):113-21.<br />

doi: 10.1007/BF02492918.<br />

Kato, N., Ando, Y., Tomokazu, A., Suzuki,<br />

H., Suzumori, K., K<strong>and</strong>a, T., & Endo, S.<br />

2008. Elastic pectoral fin actuators for biomimetic<br />

underwater vehicles. In: Bio-mechanisms<br />

of Swimming <strong>and</strong> Flying, eds. Kato, N., &<br />

Kamimura, S., Part 3, 271-82. Springer.<br />

doi: 10.1007/978-4-431-73380-5_22.<br />

Kato, N., & Furushima, M. 2002. Pectoral<br />

fin model for maneuver of underwater<br />

vehicles. In: Autonomous Underwater<br />

Vehicle <strong>Technology</strong>, 1996. AUV’96,<br />

Proceedings of the 1996 Symposium on<br />

IEEE. pp. 49-56. Durham, NH: Autonomous<br />

Undersea Systems Institute.<br />

Klausewitz, W. 1964. Der lokomotionsmodus<br />

der Flugelrochen (Myliobatoidei).<br />

Zool Anz. 173:111-20.<br />

Kocak, D., & Caimi, F. 2005. The current<br />

art of underwater imaging with a glimpse<br />

of the past <strong>and</strong> vision of the future. Mar<br />

Technol Soc J. 39(3):5-26. doi: 10.4031/<br />

002533205787442576.<br />

Lauder, G.V., & Jayne, B.C. 1996. Pectoral<br />

fin locomotion in fishes: Testing drag-based<br />

models using three-dimensional kinematics.<br />

Am Zool. 36(6):567-81.<br />

Lauder, G.V., Nauen, J.C., & Drucker, E.G.<br />

2002. Experimental hydrodynamics <strong>and</strong> evolution:<br />

function of median fins in ray-finned<br />

fishes 1. Integr Comp Biol. 42(5):1009-17.<br />

doi: 10.1093/icb/42.5.1009.<br />

Licht, S., Polidoro, V., Flores, M., Hover, F.S.,<br />

& Triantafyllou, M.S. 2004. Design <strong>and</strong><br />

projected performance of a apping foil AUV.<br />

IEEE J Oceanic Eng. 29(3):786-94.<br />

doi: 10.1109/JOE.2004.833126.<br />

Listak, M., Martin, G., Pugal, D., Aabloo, A.,<br />

& Kruusmaa, M. 2005. Design of a semiautonomous<br />

biomimetic underwater vehicle for<br />

environmental monitoring. In: Computational<br />

Intelligence in Robotics <strong>and</strong> Automation.<br />

Espoo, Finl<strong>and</strong>: IEEE.<br />

Low, K.H., & Willy, A. 2006. Biomimetic<br />

motion planning of an undulating robotic<br />

fish fin. J Vib Control. 12:1337-59.<br />

doi: 10.1177/1077546306070597.<br />

Mojarrad, M. 2000. AUV biomimetic<br />

propulsion. In: OCEANS 2000 MTS/IEEE<br />

Conference <strong>and</strong> Exhibition. pp. 2141-6.<br />

Providence, RI: IEEE.<br />

Moored, K.W. August 2010. The design of a<br />

novel tensegrity-based synthetic pectoral fin<br />

for bio-inspired propulsion. Ph.D. thesis,<br />

University of Virginia.<br />

Moored, K.W., & Bart-Smith, H. 2007. The<br />

analysis of tensegrity structures for the design<br />

of a morphing wing. J Appl Mech. 74:668-76.<br />

doi: 10.1115/1.2424718.<br />

Moored, K.W., & Bart-Smith, H. 2009. Investigation<br />

of clustered actuation in tensegrity<br />

structures. Int J Solids Struct. 46:3272-81.<br />

doi: 10.1016/j.ijsolstr.2009.04.026.<br />

Moored, K.W., Dewey, P.A., Leftwich, M.C.,<br />

Bart-Smith, H., & Smits, A.J. 2011a.<br />

Bio-inspired propulsion mechanisms based<br />

on manta ray locomotion. Mar Technol Soc J.<br />

45(4):110-8.<br />

Moored, K.W., Kemp, T.H., Houle, N.E., &<br />

Bart-Smith, H. 2011b. Analytical predictions,<br />

optimization <strong>and</strong> design of a tensegrity-based<br />

artificial pectoral fin. Int J Solids Struct.<br />

(Accepted for Publication).<br />

Palmisano, J., Geder, J., Ramamurti, R.,<br />

Liu, K.J., Cohen, J., Mengesha, T., … Ratna,<br />

B. 2008. Design, development, <strong>and</strong> testing<br />

of flapping fins with actively controlled curvature<br />

for an unmanned underwater vehicle.<br />

In: Bio-mechanisms of Swimming <strong>and</strong> Flying:<br />

Fluid Dynamics, Biomimetic Robots, <strong>and</strong><br />

Sports Science, eds. Kato, N., & Kamimura,<br />

S., 283-94. Berlin: Springer.<br />

Palmisano, J., Ramamurti, R., Lu, K.J.,<br />

Cohen, J., S<strong>and</strong>berg, W., & Ratna, B. 2007.<br />

Design of a biomimetic controlled-curvature<br />

robotic pectoral fin. In: IEEE International<br />

Conference on Robotics <strong>and</strong> Automation<br />

(ICRA). 966-73. IEEE.<br />

Pederzani, J.-N., Moored, K.W., Fish, F.E.,<br />

& Haj-Hariri, H. 2011. A numerical investigation<br />

of the hydrodynamic signature of<br />

batoid swimming. In: <strong>Society</strong> for Integrative<br />

<strong>and</strong> Comparative Biology, Annual meeting.<br />

Salt Lake City, UT: Oxford Univ Press Inc.<br />

Rosenberger, L.J. 2001. Pectoral fin locomotion<br />

in batoid fishes: Undulation versus<br />

oscillation. J Exp Biol. 204(2):379-94.<br />

Russo, R.S., Blemker, S., Fish, F.E., Moored,<br />

K.W., & Bart-Smith, H. 2011. Form function<br />

relationship between ray skeletal architecture<br />

<strong>and</strong> ray locomotion. In: Annual meeting of<br />

the <strong>Society</strong> for Integrative <strong>and</strong> Comparative<br />

Biology. Salt Lake City, UT: Oxford Univ<br />

Press Inc.<br />

Sasko, D.E., Dean, M.N., Motta, P.J., &<br />

Hueter, R.E. 2006. Prey capture behavior<br />

<strong>and</strong> kinematics of the Atlantic cownose ray,<br />

Rhinoptera bonasus. Zoology. 109(3):171-81.<br />

doi: 10.1016/j.zool.2005.12.005.<br />

Schaefer, J.T., & Summers, A.P. 2005.<br />

Batoid wing skeletal structure: Novel<br />

morphologies, mechanical implications, <strong>and</strong><br />

phylogenetic patterns. J Morphol. 264(3):<br />

298-313. doi: 10.1002/jmor.10331.<br />

Sfakiotakis, M., Laue, D., & Davies, B.<br />

2005. An experimental undulating-fin device<br />

using the parallel bellows actuator. In:<br />

Robotics <strong>and</strong> Automation, Proceedings 2001<br />

ICRA. IEEE International Conference on.<br />

pp. 2356-62. Seoul, Korea: IEEE.<br />

Sfakiotakis, M., Lane, D.M., & Davies,<br />

J.B.C. 1999. Review of fish swimming modes<br />

for aquatic locomotion. IEEE J Oceanic Eng.<br />

24(2):237-52. doi: 10.1109/48.757275.<br />

Stocking, J.B., Eberhardt, W.C., Shakhsheer,<br />

J.R., Paulus, M.A., & Calhoun, B.H. 2010.<br />

A capacitance-based whisker-like artificial<br />

sensor for fluid motion sensing. In: IEEE<br />

Sensors. Waikoloa, Big Isl<strong>and</strong>, Hawaii.<br />

Summers, A. 2000. Stifiening the stingray<br />

skeleton: An investigation of durophagy in<br />

myliobatid stingrays (Chondrichthyes, Batoidea,<br />

Myliobatidae). J Morphol. 243(2):113-26.<br />

doi: 10.1002/(SICI)1097-4687(200002)<br />

243:23.0.CO;2-A.<br />

Suzumori, K., Endo, S., K<strong>and</strong>a, T., Kato, N.,<br />

& Suzuki, H. 2007. A bending pneumatic<br />

rubber actuator realizing soft-bodied manta<br />

swimming robot. In: 2007 IEEE International<br />

108 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Conference on Robotics <strong>and</strong> Automation.<br />

pp. 4975-980. IEEE.<br />

Takagi, K., Yamamura, M., Luo, Z., Onishi,<br />

M., Hirano, S., Asaka, K., & Hayakawa, Y.<br />

2007. Development of a rajiform swimming<br />

robot using ionic polymer artificial muscles.<br />

In: Intelligent Robots <strong>and</strong> Systems, 2006<br />

IEEE/RSJ International Conference on.<br />

pp. 1861-6. Beijing, China: IEEE.<br />

Tangorra, J., Anquetil, P., Fofonoff, T., Chen,<br />

A., Zio, M., & Hunter, I. 2007. The application<br />

of conducting polymers to a biorobotic<br />

fin propulsor. Bioinspir Biomim. 2:S6.<br />

doi: 10.1088/1748-3182/2/2/S02.<br />

Tangorra, J., Davidson, S., Hunter, I.,<br />

Madden, P., Lauder, G., Haibo, D., …<br />

Mittal, R. 2008a. The development of a<br />

biologically inspired propulsor for unmanned<br />

underwater vehicles. IEEE J Oceanic Eng.<br />

32(3):533-50.<br />

Tangorra, J., Lauder, G., Madden, P., Mittal, R.,<br />

Bozkurttas, M., & Hunter, I. 2008b. A<br />

biorobotic apping fin for propulsion <strong>and</strong><br />

maneuvering. In: Robotics <strong>and</strong> Automation,<br />

2008. ICRA 2008. IEEE International<br />

Conference on. pp. 700-5. Pasadena, CA:<br />

IEEE.<br />

V. y Alvarado, P., Chin, S., Larson, W.,<br />

Mazumdar, A., & Youcef-Toumi, K. 2010. A<br />

soft body under-actuated approach to multi<br />

degree of freedom biomimetic robots: A<br />

stingray example. In: Biomedical Robotics <strong>and</strong><br />

Biomechatronics (BioRob), 2010 3rd IEEE<br />

RAS <strong>and</strong> EMBS International Conference<br />

on. pp. 473-8. Tokyo, Japan: IEEE.<br />

Walker, J.A., & Westneat, M.W. 1997.<br />

Labriform propulsion in fishes: kinematics<br />

of flapping aquatic flight in the bird wrasse<br />

Gomphosus varius (Labridae). J Exp Biol.<br />

200:1549-69.<br />

Wang, Z., Hang, G., Wang, Y., Li, J., &<br />

Du, W. 2008. Embedded SMA wire actuated<br />

biomimetic fin: a module for biomimetic<br />

underwater propulsion. Smart Mater<br />

Struct. 17:025039. doi: 10.1088/0964-<br />

1726/17/2/025039.<br />

underwater glider. In: International Symposium<br />

on Unmanned Untethered Submersible<br />

<strong>Technology</strong>. 75-85. Durham, NH: Autonomous<br />

Undersea Systems Institute.<br />

Webb, P.W. 1994. Mechanics <strong>and</strong> Physiology<br />

of Animal Swimming: The Biology of Fish<br />

Swimming. Cambridge, UK: Cambridge<br />

University Press. 45-62 pp.<br />

Yan, Q., Zhang, S., & Yang, J. 2010. Initial<br />

implementation of basic actuated unit of a<br />

flexible pectoral fin driven by SMA. In:<br />

Mechatronics <strong>and</strong> Automation (ICMA),<br />

2010 International Conference on.<br />

pp. 899-904. Xi’an, China: IEEE.<br />

Yang, S., Qiu, J., & Han, X. 2009. Kinematics<br />

modeling <strong>and</strong> experiments of pectoral<br />

oscillation propulsion robotic fish. J Bionic<br />

Eng. 6(2):174-9. doi: 10.1016/S1672-6529<br />

(08)60114-6.<br />

Yang, Y., Chen, J., Engel, J., P<strong>and</strong>ya, S.,<br />

Chen, N., Tucker, C., … Liu, C. 2006.<br />

Distant touch hydrodynamic imaging with<br />

an artificial lateral line. P Natl Acad Sci.<br />

103(50):18891.<br />

Yong-hua, Z., Shi-wu, Z., Jie, Y., & Low, K.<br />

2007. Morphologic optimal design of bionic<br />

undulating fin based on computational fluid<br />

dynamics. In: Mechatronics <strong>and</strong> Automation,<br />

2007. ICMA 2007. International Conference<br />

on. pp. 491-6. Harbin, Heilongjian, China:<br />

IEEE.<br />

Yu, J., Hu, Y., Fan, R., Wang, L., & Huo, J.<br />

2007. Mechanical design <strong>and</strong> motion<br />

control of a biomimetic robotic dolphin.<br />

Adv Robotics. 21(3):499-513.<br />

doi: 10.1163/156855307780131974.<br />

Yu, J., Tan, M., Wang, S., & Chen, E. 2004.<br />

Development of a biomimetic robotic fish<br />

<strong>and</strong> its control algorithm. IEEE Sys Man<br />

Cybern B. 34(4), 1798-810.<br />

Zhou, C., & Low, K. 2010. Better endurance<br />

<strong>and</strong> load capacity: An improved design of<br />

manta ray robot. J Bionic Eng. 7:S137-44.<br />

doi: 10.1016/S1672-6529(09)60227-4.<br />

Webb, D., & Simonetti, P. 1999. The<br />

Slocum AUV: An environmentally propelled<br />

July/August 2011 Volume 45 Number 4 109


PAPER<br />

Bioinspired Propulsion Mechanisms<br />

Based on Manta Ray Locomotion<br />

AUTHORS<br />

Keith W. Moored<br />

Peter A. Dewey<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering, Princeton University<br />

Megan C. Leftwich<br />

Physics Division, Los Alamos<br />

National Laboratory<br />

Hilary Bart-Smith<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering, University of Virginia<br />

Alex<strong>and</strong>er J. Smits<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering, Princeton University<br />

Introduction<br />

I<br />

n recent years, there has been<br />

considerable interest in developing<br />

novel underwater vehicles that use<br />

propulsion systems inspired by biology<br />

(Colgate & Lynch, 2004;<br />

B<strong>and</strong>yopadhyay, 2005). Such vehicles<br />

have the potential to open up new mission<br />

capabilities <strong>and</strong> improve maneuverability,<br />

efficiency, <strong>and</strong> speed (Fish<br />

et al., 2003, 2011). Here we explore<br />

how various aspects of biological locomotion<br />

relate to performance in the<br />

particular case of ray-like swimming,<br />

with the aim of informing the design<br />

of new vehicles.<br />

The kinematic motion of batoid fish<br />

(rays) is based on the chordwise traveling<br />

wave that is a hallmark of their motion<br />

(Rosenberger, 2001). Species are classified<br />

as being oscillatory if the traveling<br />

wave wavelength is longer than the<br />

chord of their fin <strong>and</strong> undulatory if the<br />

wavelength is less than the chord of<br />

ABSTRACT<br />

Mobuliform swimmers are inspiring novel approaches to the design of underwater<br />

vehicles. These swimmers, exemplified by manta rays, present a model for new classes<br />

of efficient, highly maneuverable, autonomous undersea vehicles. To improve our<br />

underst<strong>and</strong>ing of the unsteady propulsion mechanisms used by these swimmers,<br />

we report detailed studies of the performance of robotic swimmers that mimic aspects<br />

of the animal propulsive mechanisms. We highlight the importance of the undulatory<br />

aspect of producing efficient manta ray propulsion <strong>and</strong> show that there is a strong interaction<br />

between the propulsive performance <strong>and</strong> the flexibility of the actuating surfaces.<br />

Keywords: mobuliform, manta ray, unsteady, swimming, flexible actuators<br />

their fin. The manta ray is an example<br />

of an oscillatory swimmer. Previously,<br />

Clark <strong>and</strong> Smits (2006) explored the<br />

thrust production <strong>and</strong> efficiencies of an<br />

artificial pectoral fin thatcapturedthe<br />

traveling wave motion <strong>and</strong> observed<br />

efficiencies upwards of 50% for an oscillatory<br />

motion at a fixed flow velocity.<br />

To study the swimming of mantas,<br />

we use artificial or robotic devices that<br />

generate a simple baseline motion that<br />

approximates biological kinematics.<br />

The complexity of the motion is then<br />

progressively increased by adding more<br />

kinematic features until the motion<br />

resembles the biology very closely. At<br />

each level of complexity, various performance<br />

metrics are measured. We explore<br />

theroleofspanwisecurvature,theeffects<br />

ofaspanwisetravelingwave<strong>and</strong>tip<br />

speed modulation, which have not been<br />

previously investigated, <strong>and</strong>, the role<br />

of a chordwise traveling wave motion<br />

is investigated in the performance of<br />

an actively <strong>and</strong> passively flexible fin.<br />

Experiments<br />

Two biorobotic devices were developed<br />

<strong>and</strong> tested. First, an artificial pectoral<br />

fin able to produce root-fixed<br />

pure heaving motions was developed<br />

<strong>and</strong> will be referred to as the heaving<br />

fin. Thefin wascastfromaflexible<br />

plastic <strong>and</strong> actuated with variable degrees<br />

of spanwise curvature. The measurements<br />

on the heaving fin were<br />

conducted in a tow tank (Figure 1).<br />

This facility tows a fin throughstill<br />

water at a fixed velocity, U, inatank<br />

measuring 5-m long, 1-m deep, <strong>and</strong><br />

1.5-m wide, <strong>and</strong> we directly measure<br />

thenetforceproduced,T, <strong>and</strong>the<br />

power imparted into the fluid, P f ,<br />

by a flapping fin structure. The<br />

propulsive efficiency of the motion,<br />

η p ¼ TU= ‐ P ‐ f , can then be calculated<br />

from the thrust <strong>and</strong> power measurements<br />

averaged over a cycle,<br />

‐<br />

T <strong>and</strong> P ‐ f , respectively. If the fin were<br />

unconstrained <strong>and</strong> free-swimming,<br />

then the net force would cause the<br />

fin to accelerate or decelerate to a<br />

new velocity where there is no average<br />

net force. Constraining the fin allows<br />

for the measurement of force production<br />

<strong>and</strong> is a commonly used experimental<br />

approach (Anderson et al.,<br />

1998).<br />

110 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 1<br />

Tow tank facility consisting of an artificial pectoral fin, tow tank, motor, carriage, <strong>and</strong> a control<br />

<strong>and</strong> measurement system.<br />

Second, an artificial pectoral fin capable<br />

of generating a chordwise traveling<br />

wave motion was developed <strong>and</strong><br />

will be referred to as the traveling<br />

wave fin. This fin was used to measure<br />

the free-swimming performance in a<br />

tank measuring 6.7-m long, 1-m<br />

deep, <strong>and</strong> 1-m wide. This fin was<br />

also cast from a flexible plastic, but it<br />

was activated using a number of rigid<br />

spars in the spanwise direction. By reducing<br />

the number of actuating spars,<br />

the degree of passive flexibility of the<br />

fin could be varied. The traveling<br />

wave wavelength was controlled by<br />

changing the phase differences between<br />

adjacent spars. The steady<br />

swimming speed U <strong>and</strong> mean power<br />

input over a cycle P ‐ f were measured,<br />

<strong>and</strong> hence, the energy economy ζ<br />

could be found, where ζ ¼ U = P ‐ f .<br />

Energy economy is the inverse of<br />

cost of transport, both of which are<br />

used extensively in the biological literature<br />

(Schmidt-Nielsen, 1972; Fish<br />

et al., 1991; Liao et al., 2003; Liao,<br />

2004). Energy economy, however, is<br />

a more appropriate engineering metric<br />

as the dimensions are distance/per unit<br />

energy (the units could be miles per<br />

gallon for instance). Efficiency is also<br />

an appropriate performance measure,<br />

but from an experimental point-ofview<br />

it can only be measured when<br />

there is net thrust production or the<br />

fin is not in a free-swimming mode.<br />

Thus, for the experiments in the tow<br />

tank, efficiency was a measurable<br />

performance metric, but in the freeswimming<br />

cases, economy is used instead<br />

because the efficiency could not<br />

be directly measured.<br />

The performance is measured as a<br />

function of the Strouhal number,<br />

St = fA/U, where f is the frequency of<br />

motion, A is the peak-to-peak trailingedge<br />

amplitude of motion at the midspan,<br />

<strong>and</strong> U is the free-stream velocity.<br />

The Strouhal number is a measure<br />

of the lateral to streamwise spacing of<br />

the shed vortices in the wake <strong>and</strong> to a<br />

large extent governs the structure of<br />

the wake. It has been shown to be a<br />

critical parameter in describing the efficient<br />

propulsion of oscillating foils<br />

<strong>and</strong> plates (Anderson et al., 1998;<br />

Buchholz & Smits, 2008) <strong>and</strong> for<br />

swimming (Clark & Smits, 2006;<br />

Borazjani & Sotiropoulos, 2008,<br />

2009) <strong>and</strong> flying animals (Taylor<br />

et al., 2003).<br />

Fixed Velocity Experiments:<br />

Heaving Motion<br />

The skeletal structure of the heaving<br />

fin is composed of three connected<br />

hinged plates. The angular position of<br />

each hinge is individually controlled by<br />

a linear actuator. The fin allows for<br />

out-of-plane motion with no pitching<br />

or undulation in the chordwise direction.<br />

The skeletal structure is embedded<br />

into a compliant PVC polymer.<br />

The PVC is molded around the structure<br />

into a fin with a trapezoidal planform<br />

shape. This shape was chosen to<br />

be a simple representative shape of the<br />

manta ray as well as to maximize spacing<br />

for the internal structure. The fin<br />

has a span length of b =28cm<strong>and</strong><br />

an average chord length of ― c =19<br />

cm, with an aspect ratio, AR = b 2 /S,<br />

of 1.47, wherein S is the planform<br />

area. The cross-sectional shape is a<br />

NACA 0020 airfoil. The trailing edge<br />

is stiffened by a thin metal sheet attached<br />

to the main structure.<br />

Three flapping mode shapes were<br />

explored: flat root-fixed heave (Figure<br />

2(a)), curved root-fixed heave (Figure<br />

2(b)), <strong>and</strong> curved root-fixed heave<br />

with a span-wise traveling wave (Figure<br />

2(c)). For each of these mode<br />

shapes the tip speed of the fin can be<br />

modulated. Previous kinematics studies<br />

of ray locomotion (Heine, 1992;<br />

Rosenberger, 2001) found that in<br />

order to swim faster oscillatory rays<br />

do not vary their beat frequency, but<br />

instead they vary the tip speed of<br />

their fin while holding frequency <strong>and</strong><br />

amplitude constant. This can be<br />

viewed as modifying the actuation<br />

waveform to suit a particular mode of<br />

swimming. To implement this mode<br />

in our experiments, the time-varying<br />

waveform was varied from a pure sinusoid<br />

towards an almost square wave<br />

form (Figure 3). This allows the frequency<br />

<strong>and</strong> amplitude to be held<br />

fixed while the maximum fin tip<br />

speed is increased. The thrust <strong>and</strong> propulsive<br />

efficiency were measured for<br />

each prescribed motion.<br />

July/August 2011 Volume 45 Number 4 111


FIGURE 2<br />

Swimming modes ranging from an artificial/simple motion that approximates manta ray locomotion<br />

to a more complex biologically inspired motion: (a) flat root-fixed heaving motion, (b) curved<br />

root-fixed heaving motion, <strong>and</strong> (c) curved root-fixed heaving motion with a span-wise traveling<br />

wave (tip lag effect).<br />

Free-Swimming Experiments:<br />

Traveling Wave Motion<br />

The traveling wave fin was tested<br />

under free-swimming conditions in a<br />

stationary tank to explore the role<br />

of flexibility in ray-like propulsion.<br />

Low-friction carts were attached to<br />

the fin actuation mechanism to form<br />

a carriage that was mounted on tracks<br />

above the tank (see Figure 4). The<br />

chord of the fin was aligned parallel<br />

to the tracks, which allowed for a single<br />

degree of freedom <strong>and</strong> made the<br />

swimming direction of the carriage<br />

unidimensional, <strong>and</strong> no transverse<br />

motion of the carriage was permitted.<br />

The root of the fin abutted against an<br />

acrylic sheet that was in contact with<br />

thefreesurfaceofthewatertominimize<br />

surface waves. Upon actuating<br />

the fin, the cart propelled itself down<br />

the length of the tank.<br />

FIGURE 3<br />

An elliptical planform fin withan<br />

aspect ratio of 1.6 <strong>and</strong> a NACA 0020<br />

cross-section was cast using a flexible<br />

PVC plastic. Four aluminum spars<br />

were embedded into the fin to provide<br />

actuation. A push-rod connected each<br />

spar to a gear in a gear-train, driven by<br />

a DC motor, that produced a sinusoidal<br />

rotation of the spar about a pivot<br />

point located at the root chord of the<br />

fin. This results in a linearly increasing<br />

amplitude of motion along the span of<br />

the fin (Figure 5). The wavelength of<br />

the traveling wave λ could be varied<br />

by changing the phase difference between<br />

the actuating spars.<br />

In addition, the number of actuating<br />

spars could be varied to allow for a<br />

certain degree of passive flexibility in<br />

the fin. When four actuating spars<br />

are used, the locomotion of the fin<br />

was prescribed for the entire chord of<br />

Holding frequency <strong>and</strong> amplitude constant while modulating tip speed can be achieved by varying<br />

the actuation waveform from a sine wave to a square wave.<br />

the fin, with a traveling wave wavelength<br />

λ a , <strong>and</strong> this fin will be referred<br />

to as the active fin. When the two trailing<br />

edge spars are removed, the trailing<br />

half of the chord of the fin willpassively<br />

respond to the leading edge actuation<br />

<strong>and</strong> external fluid forces. This fin<br />

will be referred to as the passive fin<br />

(Figure 5). For the passive fin, the<br />

traveling wave along the chord is generated<br />

by the first two gears in the gear<br />

train <strong>and</strong> the passive response of the<br />

trailing edge. Due to these compounding<br />

factors, the precise wavelength<br />

along the chord for the passive fin is<br />

unknown, <strong>and</strong> we define an analogous<br />

wavelength, λ p such that λ a =<br />

λ p when the offset between the first<br />

two gears is identical for both the active<br />

<strong>and</strong> passive fin. We define the dimensionless<br />

wavelengths λ a ¼ λ a=C<br />

<strong>and</strong> λ p ¼ λ p=C, whereC is the root<br />

chord of the fin. This study focuses on<br />

cases with λ a ; λ p > 1, representative<br />

of oscillatory swimmers (Rosenberger,<br />

2001).<br />

Results <strong>and</strong> Discussion<br />

Flat <strong>and</strong> Curved Modes<br />

The first two modes of swimming,<br />

the flat mode (Figure 2(a)) <strong>and</strong><br />

thecurvedmode(Figure2(b)),were<br />

studied using the heaving fin.<br />

The thrust coefficient is defined by<br />

C T ¼ T = 1 2 ρU 2 S,whereinT is the<br />

thrust, ρ is the water density, <strong>and</strong> S<br />

is the planform area of the fin. Similarly,<br />

the power coefficient is defined<br />

by C p ¼ P= 1 2 ρU 3 S, wherein P is the<br />

power input to the water (as defined<br />

by Clark & Smits, 2006). Figure 6(a)<br />

shows that the thrust production <strong>and</strong><br />

power input increases as the Strouhal<br />

number increases, as found in previous<br />

studies by Anderson et al. (1998) <strong>and</strong><br />

Dong et al. (2006). The flat mode of<br />

swimming produces more thrust <strong>and</strong><br />

112 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 4<br />

Tank facility for the chordwise traveling wave experiments: (a) perspective view <strong>and</strong> (b) front<br />

view. Drawings not to scale.<br />

FIGURE 5<br />

Traveling wave actuation system for (left) active fin <strong>and</strong> (right) passive fin.<br />

FIGURE 6<br />

Flat mode compared to curved mode: (a) thrust performance as a function of St <strong>and</strong> (b) power<br />

coefficient as a function of St.<br />

uses more power than the curved mode<br />

throughout the Strouhal range. This<br />

results is not unexpected as the flat<br />

mode of swimming sweeps out a larger<br />

volume of fluid than the curved<br />

mode of swimming, causing the overall<br />

increase in the thrust <strong>and</strong> power<br />

coefficients.<br />

Because the two modes of swimming<br />

are dissimilar, comparing the<br />

thrust or efficiency as a function of<br />

Strouhal number may be misleading.<br />

For a fixed Strouhal number, different<br />

amounts of thrust <strong>and</strong> power are produced<br />

by each swimming mode. A<br />

better comparison is the thrust or efficiency<br />

as a function of the power<br />

coefficient because each swimming<br />

mode can be compared at a fixed<br />

power input.<br />

Figure 7a shows this comparison.<br />

For both modes of swimming the<br />

power input increases as the thrust increases,<br />

although at higher power<br />

input the thrust increases at a slower<br />

rate. For all power inputs, more thrust<br />

is produced for the flat mode of swimming<br />

than for the curved mode of<br />

swimming, while the thrust tends to<br />

the drag of the motionless fin asthe<br />

power input tends to zero. Interestingly,<br />

observations of swimming<br />

manta rays do not indicate that they<br />

use this higher-performance flat<br />

mode <strong>and</strong> instead use the curved<br />

mode for swimming. This suggests<br />

that the curved mode, coupled with<br />

another mechanism (perhaps a chordwise<br />

traveling wave), more fully characterizes<br />

the swimming mechanics of<br />

manta rays.<br />

Figure 7(b) shows the propulsive<br />

efficiency as a function of Strouhal<br />

number. The efficiency of the curved<br />

mode first rises quickly rise with<br />

increasing Strouhal number, <strong>and</strong> then<br />

apeakinefficiency is attained, followed<br />

by a slow decline in efficiency<br />

July/August 2011 Volume 45 Number 4 113


FIGURE 7<br />

Flat mode compared to curved mode: (a) thrust coefficient as a function of the power coefficient<br />

<strong>and</strong> (b) efficiency as a function of Strouhal number.<br />

at the higher Strouhal numbers. This<br />

trend is characteristic of efficiency<br />

curves for oscillating foils <strong>and</strong> plates<br />

(Clark & Smits, 2006; Anderson<br />

et al., 1998; Heathcote et al., 2006a,<br />

2006b; Heathcote & Gursul, 2007;<br />

Buchholz & Smits, 2008). In general,<br />

the efficiency crosses from negative<br />

(net drag) to positive (net thrust) <strong>and</strong><br />

continues to increase while at high<br />

values of St there is a decline in efficiency<br />

that follows potential flow theory<br />

(Jones et al., 1998). Thus, a peak<br />

in efficiency is expected at an intermediate<br />

Strouhal number. The peak efficiency<br />

for the curved mode is about<br />

20% occurring at a St = 0.2, which is<br />

in the range of 0.2 < St


FIGURE 9<br />

Tip lag mode: (a) efficiency as a function of St <strong>and</strong> (b) peak efficiency dependence on tip lag.<br />

function was used to modulate the tip speed while fixing the frequency <strong>and</strong><br />

amplitude, as given by<br />

xt ðÞ¼at ð ϕÞ 4 þbt ð ϕÞ 2 þA; 0 ≤ t ≤ T =2<br />

a ¼<br />

U sqr<br />

max<br />

2ϕ 3 þ A ϕ 4 ; b ¼ 2A<br />

ϕ 2<br />

ϕ ¼ 1=4f ; U sqr<br />

max ¼ 2παU ∞ St<br />

þ U sqr<br />

max<br />

2ϕ<br />

The amplitude, A, the frequency, f, <strong>and</strong> the maximum tip speed, U sqr<br />

max, all determine<br />

the shape of the quartic function. The period is T. By holding the frequency<br />

<strong>and</strong> amplitude constant the tip speed can be modulated by varying α<br />

between 1 <strong>and</strong> 2.667. When α is 1, the quartic function matches a sine wave,<br />

but when α is 2.667, the maximum tip speed is 2.667 times faster than the maximum<br />

tip speed of a sine wave without varying the frequency or amplitude. It<br />

should be noted, however, that many animals increase their swimming speed<br />

by modulating their frequency of motion while fixing the amplitude. Frequency<br />

modulation will increase the Strouhal number by increasing frequency.<br />

Tests were conducted at two Strouhal numbers, 0.2 <strong>and</strong> 0.25, <strong>and</strong> the amplitude<br />

was fixed at A/b = 0.44 for α =1-2.4. Figure 10a shows the thrust coefficient<br />

dependence on the tip speed, α, for a Strouhal number of 0.2. As the tip speed is<br />

increased the thrust coefficient increases linearly, indicating that tip speed modulation<br />

can be used to increase thrust production, as exhibited by rays (Heine,<br />

1992; Rosenberger, 2001).<br />

Figure 10(b) shows the thrust coefficient plotted against the power coefficient<br />

for both frequency modulation <strong>and</strong> tip speed modulation. Tip speed modulation<br />

produces less thrust than frequency modulation for the same input power, suggesting<br />

that tip speed modulation is not an efficiency strategy. However, incorporating<br />

a chordwise traveling wave may change this outcome. Additionally, as the<br />

value of α is increased, the fin has an increasing period of effectively no motion. In<br />

a free-swimming test, the increasing resting time for the fin would result in an<br />

unpowered gliding period over part of the flapping cycle. This would result in a<br />

burst-<strong>and</strong>-coast behavior that could improve the economy since forward motion<br />

would occur without any input power for part of the flapping cycle. Alternatively,<br />

(1)<br />

tip speed modulation could an effective<br />

flight/sprinting mode, where efficiency<br />

is not important.<br />

Chordwise Traveling Wave<br />

We now investigate the effects of a<br />

chordwise traveling wave, using the<br />

traveling wave fin (the experimental<br />

arrangement was described in Free-<br />

Swimming Experiments: Traveling<br />

Wave Motion).<br />

Clark <strong>and</strong> Smits (2006) investigated<br />

the thrust <strong>and</strong> efficiency of an artificial<br />

pectoral fin using a chordwise<br />

traveling wave motion <strong>and</strong> found efficiencies<br />

peaking near 50% for optimal<br />

conditions (St ≈ 0.25 <strong>and</strong> λ a ≈ 4‐6).<br />

The efficiencies were measured at predetermined<br />

Strouhal numbers. For the<br />

current work, this constraint is not imposed,<br />

<strong>and</strong> the fin was instead actuated<br />

at a given frequency <strong>and</strong> wavelength<br />

<strong>and</strong> allowed to freely swim down the<br />

length of a tow tank. In doing so, the<br />

fin attains its self-propelled swimming<br />

speed <strong>and</strong> Strouhal number for that<br />

frequency <strong>and</strong> wavelength.<br />

Figure 11 shows the steady velocity<br />

achieved as a function of input flapping<br />

frequency for different wavelengths<br />

of actuation. The velocity is<br />

giveninroot-chordlengths(CL)per<br />

second, where C root = 0.254 m. For<br />

the active fin, an almost linear increase<br />

in velocity is observed with increasing<br />

frequency, a result in agreement with<br />

previous studies that found thrust<br />

coefficients increasing with frequency<br />

(Anderson et al., 1998). Peak velocities<br />

were found to occur for λ a ¼ 6 at the<br />

highest flapping frequencies. In these<br />

instances the velocity was upwards of<br />

2 CL/s, corresponding to a dimensional<br />

velocity of 0.51 m/s, highlighting<br />

that this form of propulsion may<br />

prove fruitful for future underwater vehicle<br />

designs that dem<strong>and</strong> relatively<br />

high speeds.<br />

July/August 2011 Volume 45 Number 4 115


FIGURE 10<br />

(a) Thrust coefficient increasing with tip speed increase <strong>and</strong> (b) tip speed modulation compared<br />

to frequency modulation. The parameter α is the ratio of maximum tip speed of the square wave<br />

actuation compared to the maximum tip speed of a sine wave of the same frequency <strong>and</strong> amplitude,<br />

α ¼ Umax=U sqr<br />

max sin .<br />

Figure 11 also reveals the role of<br />

passive flexibility. Here, the fin was actuated<br />

using only the two anterior<br />

spars, leaving the posterior half of the<br />

fin to respond passively to the forcing<br />

by the actuators <strong>and</strong> the fluid forces.<br />

The velocity of this fin still increases<br />

with increasing frequency, but the<br />

trend is no longer linear. Tests in still<br />

water indicate that the passive fin has a<br />

resonant frequency of about 2.4 Hz,<br />

defined as the frequency at which the<br />

trailing edge amplitude is maximized.<br />

FIGURE 11<br />

Fin velocity in chord lengths per second as a<br />

function of flapping frequency. The wavelength<br />

λ a refers to the active fin whileλ p refers to the<br />

passive fin. (Color versions of figures available<br />

online at: http://www.ingentaconnect.com/<br />

content/mts/mtsj/2011/00000045/00000004.)<br />

FIGURE 12<br />

Strouhal number as a function of steady-state<br />

swimming velocity for the active fin compared<br />

with biological data of the manta ray<br />

(Fish, 2010).<br />

As resonance is approached, the amplitude<br />

of the trailing edge motion increases,<br />

resulting in the fin velocity<br />

increasing at an enhanced rate (in comparison<br />

to a linear trend). It should be<br />

noted that the swimming velocities for<br />

the passive fin were approximately<br />

80% of those of the active fin.<br />

The free-swimming Strouhal number<br />

for the active fin, along with manta<br />

ray field data (Fish, 2010), are displayed<br />

in Figure 12 (the Strouhal<br />

number for the passive fin isnot<br />

shown since the trailing edge excursion<br />

of the passive fin is unknown). The experimental<br />

data collapse onto a single<br />

curve, indicating that the Strouhal<br />

number for the freely swimming active<br />

fin does not depend on the wavelength.<br />

As the swimming velocity increases,<br />

the Strouhal number begins<br />

to enter the regime presumed to be efficient<br />

(St = 0.2-0.4; Taylor et al.,<br />

2003). Hence, the optimal swimming<br />

speed for the traveling wave fin, from<br />

the perspective of efficiency, is likely<br />

to be ≥2 CL/s. The biological <strong>and</strong><br />

experimental data display the same<br />

trend, whereby an increase in swimming<br />

velocity yields a decrease in<br />

Strouhal number. Borazjani <strong>and</strong><br />

Sotiropoulos (2008, 2009) were able<br />

to show, for carangiform <strong>and</strong> anguiliform<br />

swimming, that the Strouhal<br />

number for self-propulsion approaches<br />

the efficient regime observed in nature<br />

only with increasing swimming velocity.<br />

The current study supports this<br />

conclusion for mobuliform swimming<br />

as well.<br />

The energy economy, ζ =V c /P f , for<br />

the active <strong>and</strong> passive fins is shown in<br />

Figure 13. In the case of the shortest<br />

wavelength(λ* = 3), the active fin has<br />

a higher energy economy than the passive<br />

fin, but for the longer wavelengths<br />

the energy economy for the passive fin<br />

FIGURE 13<br />

Energy economy as a function of flapping frequency.<br />

The wavelength λ a refers to the active<br />

fin while λ p refers to the passive fin.<br />

116 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


exceeds that of the active fin despite a<br />

decrease in the overall swimming speed<br />

(Figure 11). Clearly, by removing the<br />

two trailing edge spars in creating the<br />

passive fin, the power consumption<br />

decreases significantly compared to<br />

the active fin. The highest energy<br />

economy recorded was 0.18 CL/J<br />

for the passive fin withλ p ¼ 12 <strong>and</strong><br />

f = 2 Hz, but its economy is still trending<br />

upward with increasing frequency,<br />

which may reflect the fact that the<br />

maximum test frequency was still<br />

below the resonant frequency of the<br />

fin (2.4Hz).Leftwich<strong>and</strong>Smits<br />

(2010) found thrust production of a<br />

passively flexible artificial lamprey tail<br />

to increase as resonance is approached.<br />

The current work suggests that the energy<br />

economy also benefits by a system<br />

exploiting the resonant modes of a fin.<br />

The increase in energy economy<br />

with frequency may also be a Strouhal<br />

number effect. Figure 12 indicates that<br />

the fin begins to enter the “efficient”<br />

regime (St = 0.2-0.4) with increasing<br />

swimming velocity (which occurs at<br />

higher flapping frequencies; see Figure<br />

11). While the energy economy is<br />

not a direct measure of efficiency, the<br />

two parameters are inherently linked,<br />

so it is not altogether surprising that<br />

the energy economy increases as the<br />

Strouhal number approaches the supposedly<br />

optimal range. The active fin<br />

with λ a ¼ 12 displays a maximum at<br />

f = 1.6 Hz. It is believed that this<br />

peak is related to changes in the wake<br />

structure that result from the threedimensionality<br />

of the wake associated<br />

with this fin. Dewey et al. (2011)<br />

found that an increasing wavelength<br />

causes the wake to bifurcate into a double<br />

wake structure that is less efficient,<br />

<strong>and</strong> it is believed to be responsible for<br />

the maximum observed in Figure 13.<br />

It may be that the other cases will<br />

eventually reach a maximum due to a<br />

similar mechanism, but further testing<br />

is required to support this suggestion.<br />

Conclusions<br />

This study has explored various kinematic<br />

modes associated with biological<br />

propulsion based on the manta ray.<br />

The results highlight the interdependence<br />

of the kinematic motions <strong>and</strong><br />

fluid–structure interactions on the performance<br />

characteristics of the animal.<br />

A purely heaving fin was used to examine<br />

three kinematic modes of swimming<br />

(flat, curved, <strong>and</strong> tip lag) as well<br />

as variation in the actuation waveform<br />

(sine waveform to square waveform). It<br />

was found that the flat mode of swimming<br />

produces higher efficiency <strong>and</strong><br />

thrust compared to the curved mode<br />

of swimming. Furthermore, there was<br />

no performance benefit foundbyincorporating<br />

tip lag into the motion or<br />

by modulating tip speed instead of frequency.<br />

These results are counter to<br />

the hypothesis that as the motion becomes<br />

more biologically similar the<br />

thrust or efficiency performance will<br />

increase. The maximum thrust coefficient<br />

was found to be about 0.7 at<br />

St =0.45,<strong>and</strong>themaximumpropulsive<br />

efficiency was about 22% at<br />

St = 0.15. These performance metrics<br />

were low, as expected, due to the absenceofachordwisetravelingwave<br />

motion.Whatwasnotexpectedwas<br />

that the performance would be insensitive<br />

to curvature, tip lag, <strong>and</strong> tip<br />

speed variations. Without flexibility,<br />

the motion of the fin may be too constrained<br />

<strong>and</strong> not “natural” enough to<br />

achieve the performance benefits of<br />

the kinematic variations explored.<br />

The presence of a chordwise traveling<br />

wave to generate an undulatory<br />

motion appears to be of prime importance<br />

for efficient propulsion. For example,<br />

Clark <strong>and</strong> Smits (2006) found<br />

efficiencies upwards of 50%. In freeswimming<br />

experiments of artificial<br />

fins similar to that studied by Clark<br />

<strong>and</strong> Smits, we found that they were<br />

able to generate speeds similar to<br />

those observed in nature, of the order<br />

2 CL/s. When the fin was actively actuated,<br />

that is to say that the traveling<br />

wave motion was defined for all points<br />

on the chord of the fin, the steady-state<br />

Strouhal number obtained by the oscillating<br />

fin was found to be independent<br />

of the wavelength <strong>and</strong> exhibited<br />

thesametrendasthemantarayin<br />

nature. That is, at low swimming<br />

velocities, both the manta ray <strong>and</strong> the<br />

artificial fin display high steady-state<br />

Strouhal numbers, but the Strouhal<br />

numbers decrease with increasing<br />

swimming speed <strong>and</strong> they approach<br />

the regime where efficient propulsion<br />

is hypothesized to exist (St =0.2-<br />

0.4). Introducing passive flexibility<br />

into the fin, by restricting the actuationtotheleadingedge<strong>and</strong>letting<br />

the rest of the fin respond passively to<br />

the actuation <strong>and</strong> the external fluid<br />

forces, improved the energy economy.<br />

It was found that the passively actuated<br />

fin achieved steady state swimming<br />

speeds that were approximately 80%<br />

of that of the actively actuated fin,<br />

but because there was a significant decrease<br />

in the power required to propel<br />

the passively actuated fin the energy<br />

economy increased.<br />

Acknowledgments<br />

The authors would like to thank<br />

Daphne Rein-Weston, Dan Quinn,<br />

<strong>and</strong> Dr. Melissa Green for their aid<br />

in developing the low-friction carriage<br />

experiment. We would also like to<br />

thank Professor Frank Fish for correspondence<br />

regarding manta rays in<br />

nature. The authors would like to acknowledge<br />

funding from the Office<br />

July/August 2011 Volume 45 Number 4 117


of Naval Research through the<br />

MURI program on Biologically-<br />

Inspired Autonomous Sea Vehicles<br />

(grant N0001408-1-0642), the<br />

David <strong>and</strong> Lucille Packard Foundation,<br />

the National Science Foundation<br />

(grant CMS-0384884), <strong>and</strong> the Virginia<br />

Space Grant Consortium.<br />

Corresponding Author:<br />

Alex<strong>and</strong>er J. Smits<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering, Princeton University<br />

Princeton, NJ 08544<br />

Email: asmits@princeton.edu<br />

References<br />

Anderson, J.M., Streitlien, K., Barrett, D.S.,<br />

& Triantafyllou, M.S. 1998. Oscillating<br />

foils of high propulsive efficiency.<br />

J Fluid Mech. 360:41-72. doi: 10.1017/<br />

S0022112097008392.<br />

B<strong>and</strong>yopadhyay, P.R. 2005. Trends in<br />

biorobotic autonomous undersea vehicles.<br />

IEEE J Oceanic Eng. 30(1):109-39.<br />

doi: 10.1109/JOE.2005.843748.<br />

Borazjani, I., & Sotiropoulos, F. 2008.<br />

Numerical investigation of the hydrodynamics<br />

of carangiform swimming in the transitional<br />

<strong>and</strong> inertial flow regimes. J Exp Biol.<br />

211:1541-58. doi: 10.1242/jeb.015644.<br />

Borazjani, I., & Sotiropoulos, F. 2009.<br />

Numerical investigation of the hydrodynamics<br />

of anguilliform swimming in the transitional<br />

<strong>and</strong> inertial flow regimes. J Exp Biol.<br />

212:576-92. doi: 10.1242/jeb.025007.<br />

Buchholz, J.H.J., & Smits, A.J. 2008.<br />

The wake structure <strong>and</strong> thrust performance<br />

of a rigid low-aspect-ratio pitching panel.<br />

J Fluid Mech. 603:331-65. doi: 10.1017/<br />

S0022112008000906.<br />

Clark, R.P., & Smits, A.J. 2006. Thrust<br />

production <strong>and</strong> wake structure of a<br />

batoid-inspired oscillating fin. J Fluid<br />

Mech. 562:415-29. doi: 10.1017/<br />

S0022112006001297.<br />

Colgate, J.E., & Lynch, K.M. 2004. Mechanics<br />

<strong>and</strong> control of swimming: A review.<br />

IEEE J Oceanic Eng. 29(3):660-73.<br />

doi: 10.1109/JOE.2004.833208.<br />

Dewey, P.A., Carriou, A., & Smits, A.J.<br />

2011. On the relationship between efficiency<br />

<strong>and</strong> wake structures of a batoid-inspired<br />

oscillating fin. J Fluid Mech. Under review.<br />

Dong, H., Mittal, R., & Najjar, F.M. 2006.<br />

Wake topology <strong>and</strong> hydrodynamic performance<br />

of low-aspect-raio flapping foils.<br />

J Fluid Mech. 566:309-43. doi: 10.1017/<br />

S002211200600190X.<br />

Fish, F.E. 2010. Private Communication.<br />

Fish, F.E., Fegely, J.F., & Xanthopoulos,<br />

C.J. 1991. Burst-<strong>and</strong>-coast swimming in<br />

schooling fish (Notemigonus crysoleucas) with<br />

implications for energy economy. Comp<br />

Biochem Phys A Phys. 100:633-7.<br />

doi: 10.1016/0300-9629(91)90382-M.<br />

Fish, F.E., Haj-Hariri, H., Smits, A.J.,<br />

Bart-Smith, H., & Iwasaki, T. 2011.<br />

Biomimetic swimmer inspired by the<br />

manta ray. In: <strong>Biomimetics</strong>: Nature-Based<br />

Innovation, ed. Bar-Cohen, Y., 495-523.<br />

Boca Raton, FL: CRC Press.<br />

Fish, F.E., Lauder, G.V., Mittal, R., Techet,<br />

A.H., Triantafyllou, M.S., Walker, J.A., &<br />

Webb, P.W. 2003. Conceptual design for the<br />

construction of a biorobotic AUV based on<br />

biological hydrodynamics. In: Proceedings<br />

of the 13th International Symposium on<br />

Unmanned Untethered Submersible<br />

<strong>Technology</strong>. Durham, NH: Autonomous<br />

Systems Undersea Institute.<br />

Heathcote, S., & Gursul, I. 2007. Jet<br />

switching phenomenon for a periodically<br />

plunging airfoil. Phys Fluids. 19(2):027104.<br />

doi: 10.1063/1.2565347.<br />

Heathcote, S., Wang, Z., & Gursul, I. 2006a.<br />

Effect of spanwise flexibility on flapping wing<br />

propulsion. In: 36th AIAA Fluid Dynamics<br />

Conference <strong>and</strong> Exhibit. Reston, VA: American<br />

Institute of Aeronautics <strong>and</strong> Astronautics Inc.<br />

Heathcote, S., Wang, Z., & Gursul, I.<br />

2006b. Flexible flapping airfoil propulsion<br />

at low Reynolds numbers. AIAA J. 45(5):<br />

1066-79. doi: 10.2514/1.25431.<br />

Heine, C. 1992. Mechanics of flapping fin<br />

locomotion in the cownose ray, Rhinoptera<br />

bonasus (Elasmobranchii, Myliobatidae).<br />

Ph.D. Thesis Duke University.<br />

Jones, K.D., Dohring, C.M., & Platzer,<br />

M.F. 1998. Experimental <strong>and</strong> computational<br />

investigation of the KnollerBetz effect.<br />

Zool Anz. 173:1240-6.<br />

Klausewitz, W. 1964. Der lokomotionsmodus<br />

der Flugelrochen (Myliobatoidei). AIAA J.<br />

37:1240-6.<br />

Leftwich, M.C., & Smits, A.J. 2010. Thrust<br />

production by a mechanical swimming<br />

lamprey. Exp Fluids. 50(5):1349-55.<br />

doi 10.1007/s00348-010-0994-x.<br />

Liao, J.C. 2004. Neuromuscular control<br />

of trout swimming in a vortex street: Implications<br />

for energy economy during the<br />

Karman gait. J Exp Biol. 207:3495-506.<br />

doi: 10.1242/jeb.01125.<br />

Liao, J.C., Beal, D.N., Lauder, G.V., &<br />

Triantafyllou, M.S. 2003. Fish exploiting<br />

vortices decrease muscle activity. Science.<br />

302:1566-9. doi: 10.1126/science.1088295.<br />

Rosenberger, L.J. 2001. Pectoral fin locomotion<br />

in batoid fishes: Undulation versus<br />

oscillation. J Exp Biol. 204(2):379-94.<br />

Schmidt-Nielsen, K. 1972. Locomotion:<br />

Energy cost of swimming, flying, <strong>and</strong> running.<br />

Science. 177:222-8. doi: 10.1126/science.<br />

177.4045.222.<br />

Taylor, G.K., Nudds, R.L., & Thomas, A.L.R.<br />

2003. Flying <strong>and</strong> swimming animals cruise<br />

at a Strouhal number tuned for high power<br />

efficiency. Nature. 425(6959):707-11.<br />

doi: 10.1038/nature02000.<br />

118 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Inspired by Sharks: A Biomimetic Skeleton<br />

for the Flapping, Propulsive Tail of an<br />

Aquatic Robot<br />

AUTHORS<br />

John H. Long, Jr.<br />

Department of Biology,<br />

Vassar College<br />

Tom Koob<br />

MiMedx Group, Inc.<br />

Justin Schaefer<br />

David Geffen School of Medicine,<br />

University of California<br />

Adam Summers<br />

Friday Harbor Labs,<br />

University of Washington<br />

Kurt Bantilan<br />

Department of Biology,<br />

Vassar College<br />

Sindre Grotmol<br />

Department of Biology,<br />

University of Bergen<br />

Marianne Porter<br />

Department of Biology,<br />

Vassar College<br />

Propulsive Functions of<br />

Vertebral Columns<br />

I<br />

n sharks <strong>and</strong> other fish, the<br />

body’s primary skeleton is the vertebral<br />

column, which runs from the<br />

head to the caudal fin (Summers &<br />

Long, 2006). The vertebral column is<br />

a jointed framework to which muscles<br />

attach <strong>and</strong> on which the muscles pull<br />

to create the traveling waves of flexure<br />

that transfer momentum from the<br />

body to the surrounding fluid. The<br />

vertebral column is composed of rigid<br />

ABSTRACT<br />

The vertebral column is the primary stiffening element of the body of fish. This<br />

serially jointed axial support system offers mechanical control of body bending<br />

through kinematic constraint <strong>and</strong> viscoelastic behavior. Because of the functional<br />

importance of the vertebral column in the body undulations that power swimming,<br />

we targeted the vertebral column of cartilaginous fishes—sharks, skates, <strong>and</strong> rays—<br />

for biomimetic replication. We examined the anatomy <strong>and</strong> mechanical properties of<br />

shark vertebral columns. Based on the vertebral anatomy, we built two classes of<br />

biomimetic vertebral column (BVC): (1) one in which the shape of the vertebrae<br />

varied <strong>and</strong> all else was held constant <strong>and</strong> (2) one in which the axial length of the<br />

invertebral joint varied <strong>and</strong> all else was held constant. Viscoelastic properties of<br />

the BVCs were compared to those of sharks at physiological bending frequencies.<br />

The BVCs with variable joint lengths were then used to build a propulsive tail, consisting<br />

of the BVC, a vertical septum, <strong>and</strong> a rigid caudal fin. The tail, in turn, was<br />

used as the propeller in a surface-swimming robot that was itself modeled after a<br />

biological system. As the BVC becomes stiffer, swimming speed of the robot increases,<br />

all else being equal. In addition, stiffer BVCs give the robot a longer stride<br />

length, the distance traveled in one cycle of the flapping tail.<br />

Keywords: biomimetics, robot, vertebral column, propulsion<br />

elements, called vertebrae, connected<br />

by flexible intervertebral joints<br />

(Grotmol et al., 2003; Koob &<br />

Long, 2000). The joints <strong>and</strong> their<br />

adjoining vertebrae limit the body’s kinematic<br />

degrees of freedom, constraining<br />

bending primarily to the lateral<br />

direction in response to loads imposed<br />

by muscle, inertia, <strong>and</strong> external fluid<br />

forces (Grotmol et al., 2006; Porter<br />

et al., 2009; Symmons, 1979; Schmitz,<br />

1995). In this way, the vertebral<br />

column functions to control dynamic<br />

reconfigurations of the self-propelling<br />

body.<br />

In roll-stable sharks <strong>and</strong> fish, lateral<br />

body bending is characterized by the<br />

overlay of harmonic <strong>and</strong> transient motions<br />

that range from small traveling<br />

flexures to large-amplitude st<strong>and</strong>ing<br />

waves (Long et al., 2010). These bending<br />

motions produce the propulsive<br />

forces that create forward swimming,<br />

turning maneuvers, <strong>and</strong> rapid accelerations.<br />

Across many different kinds of<br />

fish, the stiffness of the bending joints,<br />

measured as the apparent Young’smodulus,<br />

E, rangesfrom0.1to8MPa<br />

(Long et al., 2002). The E is a mechanical<br />

property, sometimes called<br />

the “material stiffness,” that measures<br />

the contribution of the material, independent<br />

of its geometric arrangement<br />

in the structure, to the structure’s resistance<br />

to changing shape when an external<br />

load is applied to it.<br />

July/August 2011 Volume 45 Number 4 119


Joints with any appreciable stiffness<br />

at all may at first seem to be a paradox:<br />

why not have low-stiffness joints that<br />

cost little, in terms of mechanical<br />

work, to bend The answer seems to<br />

be two-fold: (1) Stiff joints increase<br />

their resistance, in terms of the absolute<br />

bending moment, M (in units of<br />

Nm), in proportion to the magnitude<br />

of bending curvature, κ (m −1 ). This resistance,<br />

which can grow nonlinearly<br />

with κ, serves as a brake, limiting<br />

lateral bending (Long et al., 2002).<br />

(2) Stiff joints store <strong>and</strong> release more<br />

mechanical work, so-called “elastic<br />

energy,” than flexible joints (Long,<br />

1992). The amount of work released<br />

in elastic recoil is also in proportion<br />

to E <strong>and</strong> to the square of κ. Hence,<br />

the vertebral column functions best<br />

as a spring when muscles have reached<br />

their functional limits, at the end of a<br />

FIGURE 1<br />

large-κ bend when connective tissues<br />

are stiffest.<br />

To explore the mechanical design<br />

space of vertebral columns, we created<br />

biomimetic vertebral columns (BVCs).<br />

The BVCs are modeled after the vertebral<br />

column of sharks. We chose<br />

sharks’ vertebral columns since they<br />

are structurally simple, compared to<br />

those of bony fish, consisting of cylindrical<br />

centra, small neural <strong>and</strong> hemal<br />

arches, <strong>and</strong> thin intervertebral joints<br />

(Figure 1). Together, a centrum <strong>and</strong><br />

its arches are called a vertebra, <strong>and</strong> in<br />

the cartilaginous sharks, skates, <strong>and</strong><br />

rays, the vertebrae (plural form of<br />

‘vertebra’) are composed of mineralized<br />

cartilage. The compressive stiffness<br />

of these vertebrae, measured by<br />

E, ranges from 25 to 500 MPa, overlapping<br />

the lower range of E for bone<br />

(Porter et al., 2006). Compared to the<br />

Vertebral columns of two species of sharks. For scale, each centrum shown here is between 4- <strong>and</strong><br />

6-mm long in the head-to-tail direction. Head is to the left; tail is to the right.<br />

bone of mammals, for a given value of E<br />

the vertebrae of sharks are stronger, where<br />

strength is measured in terms of breaking<br />

stress (Porter & Long, 2010). Thus,<br />

in some ways, vertebral columns made of<br />

mineralized cartilage perform better than<br />

vertebral columns made of bone.<br />

In summary, vertebral columns serve<br />

at least three important propulsive functions<br />

during swimming: (1) they control<br />

dynamic reconfigurations of the body<br />

by limiting the kinematic degrees of<br />

freedom, (2) they brake high-amplitude<br />

bends by virtue of their stiffness, <strong>and</strong><br />

(3) they integrate muscle work over<br />

time by recoiling elastically.<br />

To build a BVC that can function as<br />

part of an aquatic propulsion system,<br />

we (1) characterized the morphology<br />

(size <strong>and</strong> shape) of the vertebral columns<br />

of sharks, (2) measured the mechanical<br />

properties of those vertebral<br />

columns as they underwent sinusoidal<br />

bending, (3) used that information<br />

about morphology <strong>and</strong> mechanical<br />

properties to design BVCs, (4) tested<br />

BVCs as they underwent the dynamic<br />

bending characteristic of swimming<br />

<strong>and</strong> propulsion, <strong>and</strong> (5) tested the<br />

BVCs as the primary skeleton in the<br />

flapping tail of an aquatic robot.<br />

Morphology of Sharks’<br />

Vertebral Columns<br />

In three individuals of the blacktip<br />

shark, Carcharinus limbatus, <strong>and</strong><br />

the bonnethead shark, Sphryna<br />

tiburo, we measured, from radiographs,<br />

the following features from<br />

the head to the beginning of the caudal<br />

fin (Figure 2): (1) the length of centrum,<br />

c, (2), the diameter of the centrum,<br />

d, (3) the length of the intervertebral<br />

joint, j, <strong>and</strong> (4) the cone angle, Ξ,<br />

of the capsule of the joint. Blacktip<br />

sharks, members of the family<br />

Carcharhinidae, were chosen because<br />

120 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Measuring vertebral morphology of blacktip <strong>and</strong> bonnethead sharks. Representative X-rays<br />

show the heavily mineralized vertebral centra, which possess an “X” shapeinthistwo-dimensional<br />

view that is from cone-shaped joint capsules. The dark space between vertebrae is the intervertebral<br />

joint. The morphology of each vertebra <strong>and</strong> intervertebral joint was measured from digitized l<strong>and</strong>marks<br />

(blue dots). (Color versions of figures available online at: http://www.ingentaconnect.com/<br />

content/mts/mtsj/2011/00000045/00000004.)<br />

they are known to be fast swimming<br />

predators of fish. In contrast,<br />

bonnethead sharks, members of the<br />

hammerhead family Sphyrnidae, are<br />

known less for their speed <strong>and</strong> more<br />

for their maneuverability <strong>and</strong> ability<br />

to find <strong>and</strong> eat crustaceans. Of similar<br />

adult body size, the two species<br />

represent contrasting swimming styles<br />

<strong>and</strong> ecologies. Three individuals of<br />

each species were used for this study.<br />

In the bonnethead shark, all the<br />

morphological features, except d,<br />

increased in size from the head to the<br />

end of the abdomen <strong>and</strong> then decreased<br />

towards the caudal fin (Figure 3). In<br />

blacktip shark, only Ξ <strong>and</strong> c varied<br />

from head to caudal fin. The significance<br />

(α = 0.05) of the morphological<br />

variation was determined with a<br />

multivariate analysis of covariance<br />

(MANCOVA), with species <strong>and</strong> position<br />

as main effects <strong>and</strong> individual as<br />

the covariate (JMP 8.0.2., SAS Institute,<br />

Cary, NC). Following an identity<br />

model MANCOVA, univariate<br />

ANCOVAs were also run.<br />

The variation in the morphology of<br />

these vertebral columns was used to<br />

guide the construction of the BVCs.<br />

For each of the sharks’ morphological<br />

features, we indicated which dimensions<br />

were used (see black arrows on<br />

the ordinates, Figure 3).<br />

Mechanical Properties of<br />

Sharks’ Vertebral Columns<br />

Using the same freshly dissected<br />

vertebral columns from which morphology<br />

was measured, we conducted<br />

3-point dynamic bending tests using<br />

an MTS model Mini Bionix 858<br />

(Eden Prairie, MN) with a 500 N<br />

load cell. For blacktip sharks, each<br />

vertebral column was cut into five<br />

segments of 19 vertebrae each. For<br />

bonnethead sharks, each vertebral<br />

column was cut into five segments of<br />

14 vertebrae each. The number of segments<br />

in each species was varied to<br />

keep the absolute length of each test<br />

segment approximately equal. Each<br />

segment was subjected to sinusoidal<br />

bending at a frequency, f (Hz), <strong>and</strong><br />

maximum curvature, κ (m −1 ), varied<br />

to hold constant the time rate of<br />

change of κ, which is equivalent to<br />

the strain rate (actuator displacement<br />

amplitude of 2 mm s −1 ).<br />

To characterize the viscoelastic<br />

properties of the vertebral column during<br />

bending, the apparent storage <strong>and</strong><br />

loss moduli, E′ <strong>and</strong> E″ (MPa), respectively,<br />

were measured at each combination<br />

of two species, five segment<br />

positions, <strong>and</strong> three κ. TheE′ measures<br />

the purely elastic component of<br />

the stiffness; it is the force proportional<br />

to the magnitude of the bending of the<br />

vertebral column. The E″ measures the<br />

purely viscous component of the stiffness;<br />

it is the force proportional to the<br />

velocity of the bending of the vertebral<br />

column. These properties were calculated<br />

from the following formulae:<br />

E ′ ¼ E*cos δ <strong>and</strong> E ″ ¼ E*sin δ,<br />

wherein E* ¼ F maxL 3<br />

48Iy max<br />

<strong>and</strong> δ is the<br />

phase lag (radians) between the displacement<br />

<strong>and</strong> load signals. Moreover,<br />

F max is the force (N) measured at the<br />

load cell, L is the gauge length of the<br />

specimen (m), I is the specimen’s<br />

July/August 2011 Volume 45 Number 4 121


FIGURE 3<br />

Vertebral morphology of sharks. Four dimensions were used to characterize the size <strong>and</strong> shape of<br />

the vertebral centra <strong>and</strong> the intervertebral joints. The means of three individuals for each species<br />

are shown; individuals ranged from 0.59 to 0.91 m in overall body length. The error bars indicate<br />

the st<strong>and</strong>ard error of the mean. Black arrows show the specific dimensions represented in our<br />

BVCs. MANCOVA, using the identity method, calculated a significant Wilkes λ (p < 0.0001),<br />

with a significant interaction of species <strong>and</strong> position <strong>and</strong> significant main effect of species; the<br />

covariate, individual, was also significant. Partial correlations among the response variables<br />

ranged from a low of 0.38 between j <strong>and</strong> d toahighof0.79betweenΞ <strong>and</strong> c. Significance<br />

level is indicated (*p < 0.05, **p < 0.01, ***p < 0.001).<br />

second moment of area (m 4 ), <strong>and</strong> y max<br />

(m) is the distance from the presumed<br />

neutral plane of bending (transverse<br />

center of specimen) <strong>and</strong> the lateralmost<br />

fibers of the specimen.<br />

In blacktip sharks, the E′ <strong>and</strong> E″<br />

values were of greater magnitude<br />

( p < 0.05) than those of bonnethead<br />

sharks (Figure 4). In both species,<br />

E′ increased towards the tail, an effect<br />

that is amplified at higher values of κ,<br />

as indicated by a significant (p < 0.05)<br />

interaction term. The significance of<br />

the variation in E′ <strong>and</strong> E″ was determined<br />

using ANOVA, with species,<br />

position, <strong>and</strong> κ as main effects (JMP<br />

8.0.2., SAS Institute, Cary, NC).<br />

Since the data blacktip <strong>and</strong> bonnethead<br />

sharks were taken at a single amplitude<br />

of strain rate (2.0 mm s −1 ), we<br />

sought additional information about<br />

how E′ <strong>and</strong> E″ vary with changes in<br />

strain rate. We also wanted to test the<br />

hypothesis that the intervertebral capsule,<br />

which contains liquid under<br />

above-ambient pressure, uses its internal<br />

fluid pressure to alter the apparent<br />

E′ <strong>and</strong> E″ of the vertebral column. Because<br />

blacktip <strong>and</strong> bonnethead sharks<br />

were not available for these tests,<br />

spiny dogfish, Squalus acanthias, were<br />

used. Fresh 10-vertebrae segments<br />

were removed from the region of the<br />

first dorsal fin inthreedogfish. Each<br />

segment was pressure-clamped at the<br />

terminal vertebrae <strong>and</strong> end-loaded<br />

with bending moments, M (for experimental<br />

configuration, see Long et al.,<br />

2011). The bending motion was delivered<br />

via moment arms attached to a<br />

single-axis linear actuator using an<br />

MTS model Tytron 250 (Eden Prairie,<br />

MN) <strong>and</strong> a 50-N load cell. To test the<br />

effects of both f <strong>and</strong> κ on E′ <strong>and</strong> E″,<br />

each segment was bent sinusoidally at<br />

each combination of five f values <strong>and</strong><br />

three κ values. In addition, to test the<br />

effects of the integrity of the fluid-filled<br />

intervertebral joint capsule on E′ <strong>and</strong><br />

E″, we repeated this suite of tests<br />

after (a) puncturing a single joint capsule<br />

located in the middle of the segment<br />

<strong>and</strong> (b) puncturing three joint<br />

capsules, including the first one punctured<br />

<strong>and</strong> two adjoining capsules.<br />

Increases in f increased only E′<br />

( p < 0.05) while increases in κ<br />

increased both E′ <strong>and</strong> E″ (Figure 5).<br />

The only significant effect of puncturing<br />

the intervertebral capsule was<br />

when three capsules were punctured,<br />

<strong>and</strong> even then only E″ increased. The<br />

122 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 4<br />

Mechanical properties of the vertebral columns of sharks in sinusoidal bending. Points are means<br />

from three individuals. Error bars are the st<strong>and</strong>ard error of the mean. Size of the symbol indicates<br />

the relative magnitude of the curvature, κ. Significance level is indicated (n.s. = not significant,<br />

*p < 0.05, **p < 0.01, ***p < 0.001).<br />

significance of the variation in E′ <strong>and</strong><br />

E″ was determined using ANCOVA,<br />

with puncture, f, <strong>and</strong> κ as main effects<br />

<strong>and</strong> individual as the covariate (JMP<br />

8.0.2., SAS Institute, Cary, NC).<br />

In summary, the vertebral columns<br />

of sharks have mechanical properties<br />

that are highly variable. As species<br />

<strong>and</strong> anatomical position change, so,<br />

too, do E′ <strong>and</strong> E″. Within a given vertebral<br />

segment, the apparent storage<br />

modulus, E′, <strong>and</strong> the apparent loss<br />

modulus, E″, can be altered by the<br />

bending that they undergo. Increasing<br />

the segment’s curvature,κ, increases<br />

both E′ <strong>and</strong> E″; increasing the segment’s<br />

frequency of bending, f, increases the<br />

E′. Knowing the mechanical behavior<br />

of shark vertebral columns under realistic<br />

bending conditions creates specifications<br />

for BVCs.<br />

Designing BVCs<br />

To begin to underst<strong>and</strong> how to<br />

control the mechanical behavior of<br />

BVCs, we built two classes of sharkinspired<br />

BVC: (1) BVC with variable<br />

cone angle, Ξ (BVC Ξ ): vertebrae<br />

were created with variable Ξ <strong>and</strong> the<br />

BVC had constant joint length, j,<br />

<strong>and</strong> (2) BVC with variable joint length,<br />

j (BVC j ): vertebrae were created with<br />

aconstantΞ <strong>and</strong> the BVC had variable<br />

j. In addition to exploring the effects<br />

of the structures Ξ <strong>and</strong> j, we<br />

also varied the amount of cross-linking<br />

of the hydrogel material forming the<br />

joint. Thus, we explored the BVC<br />

“morphospace,” the variety of designs<br />

described by three dimensions: Ξ, j,<br />

<strong>and</strong> cross-linking. Part of this exploration<br />

involved the challenge of making<br />

composite structures that concatenate<br />

flexible <strong>and</strong> rigid elements. After fabrication<br />

<strong>and</strong> mechanical testing of both<br />

classes of BVC, we selected a single<br />

class, the BVC j , for performance testing<br />

in a tail-flapping aquatic robot.<br />

In the BVC Ξ , vertebrae were designed<br />

in software (SolidWorks,<br />

Dassault Systèmes SolidWorks Corp.,<br />

Concord, MA) to have the following<br />

values of Ξ: 15°, 30°, <strong>and</strong> 45° (Figure<br />

6). These values correspond to<br />

low, medium, <strong>and</strong> high values of Ξ<br />

measured in sharks (see Figure 3).<br />

The diameter, d, <strong>and</strong>axiallength,c,<br />

of the vertebrae were fixed at 1 cm<br />

for both. The j of the column was<br />

fixed at 0.25 cm.<br />

Vertebrae were fabricated with a<br />

rapid prototyper (Z-Corp, model<br />

310), which produced a porous, plaster<br />

part that was subsequently infiltrated<br />

with cyanoacrylate (EZ bond<br />

5cps, K&R International, Diamond<br />

Bar, CA). This process yielded vertebrae<br />

with mean compressive moduli,<br />

E (MPa) of 43, 50, <strong>and</strong> 61 for vertebrae<br />

with values of Ξ at 15°, 30°,<br />

<strong>and</strong> 45°, respectively. These values of<br />

E are within the range measured for<br />

shark vertebrae (Porter et al., 2006).<br />

Vertebrae of a given Ξ were assembled<br />

into a BVC Ξ in two stages. First,<br />

seven vertebrae were linked together,<br />

spaced at the fixed j, with eight horse<br />

hairs (E in tension of 900 MPa)<br />

arrayedaxially<strong>and</strong>affixed to the<br />

outer circumference of the vertebrae.<br />

These horse hairs served as first<br />

July/August 2011 Volume 45 Number 4 123


FIGURE 5<br />

Mechanical properties of the vertebral column vary as a function of cycle frequency, f, <strong>and</strong> the integrity<br />

of the intervertebral joint in the spiny dogfish, Squalus acanthias. Points are means from<br />

three individuals. Error bars are the st<strong>and</strong>ard error of the mean. Size of the symbol indicates<br />

the relative magnitude of the curvature, κ. Significance level is indicated (n.s. = not significant,<br />

*p < 0.05, **p < 0.01, ***p < 0.001).<br />

10% porcine gelatin fixed in 2.5%<br />

glutaraldehyde (Long et al., 2006).<br />

Vertebrae were slid onto the hydrogel,<br />

spaced evenly at the desired j, <strong>and</strong> affixed<br />

to the hydrogel with cyanoacrylate<br />

adhesive. A total of 12 different<br />

types of BVC j were produced, one<br />

type for each of the 12 different values<br />

of j. Three replicates of each type were<br />

produced <strong>and</strong> tested. Please note that<br />

in the BVC j horse hairs were omitted<br />

because at all but the smallest values<br />

of j, the hairs cut into the hydrogel during<br />

bending.<br />

approximations of the intervertebral<br />

ligaments found in the vertebral columns<br />

of sharks. Second, a 10% porcine<br />

gelatin solution was injected in<br />

between the vertebrae; the gelatin was<br />

solidified at 4° C. Once solidified, each<br />

BVC Ξ was then subjected to one<br />

of three fixation treatments: 0, 1%,<br />

or 5 % glutaraldehyde, a chemical<br />

agent that cross-links the collagen in<br />

the hydrogel. A total of nine different<br />

types of BVC Ξ were produced, with<br />

each possible pairwise combination of<br />

Ξ <strong>and</strong> glutaraldehyde concentration.<br />

In the BVC j , vertebrae were designed<br />

to have a Ξ of 90°, which created<br />

ring-shaped vertebrae (Figure 7).<br />

The d <strong>and</strong> c of the vertebrae were<br />

fixed at 0.5 <strong>and</strong> 1.0 cm, respectively.<br />

The overall length of the BVC j was<br />

fixed at 8.4 cm. As the number of<br />

nonterminal vertebrae were varied<br />

from 0 to 11, j varied from 720 to<br />

0.5 mm. These values of j created a<br />

range that extended below <strong>and</strong> above<br />

the range of j measured in sharks (see<br />

Figure 3).<br />

Vertebrae were milled from<br />

Delrin, a polyoxymethylene thermoplastic.<br />

Delrin has a compressive E<br />

of 3.1 GPa (Delrin Design Guide,<br />

Module III, from DuPont), which<br />

liesinthemiddleoftherangeof<br />

E values reported for shark vertebrae<br />

(Porter et al., 2006).<br />

The ring vertebrae had an inner diameter<br />

of 0.8 cm, which matched the<br />

outer diameter of hydrogels made from<br />

Mechanical Properties<br />

of BVCs<br />

The E′ <strong>and</strong> E″ of the BVCs were<br />

measured in two different kinds of<br />

sinusoidal bending test, which corresponded<br />

to the tests performed on<br />

sharks’ vertebral columns. In the<br />

BVC Ξ , 3-point bending tests were<br />

conducted in a manner identical with<br />

those on the blacktip <strong>and</strong> bonnethead<br />

sharks. In the BVC j , end-loaded bending<br />

tests were conducted in a manner<br />

identical with those on the spiny dogfish<br />

sharks. The E′ <strong>and</strong> E″ data for the<br />

BVC j have been analyzed previously<br />

(Long et al., 2011). In the analysis<br />

here, the data have been reanalyzed<br />

to calculate the mechanical work required<br />

to bend the BVC j <strong>and</strong> the mechanical<br />

work recovered as recoil.<br />

In the BVC Ξ , both E′ <strong>and</strong> E″ increased<br />

as the glutaraldehyde concentration<br />

increased, E′ <strong>and</strong> E″ decreased<br />

as the Ξ increased, <strong>and</strong> E′ increased<br />

<strong>and</strong> E″ decreased as κ increased (Figure8).Thesignificance<br />

of the variation<br />

in E′ <strong>and</strong> E ″ was determined<br />

using ANOVA, with glutaraldehyde<br />

concentration, Ξ, <strong>and</strong>κ. asmain<br />

effects (JMP 8.0.2., SAS Institute,<br />

Cary, NC).<br />

124 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 6<br />

BVCs (BVC Ξ ) with variable cone angles, Ξ, <strong>and</strong> constant joint length, j.<br />

FIGURE 7<br />

BVCs (BVC j ) with variable joint lengths, j, <strong>and</strong> constant cone angle, Ξ.<br />

Compared to the mechanical properties<br />

of shark vertebral columns, the<br />

BVC Ξ have values of E′ that have a<br />

wider range, overlapping the lower<br />

values <strong>and</strong> exceeding the sharks’ higher<br />

values by an order of magnitude (compare<br />

Figures 8 <strong>and</strong> 4). In contrast, the<br />

E″ values of the BVC Ξ overlap only<br />

with those of the bonnethead shark;<br />

the BVC Ξ has much lower values of<br />

E″ than either the blacktip or spiny<br />

dogfish shark. Moreover, the E′ for<br />

BVC Ξ decreases as κ increases; we<br />

measured the opposite trend in sharks<br />

(see Figures 4 <strong>and</strong> 5). Hence, the<br />

BVC Ξ is not a good biological model<br />

in this sense. Our hypothesis as to the<br />

source of this strain softening is that<br />

the horse hairs force the column to<br />

bend primarily by compression, rather<br />

than by a combination of tension <strong>and</strong><br />

compression.<br />

In the BVC j ,bothE′ <strong>and</strong> E″ decreased<br />

nonlinearly as j increased<br />

(Figure 9). Compared to the mechanical<br />

properties of dogfish vertebral<br />

columns, the BVC j span a nearly identical<br />

range of E′ <strong>and</strong> E″ values. The<br />

greatest sensitivity to changes in j occurred<br />

at the smallest values of j (Figure<br />

9) in the region that corresponds<br />

to the j measured in the vertebral columns<br />

of sharks (Figure 3). In data<br />

shown elsewhere (Long et al., 2011),<br />

E ′ <strong>and</strong> E ″ of the BVC j increased<br />

with increasing κ, just as in sharks<br />

(see Figure 5 herein). Moreover the E′<br />

increased with increasing f, aslikewise<br />

seen in sharks (Figure 5).<br />

The mechanical work to bend the<br />

BVC j increased with increasing κ <strong>and</strong><br />

increasing E′ (Figure 9). The mechanical<br />

work recovered as elastic recoil,<br />

W recoil , is a function of the resilience,<br />

R, which, over all the testing<br />

conditions <strong>and</strong> sizes of joints, averaged<br />

76%.<br />

BVCs in Aquatic Vehicles<br />

The flexible skeletons of sharks <strong>and</strong><br />

fish are inspiring the design of novel<br />

propulsive systems <strong>and</strong> aquatic vehicles<br />

(for review, see Fish, 2006; Long,<br />

2007, 2011). Fins with life-like flexibility<br />

have been built to propel a 0.7 m<br />

long robotic turtle (Long et al., 2006)<br />

<strong>and</strong>a0.4mlongroboticknifefish<br />

(Curet et al., 2011). Bodies with<br />

life-like flexibility have been built to<br />

July/August 2011 Volume 45 Number 4 125


FIGURE 8<br />

The mechanical properties of the BVCs (BVC Ξ ) with variable cone angles, Ξ, <strong>and</strong> constant joint<br />

length, j. Horizontal bars indicate the median, the lower <strong>and</strong> upper limits of the box indicate the<br />

25th <strong>and</strong> 75th percentiles, respectively, <strong>and</strong> the whiskers indicate the range.<br />

FIGURE 9<br />

The mechanical properties of the BVCs (BVC j ) with variable joint length, j, <strong>and</strong> constant cone<br />

angle, Ξ. Top row: points represent the means of E′ <strong>and</strong> E″ pooled across f <strong>and</strong> κ; error bars<br />

are one st<strong>and</strong>ard error of the mean. Bottom row: points are not pooled.<br />

propel a 0.7-m long robotic electric<br />

ray (Krishnamurthy et al., 2010),<br />

a 0.5-m long robotic trout (Kruusmaa<br />

et al., 2011), a 0.12-m long mechanical<br />

sunfish (McHenry et al., 1995),<br />

<strong>and</strong> a 0.5-m long mechanical pickerel<br />

(Conte et al., 2010). Of these<br />

self-propelled aquatic vehicles, only<br />

the mechanical pickerel has anything<br />

resembling a vertebral column:<br />

a piece of spring steel designed to<br />

release mechanical work to power<br />

accelerations.<br />

The BVC j presented here was invented<br />

to propel a surface swimming,<br />

0.3-m long tadpole robot (Long<br />

et al., 2006; Doorly et al., 2009),<br />

known as Tadro4 (Figure 10). BVC j<br />

were attached to a servo motor that<br />

created a sinusoidally varying pitching<br />

motion of a tail. That pitch bent the<br />

BVC j , creating a bending moment<br />

that propagated down the length of<br />

the BVC j in a traveling wave that, in<br />

turn, oscillated the terminal caudal<br />

fin. In this configuration, without<br />

distributed muscles, the BVC j acts<br />

as both a transmission system, transferring<br />

momentum from the servo motor<br />

to the caudal fin, <strong>and</strong> as a propeller,<br />

directly transferring momentum to<br />

the surrounding fluid.<br />

Since Tadro4 was built to behave<br />

reactively, with sensorimotor feedback<br />

systems creating foraging <strong>and</strong> predator<br />

avoidance, we needed a version that<br />

could be programmed to swim straight<br />

using a constant flapping frequency of<br />

the tail, f, <strong>and</strong> lateral amplitude of the<br />

caudal fin. That modified version of<br />

Tadro4 was called MARMT (Mobile<br />

Autonomous Robot for Mechanical<br />

Testing), <strong>and</strong> it had a hull length of<br />

17 cm <strong>and</strong> a tail length of 10 cm<br />

(Long et al., 2011).<br />

Outfitted with a given BVC j ,<br />

MARMT’s steady swimming performance<br />

was measured as swimming<br />

126 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 10<br />

The aquatic robot, Tadro4, is propelled by a BVC (BVC j ). Tadro4 is a fully-autonomous surfaceswimmer<br />

with a flattened circular body <strong>and</strong> propulsive undulatory tail. It is modeled after fish like<br />

the extinct Drepanaspis <strong>and</strong> the living electric ray, Narcine. Using sensory input from photoresistors<br />

<strong>and</strong> IR proximity detectors, Tadro4 searches for <strong>and</strong> swims up light gradients while avoiding<br />

collisions. Tadro4 is propelled by its submerged BVC j , which is wrapped in a thin membrane, attached<br />

to a caudal fin, <strong>and</strong> actuated by an oscillating servo motor. Tadro4 was developed by Doorly<br />

et al. (2009). Photo of Drepanaspis specimen 8462, American Museum of Natural History. Photo<br />

of adult Narcine is courtesy of Dr. Steve Kajiura.<br />

of E′ (Figure 11, top panel). The stride<br />

length of MARMT increased initially,<br />

doubling as E′ doubled, before tapering<br />

off.<br />

When the BVC j operates in the<br />

flapping tail, MARMT’s swimming<br />

performance is clearly linked to the<br />

mechanical properties of the BVC j , E′<br />

in the case shown here. Those mechanical<br />

properties are, in turn, under<br />

the control of the structure of the<br />

BVC j . Thus these experiments, taken<br />

together, demonstrate the functional<br />

relationship between the structure of<br />

the BVC j <strong>and</strong> the performance of a<br />

self-propelled aquatic robot.<br />

speed, U, <strong>and</strong> stride length, the slope<br />

ofthelineofU regressed onto f,<br />

which measures the distance the<br />

robot travels over one period of the<br />

flapping tail (Figure 11). As f increased<br />

for any BVC j , so, too, did the U. The<br />

BVC j with greater values of E′ produced<br />

a more rapid increase in U,<br />

over the same range of change in f,<br />

compared to BVC j , with smaller values<br />

Summary<br />

Using the morphology <strong>and</strong> mechanical<br />

properties of the vertebral<br />

columns of sharks as our biological<br />

target, we built <strong>and</strong> tested a series of<br />

BVC. The mechanical behavior of<br />

the BVCs, measured by the storage<br />

<strong>and</strong> loss moduli over a range of bending<br />

frequencies <strong>and</strong> curvatures, can be<br />

altered by changing (1) the material<br />

properties of the hydrogel that makes<br />

up the intervertebral joint, (2) the<br />

length of the intervertebral joint, or<br />

(3) the shape of the vertebrae. BVCs<br />

are sufficient to function as propulsive<br />

elements in swimming aquatic robots:<br />

in Tadro4 <strong>and</strong> MARMT the<br />

BVC converts a simple pitch oscillation<br />

from a servo motor into a wave<br />

of bending that drives the caudal fin<br />

laterally.<br />

Having identified variables that influence<br />

the mechanical behavior of<br />

BVCs, we offer a few observations for<br />

those wishing to build jointed, flexible<br />

biomimetic skeletons for use in flexible,<br />

flapping propulsive systems:<br />

(1) Design of biomimetic systems:<br />

Engineered systems that are much<br />

July/August 2011 Volume 45 Number 4 127


FIGURE 11<br />

Swimming performance of a surface-swimming robot, MARMT, propelled by a tail with the BVC j as<br />

the primary skeleton. MARMT is a version of Tadro4 (Figure 10) modified for mechanical testing<br />

over a range of flapping frequencies of the tail, f (Hz). For all types of BVC j tested, swimming speed<br />

of MARMT, U, increased linearly with increases in f (all R 2 values > 0.92). The rate of change of<br />

U with respect to f is the stride length (distance traveled per period of the flapping cycle); it was<br />

greatest with BVC j having larger storage moduli, E′. Three replicates of each kind of BVC j were<br />

tested (N = 36). Means (N = 12) are shown here. Data reanalyzed from Long et al., 2011.<br />

functional morphology. Next, test<br />

the morphology’s mechanical behavior<br />

under physiologically relevant testing<br />

conditions. Finally, build <strong>and</strong> test<br />

simple biomimetic models of the system<br />

that change just a single structural<br />

variable over a wide range. Repeat this<br />

process with different variables, ceteris<br />

paribus, until the designer knows<br />

which variables permit the natural system<br />

<strong>and</strong> its operational range to be<br />

mimicked or extended in biomimetic<br />

form.<br />

Acknowledgments<br />

We thank Carl Bertsche, Nicole<br />

Doorly, Carina Frias, Andres Gutierrez,<br />

Jonathan Hirokawa, Kira Irving, Doug<br />

Pringle, Foster Ranney, Hannah<br />

Rosenblum, Hassan Sahktah, Sonia<br />

Roberts, Elise Stickles, Josh Sturm,<br />

<strong>and</strong> Janese Trimaldi for their help in<br />

designing, building, <strong>and</strong> testing vertebral<br />

columns, BVCs, <strong>and</strong> the aquatic<br />

robots. This work was supported by<br />

the National Science Foundation<br />

of the USA (DBI-0442269 <strong>and</strong><br />

IOS-0922605).<br />

simpler than the targeted biological<br />

system can match <strong>and</strong> extend<br />

the targeted range of mechanical<br />

behaviors.<br />

(2) Control of mechanical properties:<br />

The spacing of rigid elements<br />

in a flexible matrix is more important<br />

than the shape of the rigid elements<br />

or the material properties of<br />

the flexible material.<br />

(3) Control of reconfiguration: Because<br />

of the strain- <strong>and</strong> strain-ratedependence<br />

of viscoelastic materials,<br />

no passive, flexible propulsive system,<br />

if its E′ <strong>and</strong> E″ matches that<br />

of the vertebral column of sharks,<br />

will produce constant motions over<br />

a wide range of motor inputs.<br />

This work is a straight-forward example<br />

of one method of biomimetic<br />

design (Fish, 2006; Long, 2007):<br />

describe, test, build, <strong>and</strong> test. Start<br />

by identifying a specific operational<br />

context—aquatic undulatory propulsion<br />

in this case. Then describe,<br />

quantitatively, the biological system’s<br />

Lead Author:<br />

John H. Long, Jr.<br />

Department of Biology,<br />

Vassar College<br />

124 Raymond Avenue,<br />

Poughkeepsie, NY 12604-0513<br />

Email: jolong@vassar.edu<br />

References<br />

Conte, J., Modarres-Sadeghi, Y., Watts, M.N.,<br />

Hover, F.S., & Triantafyllou, M.S. 2010. A<br />

fast-starting mechanical fish that accelerates<br />

at 40 m s-2. Bioinspir Biomim. 5(3):1-9.<br />

doi: 10.1088/1748-3182/5/3/035004.<br />

Curet, O.M., Patankar, N.A., Lauder, G.V.,<br />

& MacIver, M.A. 2011. Aquatic manoeuvering<br />

with counter-propagating waves: A novel<br />

128 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


locomotive strategy. J R Soc Interface.<br />

8(60):1041-50. (10.1098/rsif.2010.0493).<br />

Doorly, N., Irving, K., McArthur, G.,<br />

Combie, K., Engel, V., Sakhtah, H., …<br />

Long, J.H., Jr. 2009. Biomimetic evolutionary<br />

analysis: Robotically-simulated vertebrates in<br />

a predator-prey ecology. In: Proc. 2009<br />

IEEE Symp. Artificial Life. 147-54. Nashville,<br />

TN: IEEE (Institute of Electrical <strong>and</strong><br />

Electronic Engineers). doi: 10.1109/<br />

ALIFE.2009.4937706.<br />

Fish, F.E. 2006. Limits of nature <strong>and</strong><br />

advances of technology: What does biomimetics<br />

have to offer to aquatic robots Appl.<br />

Bionics Biomech. 3(1):49-60. doi: 10.1533/<br />

abbi.2004.0028.<br />

Grotmol, S., Kryvi, H., Keynes, R., Krossoy,<br />

C., Nordvik, K., & Totl<strong>and</strong>, G.K. 2006.<br />

Stepwise reinforcement of the notochord<br />

<strong>and</strong> its intersection with the myoseptum:<br />

An evolutionary path leading to the development<br />

of the vertebra J Anat. 209:339-57.<br />

doi: 10.1111/j.1469-7580.2006.00618.x.<br />

Grotmol, S., Kryvi, H., Nordvik, K., &<br />

Totl<strong>and</strong>, G.K. 2003. Notochord segmentation<br />

may lay down the pathway for the<br />

development of the vertebral bodies in the<br />

Atlantic salmon. Anat Embryol. 207:263-72.<br />

doi: 10.1007/s00429-003-0349-y.<br />

Koob, T.J., & Long, J.H., Jr. 2000. The<br />

vertebrate body axis: Evolution <strong>and</strong> mechanical<br />

function. Am Zool. 40(1):1-18.<br />

doi: 10.1668/0003-1569(2000)040[0001:<br />

TVBAEA]2.0.CO;2.<br />

Krishnamurthy, P., Khorrami, F., de Leeuw,<br />

J., Porter, M., Livingston, K., & Long,<br />

J.H., Jr. 2010. An electric ray inspired<br />

biomimetic autonomous underwater vehicle.<br />

Am Control Conf. 2010:5224-9.<br />

Kruusmaa, M., Salumäe, T., Toming, G.,<br />

Ernits, A., & Ježov, J. 2011. Swimming speed<br />

control <strong>and</strong> on-board flow sensing of an<br />

artificial trout. In: Proceedings of IEEE<br />

International Conference of Robotics <strong>and</strong><br />

Automation (IEEE ICRA 2011). Shanghai,<br />

China: IEEE (Institute of Electrical <strong>and</strong><br />

Electronic Engineers).<br />

Long, J.H., Jr. 1992. Stiffness <strong>and</strong> damping<br />

forces in the intervertebral joints of blue marlin,<br />

(Makaira nigricans). J Exp Biol. 162:131-55.<br />

Long, J.H., Jr. 2007. Biomimetic robotics:<br />

Building autonomous, physical models to test<br />

biological hypotheses. Proc. Inst. Mech. Eng,<br />

C, J. Mech. Eng. Sci. 221:1193-200.<br />

Long, J.H., Jr. 2011. <strong>Biomimetics</strong>: Robotics<br />

based on fish swimming. In: Encyclopedia of<br />

Fish Physiology: From Genome to Environment,<br />

ed. Farrell, A.P., 1, 603-12. San Diego:<br />

Academic Press.<br />

Long, J.H., Jr., Koob-Emunds, M., Sinwell,<br />

B., & Koob, T.J. 2002. The notochord of<br />

hagfish, Myxine glutinosa: Viscoelastic properties<br />

<strong>and</strong> mechanical functions during steady<br />

swimming. J Exp Biol. 205:3819-31.<br />

Long, J.H., Jr., Koob, T.J., Irving, K.,<br />

Combie, K., Engel, V., Livingston, N., …<br />

Schumacher, J. 2006. Biomimetic evolutionary<br />

analysis: Testing the adaptive value of vertebrate<br />

tail stiffness in autonomous swimming<br />

robots. J Exp Biol. 209(23):4732-46.<br />

doi: 10.1242/jeb.02559.<br />

Long, J.H., Jr., Krenitsky, N., Roberts, S.,<br />

Hirokawa, J., de Leeuw, J., & Porter, M.E.<br />

2011. Testing biomimetic structures in<br />

bioinspired robots: How vertebrae control<br />

the stiffness of the body <strong>and</strong> the behavior<br />

of fish-like swimmers. Integr Comp Biol.<br />

51(1):158-75. doi: 10.1093/icb/icr020.<br />

Long, J.H., Jr., Porter, M.E., Liew, C.W., &<br />

Root, R.G. 2010. Go reconfigure: How fish<br />

change shape as they swim <strong>and</strong> evolve. Integr<br />

Comp Biol. 50(6):1120-39. doi: 10.1093/icb/<br />

icq066.<br />

McHenry, M.J., Pell, C.A., & Long, J.H., Jr.<br />

1995. Mechanical control of swimming speed:<br />

Stiffness <strong>and</strong> axial wave form in an undulatory<br />

fish model. J Exp Biol. 198:2293-305.<br />

Porter, M.E., Beltran, J.L., Koob, T.J., &<br />

Summers, A.P. 2006. Material properties<br />

<strong>and</strong> biochemical composition of mineralized<br />

vertebral cartilage in seven elasmobranch<br />

species (Chondrichthyes). J Exp Biol.<br />

209:2920-8. doi: 10.1242/jeb.02325.<br />

Porter, M.E., & Long, J.H., Jr. 2010.<br />

Vertebrae in compression: Mechanical<br />

behavior of arches <strong>and</strong> centra in the gray<br />

smooth-hound shark (Mustelus californicus).<br />

J Morph. 271:366-75.<br />

Porter, M.E., Roque, C.M., & Long, J.H., Jr.<br />

2009. Turning maneuvers in sharks: Predicting<br />

body curvature from body <strong>and</strong> vertebral<br />

morphology. J Morph. 270:954-65.<br />

doi: 10.1002/jmor.10732.<br />

Schmitz, R.J. 1995. Ultrastructure <strong>and</strong><br />

function of cellular components of the<br />

intercentral joint in the percoid vertebral<br />

column. J Morph. 226(1):1-24.<br />

doi: 10.1002/jmor.1052260102.<br />

Summers, A.P, & Long, J.H., Jr. 2006. Skin<br />

<strong>and</strong> bones, sinew <strong>and</strong> gristle: The mechanical<br />

behavior of fish skeletal tissues. In: Fish<br />

Biomechanics, eds. Shadwick, R.E., & Lauder,<br />

G.V., 141-77. San Diego: Academic Press.<br />

Symmons, S. 1979. Notochordal <strong>and</strong> elastic<br />

components of the axial skeleton of fishes<br />

<strong>and</strong> their functions in locomotion. J Zool.<br />

189(2):157-206. doi: 10.1111/j.1469-<br />

7998.1979.tb03958.x.<br />

July/August 2011 Volume 45 Number 4 129


PAPER<br />

Lateral-Line-Inspired Sensor Arrays for<br />

Navigation <strong>and</strong> Object Identification<br />

AUTHORS<br />

Vicente I. Fern<strong>and</strong>ez<br />

Audrey Maertens<br />

Department of Mechanical<br />

Engineering, Massachusetts<br />

Institute of <strong>Technology</strong> (MIT)<br />

Frank M. Yaul<br />

Department of Electrical<br />

Engineering <strong>and</strong> Computer<br />

Science, Massachusetts<br />

Institute of <strong>Technology</strong><br />

Jason Dahl<br />

Department of Mechanical<br />

Engineering, Massachusetts<br />

Institute of <strong>Technology</strong><br />

Jeffrey H. Lang<br />

Department of Electrical<br />

Engineering <strong>and</strong> Computer<br />

Science, Massachusetts<br />

Institute of <strong>Technology</strong><br />

Michael S. Triantafyllou<br />

Department of Mechanical<br />

Engineering, Massachusetts<br />

Institute of <strong>Technology</strong><br />

Introduction<br />

The lateral line organ is a unique<br />

sensory mechanism in fish, enabling<br />

complex behaviors based on the interpretation<br />

of local fluid mechanics. For<br />

example, the blind Mexican cavefish<br />

(Astyanax fasciatus) is able to navigate<br />

new environments at high speed<br />

without collision <strong>and</strong> to identify <strong>and</strong><br />

remember features of the environment<br />

(Montgomery et al., 2001; von<br />

Campenhausen et al., 1981). This surprising<br />

feat is accomplished relying<br />

primarily on its lateral line organ for<br />

ABSTRACT<br />

The lateral line is a critical component of fish sensory systems, found to affect<br />

numerous aspects of behavior, including maneuvering in complex fluid environments<br />

with poor visibility. This sensory organ has no analog in modern ocean vehicles,<br />

despite its utility <strong>and</strong> ubiquity in nature, <strong>and</strong> could fill the gap left by sonar<br />

<strong>and</strong> vision systems in turbid, cluttered environments.<br />

To emulate the lateral line <strong>and</strong> characterize its object-tracking <strong>and</strong> shape recognition<br />

capabilities, a linear array of pressure sensors is used along with analytic<br />

models of the fluid in order to determine position, shape, <strong>and</strong> size of various objects<br />

in both passive <strong>and</strong> active sensing schemes. We find that based on pressure information,<br />

tracking a moving cylinder can be effectively achieved via a particle filter.<br />

Using principal component analysis, we are also able to reliably distinguish between<br />

cylinders of different cross section <strong>and</strong> identify the critical flow signature information<br />

that leads to the shape identification. In a second application, we employ pressure<br />

measurements on an artificial fish <strong>and</strong> an unscented Kalman filter to successfully<br />

identify the shape of an arbitrary static cylinder.<br />

Based on the experiments, we conclude that a linear pressure sensor array for<br />

identifying small objects should have a sensor-to-sensor spacing of less than 0.03<br />

(relative to the length of the sensing body) <strong>and</strong> resolve pressure differences of at<br />

least 10 Pa. These criteria are used in the development of an artificial lateral line<br />

adaptable to the curved hull of an underwater vehicle, employing conductive polymer<br />

technologies to form a flexible array of small pressure sensors.<br />

Keywords: underwater sensing, artificial lateral line, pressure sensor arrays<br />

sensory feedback. All fish have this<br />

organ, although not all use it to the<br />

extent of the blind cavefish (see<br />

Montgomery et al., 2001). Many fundamental<br />

behaviors in fish have been<br />

identified by biologists to be lateralline-mediated,<br />

including tracking<br />

prey by their wake (Pohlmann et al.,<br />

2004) <strong>and</strong> recognizing nearby physical<br />

objects (von Campenhausen et al.,<br />

1981). Although the lateral line is a biological<br />

organ, its functionality would<br />

translate well to needs in underwater<br />

vehicle design, such as with object detection,<br />

navigation, <strong>and</strong> flow sensing.<br />

The lateral line organ consists of<br />

two subsystems responding separately<br />

to velocity <strong>and</strong> to pressure gradients<br />

on the surface of the fish, both using<br />

the same underlying sensory element,<br />

the neuromast, which responds directly<br />

to flow velocity. In the case of the<br />

pressure gradient measurements, these<br />

velocity sensors are embedded underneath<br />

the skin in canals periodically<br />

opening via pores to the external flow<br />

(van Netten, 2006). Biological studies<br />

have demonstrated the ability of the<br />

fish to use their lateral line to interrogate<br />

their environment through both<br />

active <strong>and</strong> passive sensing. In active<br />

sensing, a fish uses the flow generated<br />

by repeatedly gliding near new objects<br />

at a short distance (von Campenhausen<br />

130 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


et al., 1981) to interrogate their shape.<br />

In particular, blind cave fish were found<br />

to detect <strong>and</strong> discriminate between<br />

stationary objects or openings of different<br />

geometries in still water (von<br />

Campenhausen et al., 1981; Weissert<br />

& von Campenhausen, 1981; Burt de<br />

Perera, 2004). During passive sensing, a<br />

moving object generates a flow field that<br />

is detected by the stationary lateral line of<br />

astillfish. Vogel <strong>and</strong> Bleckmann (2000)<br />

demonstrated that goldfish use their<br />

lateral line to passively detect <strong>and</strong> discriminate<br />

the size, velocity <strong>and</strong> shape of<br />

passing rods in still water. One key element<br />

emerging from these studies is that<br />

in both active <strong>and</strong> passive settings, the<br />

object identification behavior is tied to<br />

the pressure gradient measurements of<br />

the lateral line <strong>and</strong> appears independent<br />

of the velocity measurements.<br />

While the lateral line organ’s role in<br />

many fish behaviors is becoming progressively<br />

better understood, there<br />

still remain many questions about the<br />

level of information detail available via<br />

the lateral line. The recent advances in<br />

underst<strong>and</strong>ing the central processing<br />

of the lateral line have all been with respect<br />

to the oscillating dipole stimulus<br />

(Curcic-Blake & van Netten, 2006,<br />

Goulet et al., 2007), which is inapplicable<br />

to both passive <strong>and</strong> active object<br />

sensing. In the passive case, a dipole<br />

model neglects the wake that forms<br />

about a moving object. In the active<br />

sensing situation, the interaction betweenthetwobodiesinstillwater<br />

leads to very different pressure distributions.<br />

Due to the difficulty in studying<br />

the neurological aspects of how a<br />

fish utilizes the stimulus of the lateral<br />

line, an artificial representation of the<br />

lateral line can give insight on the use<br />

of pressure sensing for biologically inspired,<br />

engineered applications. This<br />

paper investigates the ability of pressure<br />

sensor arrays, emulating the lateral<br />

line organ, to distinguish the shapes of<br />

physical objects through both active<br />

<strong>and</strong> passive sensing, while also identifying<br />

physical requirements for a constructed<br />

artificial lateral line to be used<br />

for underwater vehicle navigation applications.<br />

The results demonstrate<br />

the promise of a lateral-line-like sensor<br />

for autonomous underwater vehicles<br />

(AUVs) in severe environments.<br />

Previous approaches with artificial<br />

lateral lines have taken a biomimetic<br />

approach of reproducing the lateral<br />

line at the structural level, by recreating<br />

a sophisticated neuromast-like<br />

cantilevered sensors <strong>and</strong> using a canal<br />

system similar to that of the fish in<br />

order to obtain pressure gradient<br />

measurements (Chen et al., 2006;<br />

Yang et al., 2008). Yet, at its core,<br />

the portion of the biological lateral<br />

line associated with the behaviors of<br />

interest can be viewed as a more elementary<br />

processing unit, taking distributed<br />

pressure measurements as<br />

inputs <strong>and</strong>, via some processing that<br />

remains to be understood, extracting<br />

information about the environment.<br />

Thus, the current work instead takes<br />

a bioinspired approach in which the<br />

lateral line is abstracted by a linear<br />

pressure sensor array <strong>and</strong> the resulting<br />

information is processed relying on<br />

modern inference algorithms. In addition<br />

to providing potential clues as to<br />

the fundamental processes taking<br />

place in the biological sensory system,<br />

the present abstracted approach, by directly<br />

focusing on the desired functionality<br />

of the artificial lateral line,<br />

bears significant engineering benefits.<br />

As a sparse artificial lateral line can be<br />

readily implemented, making use of<br />

diaphragm-based pressure sensors, the<br />

initial focus is able to shift from sensor<br />

fabrication to the inference <strong>and</strong> processing.<br />

In turn, insights from the inference<br />

results are then used to develop<br />

specifications for optimal sensor design.<br />

As another benefit of using pressure<br />

sensors to approximate the lateral<br />

line, note that the available pressure<br />

information includes, but is not limited<br />

to, discrete pressure gradients (as in the<br />

lateral line canal system).<br />

First, we present two experiments<br />

demonstrating passive object detection<br />

through the use of an artificial lateral<br />

line. In this study, an artificial lateral<br />

line is used to track the size <strong>and</strong> location<br />

of a moving cylinder <strong>and</strong> separately to<br />

identify the shape of a moving cylinder<br />

between known possibilities. Second,<br />

we present an experiment demonstrating<br />

active object detection, where a<br />

moving artificial lateral line is used to<br />

identify the completely unknown<br />

shape of an object. In this study, the<br />

sensor arrangement is studied with<br />

relation to the applied problem in<br />

order to identify constraints of sensor<br />

spacing. Finally, we discuss the construction<br />

<strong>and</strong> testing of a prototype artificial<br />

lateral line using conductive<br />

polymer materials for use in object detection<br />

<strong>and</strong> navigation applications.<br />

Passive Cylinder<br />

Identification<br />

In passive object identification, an<br />

interaction between the object <strong>and</strong> the<br />

environment generates a pressure field<br />

that stimulates the sensor array. In the<br />

examples considered here, this interaction<br />

is between a moving cylinder<br />

<strong>and</strong> still water. The sensor array is also<br />

stationary in the water. The problem of<br />

identifying a moving cylinder via the<br />

pressure along a lateral-line-like sensor<br />

is addressed in two stages. First, we discuss<br />

an approach for determining the<br />

position <strong>and</strong> size of a cylinder, focusing<br />

on a single shape, <strong>and</strong> subsequently, we<br />

consider the question of distinguishing<br />

between different shapes.<br />

July/August 2011 Volume 45 Number 4 131


Experimental Setup<br />

Two experiments were used to analyze<br />

passive object detection techniques.<br />

In the first experimental setup<br />

(see Figure 1A), the position tracking<br />

ofacircularcylinderwasexperimentally<br />

tested using off-the-shelf pressure<br />

sensors (Honeywell 19C015PG4K)<br />

arranged in a linear array. Seven sensors<br />

were embedded along a 46-cm<br />

square flat plate spaced 1.9 cm<br />

(0.75 inch) apart, as shown in Figure<br />

1A. Using a mechanical stage<br />

with millimeter accuracy in the<br />

x-y plane, a 32-mm diameter circular<br />

cylinder was passed in front of the sensor<br />

array at constant velocity <strong>and</strong><br />

known position. These experiments<br />

were performed in a 3.6 m × 1.2 m ×<br />

1.2 m water tank at the Singapore-<br />

MIT Alliance for Research <strong>and</strong> <strong>Technology</strong><br />

(SMART) Centre. Pressure<br />

signals were amplified by a factor of<br />

1000 immediately adjacent to the sensors<br />

using AD620 instrumentation<br />

amplifiers <strong>and</strong> were next sampled<br />

using an NI USB-6210 analog-todigital<br />

converter. The sensors were<br />

calibrated using static pressure.<br />

In the second passive detection experimental<br />

setup, as with the first, a<br />

verticallymountedcylinderpassesin<br />

front of a stationary linear pressure<br />

sensor array at constant velocity (Figure<br />

1B). In this experiment, four<br />

Honeywell 242PC15M pressure sensors<br />

were enclosed in a streamlined<br />

cylindrical enclosure. The sensors<br />

were mounted rigidly with a small diameter<br />

outlet in order to minimize<br />

noise. Due to onboard amplification,<br />

the sensors were only amplified by a<br />

factor of 10 before data acquisition<br />

(NI USB 6210). These experiments<br />

were performed in the MIT towing<br />

tank facility (36.6 m × 2.4 m ×<br />

1.2 m). As the goal of these experiments<br />

<strong>and</strong> subsequent analysis is to<br />

FIGURE 1<br />

Schematics of two experimental apparatuses for passive object identification <strong>and</strong> sample pressure<br />

data from one apparatus. (A) A moving circular cylinder that is towed past an array of seven stationary<br />

pressure sensors on a flat wall. (B) An array of four pressure sensors mounted on a stationary streamlined<br />

body with a square cylinder towed past the body. (C) A representative data set corresponding to<br />

the layout in part B, filtered with a cut-off frequency of 100 Hz. (Color versions of figures available<br />

online at: http://www.ingentaconnect.com/content/mts/mtsj/2011/00000045/00000004.)<br />

distinguish between two cylinder<br />

cross-section shapes known apriori,<br />

both square <strong>and</strong> round cross-section<br />

cylinders were towed past the sensor<br />

array. In order to provide a reliable<br />

common denominator for shape classification,<br />

the data was gathered from<br />

a variety of speeds, cylinder sizes, <strong>and</strong><br />

distances from the sensor array as<br />

shown in Table 1. Calibration of the<br />

pressure sensors was completed in situ<br />

by comparing the amplitude of pressure<br />

oscillations due to regularly generated<br />

waves, using a wavemaker, to the<br />

theoretical amplitude based on linearized<br />

surface wave theory.<br />

Passive Cylinder Tracking<br />

Tracking the position <strong>and</strong> size of a<br />

circular cylinder can be interpreted as<br />

the first step or lowest level of shape<br />

identification. This task is complicated<br />

by the wake that forms when a cylinder<br />

moves. In order to track the position<br />

132 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


<strong>and</strong> size, an accurate <strong>and</strong> preferably<br />

simple model for relating the cylinder<br />

state to the pressure measurements is<br />

needed as well as a technique for estimating<br />

the state in real time. Potential<br />

flow solutions are an immediate c<strong>and</strong>idate<br />

for modeling an object in a fluid<br />

since they provide an analytic solution<br />

with relatively few parameters to define<br />

the complete flow field.<br />

The base model chosen for tracking<br />

a circular cylinder in the present experimentsisaRankinehalf-bodyin<br />

potential flow. This well-known structure<br />

is described by the superposition<br />

of a source <strong>and</strong> a free stream. A mirror<br />

image of the half-body provides the<br />

necessary boundary conditions to<br />

define the wall in the model. This<br />

model was found to approximate the<br />

pressure field of a cylinder with a<br />

wake when compared with pressure<br />

measurements obtained through a viscous<br />

numerical simulation. The radius<br />

of the cylinder was taken as the distance<br />

from the location of the source<br />

to the nearby forward stagnation<br />

point in the flow. It is important to<br />

note that the dipole model frequently<br />

used in a vibratory setting in studies<br />

with fish (e.g., Coombs & Janssen,<br />

1990) does not model the flow well<br />

about a steadily translating body due<br />

to the absence of the wake in the<br />

model. It is not surprising that a potential<br />

flow model like the Rankine halfbody<br />

approximates the experimental<br />

situation, since the flow outside the<br />

separated region in the wake can be<br />

reasonably considered as potential<br />

flow. Although the real situation is<br />

not steady <strong>and</strong> the pressure far from<br />

the cylinder was found to converge<br />

more slowly than in numerical simulations,<br />

a Rankine model captures the<br />

amplitude <strong>and</strong> shape of the pressure response<br />

well in the immediate vicinity<br />

of the cylinder. Using the velocity potential<br />

from this model, the pressure at<br />

the sensor locations is calculated using<br />

Bernoulli’s equation.<br />

Although the pressure sensors used<br />

in these experiments measure pressure<br />

with respect to a fix reference value,<br />

one key element in the analysis for<br />

tracking a cylinder is to consider the<br />

difference in pressure between adjacent<br />

pressure sensors instead of the absolute<br />

pressure. Long-wavelength,<br />

small-amplitude disturbances at the<br />

free surface of the tank, caused by the<br />

motion of test cylinder, were found to<br />

cause pressure fluctuations on the<br />

order of 20 Pa, significant compared<br />

to the 100 Pa pressure signals from<br />

the cylinder. Due to the long wavelength,<br />

taking the difference in pressure<br />

signals effectively removes this contamination.<br />

As mentioned in the Introduction,<br />

the portion of the lateral line that<br />

has been associated with object identification<br />

responds to pressure differences<br />

(Coombs, 2001), so it is interesting<br />

that the use of pressure differences is<br />

necessary even when absolute pressure<br />

measurements are available.<br />

For the purpose of tracking the<br />

position <strong>and</strong> size of a cylinder, the forward<br />

model relating the hidden parameters<br />

of interest to the available<br />

measurements has been defined by<br />

the Rankine half-body representation<br />

of the cylinder <strong>and</strong> wake, in conjunction<br />

with the Bernoulli equation to<br />

obtain the pressure at the sensors. To<br />

tackle the inverse problem <strong>and</strong> estimate<br />

the cylinder state based on the<br />

pressure measurements, several additional<br />

issues require consideration.<br />

First, the variables of interest are the<br />

position (x <strong>and</strong> y as labeled in Figure<br />

1A) <strong>and</strong> radius (R) of the cylinder.<br />

In order to cast the problem in the<br />

form of a hidden Markov model (see,<br />

for example, Cappe et al., 2005), in<br />

which the state at each timestep depends<br />

solely on the previous iteration,<br />

the velocity (u) must to be included in<br />

the state vector. Using this hidden<br />

Markov model formulation allows for<br />

the use of well-known estimation algorithms<br />

for solving the inverse problem,<br />

such as the Kalman filter extensions for<br />

nonlinear models (Gelb, 1974; Julier<br />

& Uhlmann, 2004). Second, the forward<br />

model requires a strict constraint<br />

on two of the state variables, R <strong>and</strong> y, in<br />

order to correspond to a physical realization:<br />

0 < R < y. Unfortunately, there<br />

is no natural boundary in the measurement<br />

model that corresponds to these<br />

physical constraints, as the potential<br />

flow model works uniformly well <strong>and</strong><br />

there is no sudden shift in the predicted<br />

pressure. As a consequence of this, an estimation<br />

technique such as the extended<br />

Kalman filter fails when it is applied to<br />

naturally noisy experimental data.<br />

The successful tracking of the position<br />

<strong>and</strong> size of a cylinder was accomplished<br />

instead using a particle filter<br />

(see Branko et al., 2004). This general<br />

inverse problem technique operates by<br />

using a number of samples, which represent<br />

possible states. To each sample,<br />

there is a corresponding weight that<br />

roughly tracks the quality of that sample.<br />

At each time step, each sample is<br />

updated to a new value r<strong>and</strong>omly chosen<br />

from a distribution based on the<br />

previous sample, <strong>and</strong> the weight of<br />

that sample is updated based on how<br />

likely it is to have generated the measurements.<br />

The key advantage of this<br />

approach is that, in specifying the<br />

noise distributions that govern the update<br />

of the samples, it is possible to<br />

limit the range of state vectors to<br />

those that correspond to a physical system.<br />

In the present case, the x position<br />

<strong>and</strong> the velocity u are assumed to have<br />

Gaussian noise distributions, which is<br />

the typical assumption for an unconstrained<br />

variable where the noise is<br />

July/August 2011 Volume 45 Number 4 133


generated by the accumulation of<br />

many small r<strong>and</strong>om influences. The<br />

noise distribution associated with the<br />

y position is assumed to be a lognormal<br />

distribution, <strong>and</strong> for the radius<br />

it is assumed to be a Gaussian distribution<br />

that is truncated at 0 <strong>and</strong> y <strong>and</strong> renormalized.<br />

These distributions satisfy<br />

the constraints between the variables<br />

<strong>and</strong> have intuitive limiting behaviors.<br />

When the mean is far from zero with<br />

respect to the st<strong>and</strong>ard deviation, the<br />

lognormal is approximately symmetric<br />

about the mean. In the case of the<br />

truncated Gaussian distribution, if<br />

the previous value of the radius is<br />

greater than the new y value, then the<br />

distribution is weighted heavily to<br />

larger values <strong>and</strong> has a maximum<br />

likelihood of y. Themaincostto<br />

using a particle filter is that it frequently<br />

requires a large number of<br />

samples to generate repeatable accurate<br />

results. For this application, it was<br />

found that approximately 150 particles<br />

are sufficient.<br />

A typical result of the particle filter<br />

implementation on experimental data<br />

is shown in Figure 2. In the figure,<br />

the black line corresponds to the estimate<br />

of the variable being output<br />

from the filter, <strong>and</strong> the red line corresponds<br />

to the true value. Note that<br />

the filter is initialized with a uniform<br />

distribution <strong>and</strong> does not begin to<br />

converge until there are significant<br />

pressure differences. As seen from the<br />

results, the particle filter estimate generally<br />

tracks the x position <strong>and</strong> radius<br />

well but underestimates the y position.<br />

The velocity is also underestimated in<br />

this example, although both overestimates<br />

<strong>and</strong> underestimates of the velocity<br />

were recorded in general. These<br />

results demonstrate that using a steady<br />

potential flow model, the radius <strong>and</strong><br />

aspects of the position can be accurately<br />

tracked. With further refinements of<br />

the model, the underestimation of<br />

the y position could likely be overcome.<br />

The use of a simple analytical<br />

model here makes it more likely that<br />

equivalent information is also available<br />

to a fish via its lateral line, <strong>and</strong> similarly<br />

likely that it could be adapted for<br />

quickly tracking objects near a hull.<br />

TABLE 1<br />

Matrix of experiments for passive cylinder shape classification.<br />

Passive Cylinder<br />

Shape Classification<br />

The second set of experiments is<br />

concerned with the more complicated<br />

problemofshapeclassification. Fundamentally,<br />

we consider whether it is<br />

possible to distinguish between two<br />

similar shapes after a single pass of an<br />

object past the sensor array, using a<br />

robust test for the decision. The test<br />

in question must identify whether<br />

a passing cylinder has a square or<br />

round cross section. Since the decision<br />

is based on a single pass, the stochastic<br />

nature of the flow in the wake of the<br />

cylinder must be overcome.<br />

The large data set of pressure obtained<br />

from the experimental setup in<br />

Figure 1B, corresponding to cylinders<br />

of different sizes, velocities, distances,<br />

<strong>and</strong> shapes provides a broad range of<br />

parameters for verifying the ability of<br />

any decision rule. In addition, the<br />

large number of runs for each point<br />

in the test matrix allows for an accurate<br />

comparison of the mean pressure responses.<br />

As shown in Figure 3, there<br />

are distinct differences between the<br />

average pressure measured due to the<br />

circular <strong>and</strong> square cross sections; analogous<br />

differences exist for comparisons<br />

in all the tested sizes <strong>and</strong> velocities. It is<br />

notable that the most easily distinguished<br />

differences in the pressure signals<br />

occur after the zero crossing point<br />

(circled).<br />

The analysis presented here develops<br />

a test to determine, after the fact,<br />

whether a square or circular cylinder<br />

has passed the sensors. Based on the<br />

observations of the mean pressure<br />

traces, it is tempting to choose features<br />

such as the location of the minimum<br />

pressure to classify the shape of the<br />

cylinder. This approach is limited in<br />

that the features would be localized<br />

in the pressure trace <strong>and</strong>, therefore,<br />

highly susceptible to the type of noise<br />

observed in Figure 1C. Instead, we use<br />

a principal component analysis (PCA)<br />

to identify a limited number of features<br />

in the form of weighted sums over the<br />

full pressure data from a single run.<br />

This type of feature depends on all the<br />

data <strong>and</strong> is less susceptible to localized<br />

perturbations such as those in the<br />

wake. By carefully applying the PCA<br />

to a training data subset, the results<br />

can be directed towards a decision rule<br />

in classifying the shape of the cylinders.<br />

In implementation, PCA works as a<br />

singular value decomposition of the<br />

7.62 cm Diameter 5.08 cm Diameter<br />

0.5 m/s 0.75 m/s 0.5 m/s 0.75 m/s<br />

Square 100 runs 100 runs 100 runs 100 runs<br />

Round 100 runs 100 runs 100 runs 100 runs<br />

Each of the experiments listed corresponds to the cylinder passing at a distance of 5.1 mm from the<br />

sensors. Additional data, not listed in the table, were collected with the 5.08-cm-diameter cylinders passing<br />

1.27 cm from the sensors at 0.75 m/s.<br />

134 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Results of a cylinder tracking experiment using a particle filter. In the first four parts, each vertical<br />

slice in the surface corresponds to the probability density function for that variable at the time of<br />

the slice, given the previous pressure measurements. In these sections, the red line represents the<br />

true value of the parameter at each time instant, <strong>and</strong> the black line corresponds to the expected<br />

value of the distribution. The final part shows the corresponding pressure difference measurements,<br />

filtered at 60 Hz for anti-aliasing.<br />

sample covariance matrix of the data<br />

(Jolliffe, 2002; Jackson, 2003). One important<br />

detail in the implementation is<br />

that the mean of each initial feature (a<br />

sample point in the pressure trace)<br />

must be removed. This mean is taken<br />

across all training data, not over each<br />

class. With the mean removed, the<br />

sample covariance is straightforward<br />

to calculate. Each resulting principal<br />

component is a vector that accounts<br />

for the maximum possible variance,<br />

subject to having unit area <strong>and</strong> being<br />

uncorrelated with all the previous principal<br />

components.<br />

Since only four sensors were available<br />

for this experiment (as opposed<br />

to seven in the previous cylinder tracking<br />

experiments) <strong>and</strong> the cylinder<br />

stimuli are translating at constant velocity,<br />

an equivalence between the<br />

pressure time history <strong>and</strong> the spatial<br />

pressure field was used in the analysis.<br />

This approximation is very accurate in<br />

regions in front of the cylinder, but less<br />

so in the unsteady wake areas. In order<br />

to compare data sets with different cylinder<br />

velocities on the same spatial<br />

scale, the 0.5 m/s velocity data is<br />

down-sampled appropriately. In addition,<br />

the data is aligned by the point<br />

at which the pressure crosses zero<br />

(marked in Figure 3), which forms a<br />

reliable internal placement marker.<br />

Finally, the PCA approach does not<br />

naturally produce features which capture<br />

the cylinder shape. The decomposition<br />

of the covariance matrix<br />

generates principal components that<br />

capture the majority of the data variance<br />

in the first few components.<br />

Therefore, the resulting principal components<br />

will be most useful if the primary<br />

cause of differences in the data is<br />

due to the shape of the cylinders.<br />

While the variation in the data due to<br />

the flow separation <strong>and</strong> the r<strong>and</strong>om<br />

phase of the wake is impossible to<br />

July/August 2011 Volume 45 Number 4 135


FIGURE 3<br />

A comparison of the mean pressure measured at one experimental configuration reveals clear<br />

differences between the two cylinder shapes. The st<strong>and</strong>ard deviation is also plotted, offset by<br />

−2.0 kPa.<br />

remove, the velocity of the cylinder<br />

strongly correlates with the pressure<br />

signal amplitude <strong>and</strong> masks the response<br />

to shape. Normalizing each<br />

pressure trace by the maximum pressure<br />

effectively removes this masking<br />

effect since the pressure amplitude is<br />

consistently proportional to the velocity<br />

squared.<br />

Using the first three principal components<br />

based on a training set of data<br />

from a single sensor <strong>and</strong> covering all of<br />

the experimental parameters, a highly<br />

successful classification test is obtained.<br />

These three principal components<br />

form a space in which the data<br />

points generated from experimental<br />

runs are grouped into two roughly<br />

ellipsoidal clouds, which can be optimally<br />

separated by a decision plane,<br />

minimizing the sum of squared error<br />

on the training data. Since the projections<br />

of the data points on each axis are<br />

based on a linear combination of the<br />

pressure data with the corresponding<br />

FIGURE 4<br />

principal component, a single vector<br />

of weights can be found that corresponds<br />

to an axis perpendicular to<br />

the decision plane (Figure 4B). This<br />

results in a single linear combination<br />

using these weights with the<br />

normalized pressure measurements<br />

to determine the score of an experimental<br />

run, with a positive value<br />

implying a square cross section. When<br />

applied to the test data (excluding the<br />

training data), this test produced<br />

a very small misclassification rate<br />

of 1.2%.<br />

While the decision test encapsulated<br />

in the first two parts of Figure 4<br />

results in a surprisingly high accuracy,<br />

the difficulty with using principal<br />

components is that there is generally<br />

little corresponding intuition for why<br />

it works so well. The key element of<br />

the decision test is given by the decision<br />

weights, illustrated in Figure 4B.<br />

There are three distinct main sections<br />

The elements of the PCA-based classification test (A <strong>and</strong> B), <strong>and</strong> the accuracy of the test using<br />

different subsections of the full decision coefficients (C).<br />

136 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


to the decision weights based on the<br />

PCA results, labeled II, III, <strong>and</strong> IV in<br />

Figure 4B. Region II is the smallest in<br />

magnitude <strong>and</strong> corresponds to the region<br />

between the maximum pressure<br />

<strong>and</strong> the zero crossing. Region III corresponds<br />

to the area of largest difference<br />

between the two shapes shown in Figure<br />

3. Region IV weighs an area of the<br />

pressure response that is directly influenced<br />

by the wake of the cylinder. This<br />

region has considerable variation in the<br />

data, but it is unclear whether it is related<br />

to the cylinder shape. Regions I<br />

<strong>and</strong> IV on either side are very lightly<br />

weighted. By zeroing out regions, it is<br />

possible to examine the importance of<br />

the remaining portion of the decision<br />

weights to the accuracy in classifying<br />

the shape. In this zeroing process (Figure<br />

4C), we find that data on the rear<br />

side of the zero-crossing point is the<br />

most important in classifying cylinder<br />

shape. In fact, the data before the zerocrossing<br />

point appears to add almost<br />

no new information for classifying<br />

the shape.<br />

The emphasis on region IV implies<br />

that the wake of the cylinder contains a<br />

substantial portion of the information<br />

being used to classify the shape. Initially,<br />

this may appear to be through<br />

vortex spacing, governed by the Strouhal<br />

number. Although this could help<br />

in classifying the shape if particular<br />

sizes were being compared, in these experiments<br />

the large round cylinder <strong>and</strong><br />

the small square cylinder had nearly<br />

equal vortex spacing, meaning that<br />

half of the data would be difficult to<br />

separate by this trait. This observation is<br />

inconsistent with the 1% error rate.<br />

The results based on principal components<br />

demonstrate that it is possible<br />

to distinguish between two relatively<br />

similar cross sections of moving cylinders<br />

based on an artificial lateral line<br />

sensor. This has been hinted at based<br />

on experiments with fish, but the<br />

clear identification of shape without<br />

regard to changes in velocity or size<br />

has not been demonstrated. Given the<br />

importance of the wake in identifying<br />

the shape in the experiment, it is possible<br />

that the ability of the lateral line,<br />

<strong>and</strong> any artificial analogs, to identify<br />

shapes does not extend to arbitrary<br />

shapes. In particular, the bluntness<br />

of the leading face of the cylinder<br />

has a strong impact on the location<br />

<strong>and</strong> flow direction at the point of<br />

flow separation. If two shapes with<br />

similar sharp flow separation are employed,<br />

it may be much harder to distinguish<br />

them.<br />

Active Cylinder<br />

Identification<br />

In active object identification, we<br />

use pressure measurements from an artificial<br />

lateral line sensor array to locate<br />

<strong>and</strong> identify stationary objects in still<br />

water without prior information on<br />

their shape. The problem is considered<br />

from a two-dimensional perspective<br />

in which a moving fish-like<br />

body glides at constant speed past a<br />

column-like stationary object of<br />

unknown location, size, <strong>and</strong> crosssection<br />

geometry. The moving body<br />

is equipped with a finite number of<br />

pressure sensors distributed over its<br />

surface. The viscous effects are<br />

ignored in the model used to calculate<br />

the pressure <strong>and</strong> the limits of this assumption<br />

are discussed.<br />

Hassan (1985, 1992) showed that<br />

in the case of a fish gliding towards a<br />

cylinder or a wall, the pressure increases<br />

noticeably in the front region<br />

when the fish gets close to the obstacle.<br />

He also showed (Hassan, 1985) that<br />

for a fish gliding past a cylinder, there<br />

is a univocal relationship between<br />

spacio-temporal hydrodynamic signature<br />

<strong>and</strong> size <strong>and</strong> distance to the cylinder,<br />

an indication that an artificial<br />

lateral line may be able to distinguish<br />

these features. More recent studies<br />

(Sichert et al., 2009; Bouffanais et al.,<br />

2011) have considered parameterizations<br />

of the shape of an object relevant<br />

to potential flow models, giving an<br />

analytic representation of the flow<br />

field that can be easily related to pressure<br />

<strong>and</strong> velocity measurements.<br />

Experimental Setup<br />

Experiments were conducted in the<br />

SMART Centre water tank. The complete<br />

setup is shown in Figure 5A. The<br />

FIGURE 5<br />

Active sensing experimental layout. (A) A picture<br />

of the experimental set-up used for active object<br />

identification (the foil is moving towards<br />

the photographer). (B) A schematic of the<br />

cross-section of the same experiment showing<br />

the location of the pressure ports.<br />

July/August 2011 Volume 45 Number 4 137


moving body used to both generate the flow <strong>and</strong> sense pressure was a NACA<br />

0018 foil (chord c = 15 cm <strong>and</strong> span s = 60 cm) cast with internal 0.318 cm PVC<br />

tubing to transmit pressure from taps at the foil’s midspan to the top. Honeywell<br />

19C015PG4K pressure sensors were mounted on top of the foil, <strong>and</strong> measurements<br />

were collected at a sampling rate of 500 Hz via a NI USB-6289 DAQ.<br />

The location of the sensor ports is shown in Figure 5B. The foil was dragged at<br />

velocity v = 0.5 m/s past a static cylinder of elliptical cross section oriented at various<br />

angles. At its closest point, the foil was 5-10 mm away from the cylinder.<br />

The Forward <strong>and</strong> Inverse Problem<br />

It is believed that blind cave fish can encode separately the distance, size <strong>and</strong><br />

shape of objects (Burt de Perera, 2004). Therefore, a convenient <strong>and</strong> potentially<br />

biologically relevant way to parameterize the problem is to characterize an object<br />

by two parameters accounting for its position, one size parameter <strong>and</strong> several shape<br />

parameters. Another desirable feature of this parameterization is that the number<br />

of shape parameters needed to account for the pressure decreases with the<br />

distance to the object. Bouffanais et al. (2011) proposed such a characterization:<br />

<br />

<br />

SðθÞ ¼ a þ R e iθ ∞<br />

þ ∑ μk e ikθ<br />

k¼1<br />

¼ x þ iy þ R e iθ þ ∑<br />

∞ jμk je ik ð θ<br />

k¼1<br />

α kÞ<br />

<br />

; θ ∈ ½0; 2πŠ<br />

wherein a (a = x + iy) refers to the location of the object, R to its size <strong>and</strong> each μ k<br />

(μ k =|μ k |e ikα k<br />

)termisassociatedwitha(k +1)-gonaltypeofperturbationofthe<br />

shape from that of a circle. As k increases, the impact of the μ k term on the pressure<br />

field decays very quickly with the distance from the cylinder. For an ellipse,<br />

as in these experiments, only the first shape parameter μ 1 is non-zero <strong>and</strong> therefore<br />

the subscripts will be dropped.<br />

Given the moving object, its trajectory, <strong>and</strong> the location, size <strong>and</strong> shape (a,<br />

R, μ) of the stationary object, the velocity potential anywhere in the flow field<br />

can be expressed in terms of a singularity distribution over the surface of the objects.<br />

The problem is solved numerically using a source panel method: the surface<br />

of each object is broken up into line segments of constant source strength (364 line<br />

segments were used for the moving body <strong>and</strong> 150 for the stationary object). The<br />

pressure at the sensor locations on the surface of the moving body can then be<br />

expressed in terms of the potential at these points using the unsteady Bernoulli<br />

equation. The combination of the parameterization of the object shape <strong>and</strong><br />

the numerical potential flow model defines the forward methodology, relating<br />

the pressure measurements to the variables of interest (location, size, <strong>and</strong><br />

shape). In experiments, the pressure measured by the sensors is corrupted<br />

by noise. The two main sources of noise are laboratory noise (electrical <strong>and</strong><br />

mechanical) <strong>and</strong> the background noise of the fluid flow (which includes viscous<br />

effects).<br />

For proper inversion, the technique must be (1) robust to noise, (2) capable of<br />

h<strong>and</strong>ling non-linearity since the pressure does not depend linearly on the characterizing<br />

parameters of the stationary object, <strong>and</strong> (3) dynamic, in order to be used<br />

for navigation of underwater vehicles.<br />

A particularly suitable algorithm for<br />

such inversion is the unscented Kalman<br />

filter (UKF) ( Julier & Uhlmann,<br />

2004), which is a robust dynamic<br />

probabilistic signal filtering technique<br />

for highly nonlinear systems. The<br />

UKF is more accurate than the more<br />

traditional extended Kalman filter for<br />

highly nonlinear problems <strong>and</strong> does<br />

not require the computation of derivatives<br />

for which no analytic expressions<br />

are available. It also propagates the statistics<br />

with fewer samples than the<br />

more powerful particle filter, which<br />

makes it better suited for real time<br />

applications.<br />

Subtle modifications to the UKF<br />

are necessary before applying it to our<br />

problem. Non-physical configurations,<br />

such as bodies intersecting each<br />

other cannot be identified by an UKF<br />

assuming Gaussian distributions. This<br />

can skew statistics, producing unusable<br />

results. To avoid this problem, if after<br />

updating the location of the moving<br />

object, the estimated mean configuration<br />

is not valid, the estimated position<br />

is shifted ‘out of the way.’ If the mean<br />

configuration is valid, the parameter α<br />

that determines the spread of the samples<br />

around the mean in the unscented<br />

transform (Wan & van der Merwe,<br />

2001) is chosen small enough (for<br />

each time step) that all the samples<br />

are valid.<br />

TheUKF<strong>and</strong>theforwardmodel<br />

are combined to solve the inverse problem:<br />

locating <strong>and</strong> identifying a cylinder<br />

using pressure measurements. In the<br />

results <strong>and</strong> analysis discussed here,<br />

the steady pressure due to the constant<br />

velocity of the foil has been subtracted<br />

from the pressure signal. The measurement<br />

covariance matrix used in the<br />

UKF was calculated for each run<br />

based on the pressure measured 0.5-<br />

0.3 s before the characteristic drop<br />

138 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


of pressure at the second sensor (see<br />

Figure 7C).<br />

FIGURE 6<br />

Results of a cylinder detection <strong>and</strong> identification in two passes (A <strong>and</strong> B) using an UKF. The colored<br />

dashed lines show the true value of the parameters. (a–f) The corresponding shape estimate (red<br />

circle or ellipse), the actual cylinder (green dotted ellipse), <strong>and</strong> the position of the foil at various<br />

times. (Color versions of figures available online at: http://www.ingentaconnect.com/content/mts/<br />

mtsj/2011/00000045/00000004.)<br />

Active Sensing Results<br />

<strong>and</strong> Analysis<br />

Due to the amount of noise in the<br />

experiments <strong>and</strong> the fact that it was<br />

highly correlated, all attempts to simultaneously<br />

fully identify the location<br />

<strong>and</strong> geometry of the cylinder<br />

were unsuccessful. However, fish<br />

have been observed to pass several<br />

times in front of new objects (von<br />

Campenhausen et al., 1981), <strong>and</strong> it<br />

seems reasonable to assume that they<br />

first locate the objects <strong>and</strong> estimate<br />

their size before refining their estimate<br />

of the shape. A similar approach is used<br />

here: the first pass is used to get a first<br />

estimate of the position (x <strong>and</strong> y) <strong>and</strong><br />

size (R) of the cylinder. A second pass<br />

using the same data refines the first estimates<br />

<strong>and</strong> estimates the shape parameters<br />

(|μ| <strong>and</strong>α). This approach is<br />

consistent with the dynamic hierarchical<br />

access associated with the shape<br />

characterization parameters used by<br />

Bouffanais et al. (2011). An example<br />

of such a process of cylinder geometry<br />

reconstruction is shown in Figure 6 for<br />

a cylinder with shape parameters R =<br />

3.81 cm <strong>and</strong> μ = 0.2i <strong>and</strong> the foil passing<br />

6.5 mm away from it.<br />

Both behavioral studies (Weissert<br />

& von Campenhausen, 1981) <strong>and</strong><br />

mathematical modeling (Hassan,<br />

1985) have shown that the presence<br />

of the static object in still water cannot<br />

be detected until the fish (or here the<br />

foil) is very close to it (on the order<br />

of the width of the moving body). As<br />

canbeseeninFigure6A,assoonas<br />

the foil is close enough to the cylinder,<br />

the position (x <strong>and</strong> y) <strong>and</strong> size (R) estimates<br />

converge steadily towards the<br />

true value of the parameters. The second<br />

run allows for very good orientation<br />

(α) estimation (Figure 6B). The<br />

aspect ratio of the ellipse (|μ|) is the<br />

more difficult feature to reconstruct.<br />

The estimated parameter |μ| first w<strong>and</strong>ers<br />

around the actual value of the<br />

parameter before decreasing towards<br />

a lower value. As can be seen in Figure<br />

6F even though the shape of the<br />

ellipse has not been exactly reconstructed,<br />

the estimate of the half of<br />

the ellipse that is closest to the foil is<br />

reasonably accurate. This observation<br />

confirms that object identification<br />

based on pressure sensing becomes<br />

less reliable for features further from<br />

the sensors.<br />

The results demonstrate that object<br />

localization <strong>and</strong> object recognition are<br />

possible with experimental pressure<br />

measurements; however, there are limitations<br />

to the proposed method. The<br />

main limitation is the use of potential<br />

flow models to establish the governing<br />

equations for the filter <strong>and</strong> the simulations<br />

(the importance of viscosity on<br />

the stimulus to the lateral line system<br />

of fish is also discussed in Windsor &<br />

McHenry, 2009). A fourth pressure<br />

sensor placed near the tail of the foil<br />

measured an oscillating pressure signal<br />

characteristic of the wake, unlike the<br />

first three sensors. Particle image<br />

velocimetry was used to visualize the<br />

flow <strong>and</strong> compare it to the flow predicted<br />

by the inviscid model. As can<br />

be seen in Figures 7A <strong>and</strong> 7B, the<br />

two flows are nearly identical almost<br />

everywhere, but as the foil passes the<br />

cylinder, separation occurs over a<br />

small region on the foil (in the orange<br />

ellipse). Comparing the pressure measurements<br />

<strong>and</strong> simulations (Figure<br />

7C), we can see that the potential<br />

flow model gives very good approximations<br />

of the pressure measurements<br />

until the second sensor passes the cylinder<br />

(black dotted line). After that<br />

point at which separation occurs, the<br />

inviscid assumption is violated <strong>and</strong><br />

the model is no longer valid. Therefore,<br />

only the measurements before<br />

the black dotted line on the figure<br />

July/August 2011 Volume 45 Number 4 139


FIGURE 7<br />

Comparison between simulated (A) <strong>and</strong> experimental (B) flow field. The green <strong>and</strong> orange ellipses<br />

show where the flows differ. (C) The pressure data (plain line) is filtered with a cut-off frequency of<br />

100 Hz. (Color versions of figures available online at: http://www.ingentaconnect.com/content/<br />

mts/mtsj/2011/00000045/00000004.)<br />

have been used to locate <strong>and</strong> identify<br />

the cylinder.<br />

some of the experiments that still provided<br />

successful tracking results. The<br />

pressure data for the active sensing experiments<br />

was of a similar scale. Any<br />

sensors used for an artificial lateral<br />

line application must provide the sensitivity<br />

<strong>and</strong> correspondingly low noise<br />

floor to distinguish small changes in<br />

pressure of at least 10 Pa.<br />

The experiments in active sensing<br />

ofobjectsprovideaguidetotheoptimal<br />

density of sensors needed for<br />

similar applications, since the curved<br />

surface of the foil <strong>and</strong> the self-generated<br />

flow increase in the importance<br />

of the instantaneous measurements<br />

of the spatial pressure distribution.<br />

Simulations were performed to examine<br />

the affect of sensor density on the<br />

ability to identify a stationary object.<br />

The simulations consisted of the<br />

same foil described earlier gliding at<br />

0.5 m/s <strong>and</strong> passing 7 mm away from<br />

a cylinder with the geometry parameters<br />

R 0 = 1.5 cm <strong>and</strong> μ =0.2i.<br />

White Gaussian noise of st<strong>and</strong>ard<br />

deviation 2 Pa was added to the<br />

simulated pressure measurements. Between<br />

10 <strong>and</strong> 70 pressure measurement<br />

points were evenly distributed<br />

along the front three quarters of the<br />

foil (see Figure 8C), <strong>and</strong> the sampling<br />

Sensor Array Constraints<br />

The use of off-the-shelf sensors severely<br />

limits the density which can be<br />

achieved in the sensor array while<br />

maintaining the sensitivity to deviations<br />

from the static pressure needed<br />

for engineering applications. From<br />

the perspective of sensor sensitivity,<br />

the weakest signals are those associated<br />

with tracking the position <strong>and</strong> size of a<br />

moving cylinder. The maximum absolute<br />

pressure was as small as 30 Pa in<br />

FIGURE 8<br />

Convergence time (A) <strong>and</strong> final error (B) of the object identification as a function of sensor<br />

spacing. (C) The location of the sensors for a spacing of 6 mm.<br />

140 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


ate was chosen such that f = (number of sensors) / 20,000. The error was calculated<br />

as E ¼ 1 ∇R<br />

5 R 0<br />

þ ∇x<br />

R 0<br />

þ ∇y<br />

R 0<br />

þ ∇μ x<br />

þ ∇μ y<br />

. For each simulation, the convergence<br />

time (time elapsed before E < 0.2) <strong>and</strong> the final error were calculated.<br />

Each case was simulated 12 times, <strong>and</strong> the mean <strong>and</strong> st<strong>and</strong>ard deviations of the<br />

time of convergence <strong>and</strong> final error were computed (Figures 8A <strong>and</strong> 8B). The<br />

goal of this procedure was to identify a sensor density below which the lack of<br />

spatial density cannot be compensated by a higher sampling frequency.<br />

Both the convergence time <strong>and</strong> final error plots suggest that, at least for the<br />

configuration considered, the performance of the object identification decreases<br />

as the sensor spacing exceeds 5 mm, which corresponds to a spacing of 0.03 relative<br />

to the length of the foil. The optimal spacing that these observations suggest<br />

scales favorably with the actual spacing of lateral line canal neuromasts in the<br />

trunk canal of the blind Mexican cave fish, which is roughly 0.02 body lengths<br />

(measured from image in Windsor & McHenry, 2009). This optimal spacing<br />

derived from simulations is specific to identifying the shape of cylindrical objects<br />

in which the cross section size is on the same order or slightly smaller than the<br />

body length. In general, the optimal spacing for measuring the pressure distribution<br />

may scale with the size of the object being detected as well as the body<br />

length. In addition, based on the observed separation that occurs on the surface<br />

of the sensing body, the pressure sensors should be distributed only over roughly<br />

the forward half of the body in order to maximize the effectiveness of the sensor<br />

array without waste.<br />

A New Approach for a Flexible Pressure Sensor Array<br />

In order to achieve the optimal spatial distributions <strong>and</strong> sensitivity of pressure<br />

sensors in an effective artificial lateral line, it is necessary to develop sensor<br />

arrays on smaller scales than those possible using off-the-shelf technologies. In<br />

addition to the spatial <strong>and</strong> sensitivity constraints, it would be necessary that the<br />

sensor be applied directly on a surface in order for a lateral line sensor to be<br />

practical for ocean vehicles, instead of fabricating the sensing body around<br />

the sensor as was necessary in the experiments reported here. To address<br />

these problems, we have begun to develop a thin, flexible, one-dimensional<br />

array of pressure sensors to meet the pressure <strong>and</strong> spatial resolution requirements.<br />

The array is designed to conform to an AUV’s hull without protruding<br />

significantly.<br />

While MEMS pressure sensors are typically made with a silicon substrate<br />

(Senturia, 2001), the flexible sensor array described here is made entirely of a<br />

silicone elastomer material, allowing it to be rugged, waterproof, <strong>and</strong> flexible.<br />

A conductive polymer is used as the pressure sensitive element in order to function<br />

while maintaining mechanical compatibility with the rest of the flexible<br />

structure.<br />

The conductive polymer is composed of the silicone Polydimethylsiloxane<br />

(PDMS) doped with conductive carbon black particles (Ding et al., 2007). This<br />

material has been used for chemical sensors (Andreadis et al., 2007) but not for a<br />

high-resolution pressure sensor. Prior work involving large pressure sensing arrays<br />

has concentrated on tactile pressure sensing, which requires reduced sensitivity<br />

<strong>and</strong> greater dynamic range (Someya et al., 2004; Harsanyi, 2000). In contrast,<br />

the carbon black pressure sensors are designed specifically for the lateral line<br />

application, where small pressure variations<br />

in the tens of Pascals are of<br />

interest.<br />

Pressure Sensor Design<br />

<strong>and</strong> Fabrication<br />

An individual pressure sensing cell<br />

is depicted <strong>and</strong> diagrammed in Figure<br />

9. Its active components consist<br />

of a resistive strain gauge patterned<br />

on the surface of a 10-mm-wide,<br />

1-mm-thick elastomer membrane,<br />

as shown in Figures 1A <strong>and</strong> 1B.<br />

PDMS was chosen as the membrane<br />

material due to its low tensile modulus<br />

(Schneider et al., 2008), which<br />

improves sensitivity. A differential<br />

pressure across the surfaces of the<br />

membrane causes the membrane to<br />

deflect. For small pressures, the deflection<br />

is linear (Senturia, 2001). The deflection<br />

then induces strain in the<br />

resistive strain gauge. The resulting resistance<br />

change is measured using the<br />

four terminals of the strain gauge,<br />

as depicted in Figures 9B <strong>and</strong> 9C. A<br />

constant current is applied through<br />

the outer terminals, <strong>and</strong> the voltage is<br />

measured across the inner voltage-tap<br />

terminals. This four-point probe measurement<br />

desensitizes the device to variations<br />

in contact resistance between<br />

the wires <strong>and</strong> the strain gauge. The<br />

voltage taps are positioned to capture<br />

the greatest resistance change at the<br />

central edge of the membrane where<br />

the greatest strain occurs (Senturia,<br />

2001).<br />

For a lateral line application on an<br />

underwater vehicle, the differential<br />

pressure between sensors in the array<br />

is of interest, but the depth of the<br />

aquatic vehicle, which corresponds to<br />

the absolute pressure, is not. The<br />

slow-varying absolute pressure may<br />

be cancelled out by low frequency<br />

pressure equilibration using the air<br />

July/August 2011 Volume 45 Number 4 141


FIGURE 9<br />

Diagrams <strong>and</strong> photo of a single pressure sensing cell with a 10-mm square membrane. The device<br />

is 3-mm thick.<br />

channel diagrammed in Figure 9A.<br />

This equilibration prevents damage<br />

to the sensors caused by the water<br />

depth at which the vehicle operates.<br />

It also relaxes the design requirements<br />

on the sensor, allowing it to have a<br />

thin membrane that is sensitive while<br />

not requiring it to withst<strong>and</strong> large<br />

pressures.<br />

In order to verify the performance<br />

of the carbon black sensing element<br />

in the context of a flexible pressure sensor,<br />

a single pressure sensing cell was<br />

fabricated (Figure 9C). PDMS was<br />

cast in a mold to form the membrane<br />

structure overhanging a cavity as<br />

showninFigure9A.Toformthe<br />

strain gauge, conductive carbon black<br />

particles approximately 1 μm in diameter<br />

were uniformly mixed into<br />

PDMS. Patterning of the strain gauge<br />

was accomplished with a stencil mask,<br />

which yielded a 100-μm-thick layer.<br />

The carbon black strain gauge terminals<br />

were bonded to aluminum wire<br />

with conductive epoxy. For testing<br />

purposes, a glass slide was used to simulate<br />

the rigid hull.<br />

Sensor Performance<br />

The dynamic response of the single<br />

test pressure sensor was characterized<br />

using a manual pressure source independently<br />

measured using a Honeywell<br />

pressure sensor. The resistance of<br />

the test sensor’s carbon black strain<br />

gauge was recorded as the pressure<br />

varied. Figures 10A <strong>and</strong> 10B show<br />

the resistance of the strain gauge as a<br />

function of time, compared to the<br />

pressure measured by the off-the-shelf<br />

sensor.<br />

From the dynamic response data in<br />

Figure 10, the sensitivity can be estimated<br />

as 0.55% change in resistance<br />

over 100 Pa. An amplification circuit<br />

was used to cancel the DC offset voltage<br />

<strong>and</strong> amplify the small resistance<br />

changes by 100, resulting in a roughly<br />

55% change in voltage over 100 Pa.<br />

This is more than sufficient for detecting<br />

pressure variations on the order of<br />

10 Pa, so it is adequate for the lateral<br />

line application.<br />

Figure 10C depicts the applied<br />

pressure plotted against the measured<br />

resistance for the dynamic tests in Figures<br />

10A <strong>and</strong> 10B, as well as a static<br />

test. The static test was performed by<br />

applying a series of pressure steps<br />

with a 1-min hold time <strong>and</strong> recording<br />

the resistance at the end. The pressure<br />

was stepped up, down, <strong>and</strong> up again to<br />

characterize sensor hysteresis. The<br />

slope of the curves represents the sensitivity<br />

of the sensor, <strong>and</strong> the asymmetry<br />

between the up <strong>and</strong> down steps<br />

is indicative of hysteresis. This is expected<br />

because the PDMS membrane<br />

is a viscoelastic material that experiences<br />

both creep <strong>and</strong> stress relaxation<br />

(Schneider et al., 2008). Both behaviors<br />

are byproducts of its low tensile modulus<br />

but are well understood <strong>and</strong> can be<br />

modeled <strong>and</strong> accounted for using signal<br />

processing techniques. The creep<br />

causes increased strain for a given pressure,<br />

resulting in the static response<br />

having a greater slope than the dynamic<br />

response. The dynamic responses are<br />

more linear <strong>and</strong> consistent from cycle<br />

to cycle than the static response, because<br />

the creep is not significant on<br />

the timescale of these tests.<br />

In addition to the polymer creep,<br />

there is the time-dependent relaxation<br />

behavior inherent in the conductive<br />

polymer. From the data in Figures<br />

10A <strong>and</strong> 10B, there is a very<br />

short time constant in the carbon<br />

black strain gauge resistance data<br />

when the pressure is being stepped<br />

up, but a much longer time constant<br />

when the pressure is being stepped<br />

down. This is thought to be a product<br />

of the way the carbon black conductive<br />

polymer responds to strain (Ding et al.,<br />

2007). However, as shown in dynamic<br />

response transfer curves of Figure 10C,<br />

142 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 10<br />

Dynamic <strong>and</strong> static responses of the single cell pressure sensor, normalized by the strain gauge<br />

resistance at atmospheric pressure R initial ∼ 20 kΩ. Parts A <strong>and</strong> B compare the output of an offthe-shelf<br />

Honeywell sensor with the output of the carbon black pressure sensor when the same<br />

pressure is applied to both. Part C presents dynamic transfer curves using the dynamic data<br />

from parts A <strong>and</strong> B as well as static data obtained by applying a constant pressure to the sensor<br />

<strong>and</strong> holding it for 1 min at each data point.<br />

simulations. Ultimately, the membrane<br />

thickness <strong>and</strong> the width of the<br />

strain gauge patterns will be the limiting<br />

factors. The carbon black–based<br />

flexible sensor array is a promising<br />

technology for the lateral line application<br />

because it meets the pressure <strong>and</strong><br />

spatial resolution criteria, it has repeatable<br />

<strong>and</strong> predictable performance, <strong>and</strong><br />

it can conform to the hull of an aquatic<br />

vehicle without protruding.<br />

this behavior is repeatable <strong>and</strong> consistent<br />

during cyclic loading, so future<br />

work will consist of using signal processing<br />

to compensate for this effect.<br />

Creating a Sensor Array<br />

Extending the single carbon black<br />

sensor cell into an array involves both<br />

miniaturization of each sensor cell<br />

<strong>and</strong> routing of the carbon black strain<br />

gauges. A prototype array of four pressure<br />

sensing cells is shown in Figure<br />

11. The four sensors share two<br />

common current terminals, <strong>and</strong> each<br />

individual sensor has two voltage<br />

taps, reducing the necessary wiring.<br />

Each cell has a 5-mm-wide, 0.5-mmthick<br />

square membrane. The dimensions<br />

of the cells have been scaled<br />

down by a factor of 2 from the single<br />

test pressure sensor. The spatial resolution<br />

of the array is determined by the<br />

sensor cell spacing, which is 7-mm<br />

center-to-center for this array.<br />

Further miniaturization is possible<br />

by reducing the dimensions of the<br />

membrane <strong>and</strong> strain gauge pattern<br />

in order to achieve the sub-5 mm spacing<br />

found optimal in the active sensing<br />

Conclusion<br />

This paper has considered several<br />

aspects of object identification based<br />

entirely on a lateral-line-like pressure<br />

sensor array, with the aim of examining<br />

its potential for application<br />

in underwater vehicles. Object identification<br />

was divided into passive object<br />

identification, where an external flow<br />

interacts with the object to stimulate<br />

the sensor array, <strong>and</strong> active object<br />

identification, where self-generated<br />

flow is used to interrogate an object<br />

in still water. In each case, experiments<br />

were used to examine the problem of<br />

estimating the position, size, <strong>and</strong><br />

shape of a cylinder based on a linear<br />

array of pressure sensors in a realistic<br />

noise environment. Constraints with<br />

off-the-shelf sensors in the experimental<br />

implementation of an artificial lateral<br />

line, have led to the development<br />

of a flexible artificial lateral line that<br />

makes use of a new polymer sensing<br />

technology.<br />

The passive sensing of a moving<br />

cylinder separated the identification<br />

of the position <strong>and</strong> size from the<br />

shape into different experiments. The<br />

wake generated by the motion of<br />

the cylinder was a critical component<br />

in the success of each part. For the cylinder<br />

tracking, a simple steady potential<br />

model of the wake was sufficient<br />

to allow a particle filter to track the<br />

July/August 2011 Volume 45 Number 4 143


FIGURE 11<br />

Diagram <strong>and</strong> photo of an array of four pressure sensors, each with a 5-mm transparent square<br />

membrane. The sensors share common current terminals. Each individual sensor has two voltage<br />

tap terminals to measure the resistance of the part of the carbon black strain gauge located on the<br />

corresponding membrane.<br />

position <strong>and</strong> size in real time. In order<br />

to distinguish between a square <strong>and</strong><br />

circular cylinder, a decision rule derived<br />

from PCA was highly successful.<br />

This rule demonstrated robustness in<br />

dealing with cylinders of different<br />

sizes, distances, <strong>and</strong> speeds. In analyzing<br />

the decision rule, it was found that<br />

pressure data in the highly variable<br />

region near <strong>and</strong> aft of the flow separation<br />

is vital in determining the correct<br />

shape. This implies limitations<br />

in the types of shapes that can be distinguished<br />

passively, as dissimilar<br />

shapes with similar wakes may cause<br />

difficulty.<br />

By recreating an active sensing encounter<br />

with a foil st<strong>and</strong>ing in for the<br />

vehicle or fish, we demonstrated the<br />

possibility of identifying the location<br />

<strong>and</strong> shape of a cylinder without prior<br />

knowledge of its shape. A potential<br />

flow model was chosen for simplicity<br />

<strong>and</strong> to avoid heavy calculations that<br />

would make it impossible to run the algorithm<br />

in real time. Using this model<br />

with an UKF allowed us to obtain a<br />

very good object location estimate<br />

<strong>and</strong> reasonable ellipse identification<br />

using three pressure sensors. The results<br />

stress the importance of the<br />

head lateral line of the fish (that had<br />

been emphasized in the case of the<br />

fish moving towards an obstacle by<br />

Hassan, 1986) when it passes an object,<br />

due to the timing <strong>and</strong> location<br />

of flow separation. To make more use<br />

of a biomimetic trunk lateral line, the<br />

next step is to develop a model that<br />

takes the separation into account to<br />

be able to extract information from<br />

the second half of the data.<br />

For both active <strong>and</strong> passive sensing,<br />

we have found that information about<br />

thelocation<strong>and</strong>sizeofacylinderis<br />

available via an artificial lateral line.<br />

Beyond this point, however, there are<br />

strong differences in the two scenarios.<br />

In the case of passive sensing, the flow<br />

separation <strong>and</strong> wake are integral components<br />

to the pressure distribution<br />

<strong>and</strong> must be accounted for from the<br />

beginning. In fact, it appears the<br />

shape information may largely be available<br />

through these components. In<br />

contrast, for active sensing flow separation<br />

does not occur immediately,<br />

allowing it to be ignored in the initial<br />

shape estimation as done in this<br />

paper. Using pre-separation data limits<br />

the number of measurements<br />

available but allows the use of a general<br />

model for identifying arbitrary<br />

shapes.<br />

Simulations based on the active<br />

sensing experiments have shown that<br />

increasing the sensor spacing of<br />

approximately 0.03 body lengths<br />

achieves the fastest <strong>and</strong> most accurate<br />

object identification for cylindrical objects<br />

with a diameter on the same scale<br />

as the sensing body. Coupled with the<br />

sensitivity (less than 10 Pa) required to<br />

suitably measure the pressure signals in<br />

the experiments, this defines the specifications<br />

for an artificial lateral line<br />

that could be used to identify or locate<br />

similar objects for an underwater<br />

vehicle. Of significant additional concern<br />

is the need for a sensor that can<br />

be easily mounted to a hull without<br />

significantly disturbing the flow.<br />

Based on the design <strong>and</strong> test results<br />

shown, the carbon black-based flexible<br />

sensor array is a promising technology<br />

for the lateral line application. Used as<br />

a sensing element, it can meet the pressure<br />

<strong>and</strong> spatial resolution criteria, it<br />

has repeatable <strong>and</strong> predictable performance,<br />

<strong>and</strong> it is able to conform to<br />

the smooth curves of a hull while protruding<br />

only 3 mm.<br />

144 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Acknowledgments<br />

The authors gratefully acknowledge<br />

the support of the Singapore-MIT<br />

Alliance for Research <strong>and</strong> <strong>Technology</strong><br />

(SMART) program’s Center for Environmental<br />

Sensing <strong>and</strong> Modeling<br />

<strong>and</strong> that of the National Oceanic <strong>and</strong><br />

Atmospheric Administration’s Sea<br />

Grant program under project number<br />

R/RT-2/RCM-17.<br />

Lead Author:<br />

Vicente I. Fern<strong>and</strong>ez<br />

Department of<br />

Mechanical Engineering,<br />

Massachusetts Institute of <strong>Technology</strong><br />

5-424, 77 Massachusetts Avenue<br />

Cambridge, MA 02139<br />

Email: vicentef@mit.edu<br />

References<br />

Andreadis, N., Chatz<strong>and</strong>roulis, S., Goustouridis,<br />

D.,Kosma,V.,Beltsios,K.,&Raptis,I.2007.<br />

Fabrication of conductometric chemical sensors<br />

by photolithography of conductive polymer<br />

composites. J Microelec Eng. 84(5-8):<br />

1211-4. doi: 10.1016/j.mee.2007.01.050.<br />

Bouffanais, R., Weymouth, G.D., & Yue,<br />

D.K.P. 2011. Hydrodynamic object recognition<br />

using pressure sensing. P R Soc A. 467(2125):<br />

19-38. doi: 10.1098/rspa.2010.0095.<br />

Branko, R., Arulampalam, S., & Gordon, N.<br />

2004. Beyond the Kalman Filter: Particle<br />

Filters for Tracking Applications. Boston:<br />

Artech House. 318 pp.<br />

Burt de Perera, T. 2004. Spatial parameters<br />

encoded in the spatial map of the blind<br />

Mexican cave fish, Astyanax fasciatus.<br />

Anim Behav. 68(2):291-5. doi: 10.1016/<br />

j.anbehav.2003.11.009.<br />

Cappe, O.,Moulines,E.,&Ryden,T.2005.<br />

InferenceinHiddenMarkovModels.NewYork:<br />

Springer Science+Business Media, Inc. 652 pp.<br />

Chen, J., Engel, J., Chen, N., P<strong>and</strong>ya, S.,<br />

Coombs, S., & Liu, C. 2006. Artificial lateral<br />

line <strong>and</strong> hydrodynamic object tracking. In:<br />

Micro Electro Mechanical Systems 2006,<br />

MEMS 2006 Istanbul. 19th IEEE International<br />

Conference on Micro Electro Mechanical<br />

Systems. pp. 694-7. Istanbul: IEEE.<br />

doi: 10.1109/MEMSYS.2006.1627894.<br />

Coombs, S. 2001. Smart skins: Information<br />

processing by lateral line flow sensors. Auton<br />

Rob. 11(3):255-61.<br />

Coombs, S., & Janssen, J. 1990. Behavioral<br />

<strong>and</strong> neurophysiological assessment of the lateral<br />

line sensitivity in the mottled sculpin,<br />

Cottus bairdi. J Comp Physiol A. 167(4):<br />

557-67. doi: 10.1007/BF00190827.<br />

Curcic-Blake, B., & van Netten, S. 2006.<br />

Source location encoding in the fish lateral<br />

line canal. J Exp Biol. 209(8):1548-59.<br />

doi: 10.1242/jeb.02140.<br />

Ding, T., Wang, L., & Wang, P. 2007.<br />

Changes in electrical resistance of carbonblack-filled<br />

silicone rubber composite during<br />

compression. J Polym Sci Part B Polym<br />

Phys. 45(19):2700-6. doi: 10.1002/<br />

polb.21272.<br />

Gelb, A. 1974. Applied Optimal Estimation.<br />

Massachusetts: MIT Press. 382 pp.<br />

Goulet, J., Engelmann, J., Chagnaud, B.,<br />

Franosch, J., Suttner, M., & Hemmen, J.<br />

2007. Object localization through the lateral<br />

line system of fish: Theory <strong>and</strong> experiment.<br />

J Comp Phys A 194(1):1-17. doi: 10.1007/<br />

s00359-007-0275-1.<br />

Harsanyi, G. 2000. Polymer films in sensor<br />

applications: A review of present uses <strong>and</strong><br />

future possibilities. Sens Rev. 20(2):98-105.<br />

doi: 10.1108/02602280010319169.<br />

Hassan, E.S. 1985. Mathematical analysis<br />

of the stimulus for the lateral line organ.<br />

Biol Cybern. 52(1):23-36. doi: 10.1007/<br />

BF00336932.<br />

Hassan, E.S. 1986. On the discrimination<br />

of spatial intervals by the blind cave fish<br />

(Anoptichthys jordani). J Comp Physiol A.<br />

159(5):701-10. doi: 10.1007/BF00612042.<br />

Hassan, E.S. 1992. Mathematical description<br />

of the stimuli to the lateral line system derived<br />

from a three dimensional flow field analysis:<br />

I. The case of moving in open water <strong>and</strong> of<br />

gliding toward a plane surface. Biol Cybern.<br />

66:443-52. doi: 10.1007/BF00197725.<br />

Jackson, J.E. 2003. A User’s Guide to<br />

Principal Components. New Jersey: Wiley.<br />

592 pp.<br />

Jolliffe, I.T. 2002. Principal Component<br />

Analysis. New York: Springer. 502 pp.<br />

Julier, S.J., & Uhlmann, J.K. 2004. Unscented<br />

filtering <strong>and</strong> nonlinear estimation.<br />

Proc IEEE. 92(3):401-22. doi: 10.1109/<br />

JPROC.2003.823141.<br />

Montgomery, J.C., Coombs, S., & Baker,<br />

C.F. 2001. The mechanosensory lateral line<br />

system of the hypogean form of Astyanax<br />

fasciatus. Eviron Biol Fishes. 62(1-3):87-96.<br />

doi: 10.1023/A:1011873111454.<br />

Pohlmann, K., Atema, J., & Breithaupt, T.<br />

2004. The importance of the lateral line<br />

in nocturnal predation of piscivorous<br />

catfish. J Exp Biol. 207(17):9271-8.<br />

doi: 10.1242/jeb.01129.<br />

Schneider, F., Fellner, T., Wilde, J., &<br />

Wallrabe, U. 2008. Mechanical properties<br />

of silicones for MEMS. J Micromech<br />

Microeng. 18(6):065008. doi: 10.1088/<br />

0960-1317/18/6/065008.<br />

Senturia, S.D. 2001. Microsystem Design.<br />

Boston: Kluwer Academic Publishers.<br />

720 pp.<br />

Sichert, A., Bamler, R., & van Hemmen, J.<br />

2009. Hydrodynamic object recognition:<br />

When multipoles count. Phys Rev Lett.<br />

102(5):058104. doi: 10.1103/<br />

PhysRevLett.102.058104.<br />

Someya, T., Sekitani, T., Iba, S., Kato, Y.,<br />

Kawaguchi, H., & Sakurai, T.A. 2004.<br />

Large-area, flexible pressure sensor matrix<br />

with organic field-effect transistors for artificial<br />

skin applications. P Natl Acad Sci.<br />

101(27):9966-70. doi: 10.1073/pnas.<br />

0401918101.<br />

van Netten, S. 2006. Hydrodynamic detection<br />

by cupulae in a lateral line canal:<br />

functional relations between physics <strong>and</strong><br />

July/August 2011 Volume 45 Number 4 145


physiology. Biol Cybern. 94(1):67-85.<br />

doi: 10.1007/s00422-005-0032-x.<br />

Vogel, D., & Bleckmann, H. 2000. Behavioral<br />

discrimination of water motions caused<br />

by moving objects. J Comp Physiol A. 186(12):<br />

1107-17. doi: 10.1007/s003590000158.<br />

von Campenhausen, C., Riess, I., &<br />

Weissert, R. 1981. Detection of stationary<br />

objects by the blind cave fish Anoptichthys<br />

jordani (Characidae). J Comp Physiol A Sens<br />

Neural Behav Physiol. 143(3):369-74.<br />

doi: 10.1007/BF00611175.<br />

Wan, E., & van der Merwe, R. 2001.<br />

The unscented Kalman filter. In: Kalman<br />

filtering <strong>and</strong> neural networks, ed. Haykin, S.,<br />

221-80. New York: Wiley. doi: 10.1002/<br />

0471221546.ch7.<br />

Weissert, R., & von Campenhausen, C.<br />

1981. Discrimination between stationary<br />

objects by the blind cave fish Anoptichthys<br />

jordani (characidae). J Comp Physiol A<br />

Sens Neural Behav Physiol. 143(3):375-81.<br />

doi: 10.1007/BF00611176.<br />

Windsor, S.P., & McHenry, M.J. 2009. The<br />

influence of viscous hydrodynamics on the<br />

fish lateral-line system. Integr Comp Biol.<br />

49(6):691-701. doi: 10.1093/icb/icp084.<br />

Yang, Y., Chen, N., Tucker, C., P<strong>and</strong>ya, S.,<br />

Jones, D., & Liu, C. 2008. Biomimetic flow<br />

sensing using artificial lateral lines. ASME<br />

Conf Proc. 2007(43025):1331-8.<br />

146 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

A Conserved Neural Circuit-Based<br />

Architecture for Ambulatory <strong>and</strong><br />

Undulatory Biomimetic Robots<br />

AUTHORS<br />

Joseph Ayers<br />

Anthony Westphal<br />

Daniel Blustein<br />

Department of Biology <strong>and</strong><br />

<strong>Marine</strong> Science Center,<br />

Northeastern University<br />

Introduction<br />

The innate behavior of underwater<br />

animals provides an effective model for<br />

the adaptive behavior of unmanned<br />

underwater vehicles (Ayers, 2004).<br />

Underwater animals must respond to<br />

a broad variety of environmental challenges<br />

including turbidity, hydrodynamic<br />

flow, heterogeneous <strong>and</strong><br />

highly structured bottom types <strong>and</strong><br />

impediment. Their relatively neutral<br />

buoyancy renders them especially susceptible<br />

to hydrodynamic perturbation.<br />

As a result, they have evolved a<br />

behavioral set that includes a broad<br />

variety of compensatory responses to<br />

perturbation. This behavioral set results<br />

from layered exteroceptive reflexes<br />

responding to exteroceptive<br />

sensor input resulting from changes<br />

in orientation relative to gravity, impediment,<br />

chemical cues, <strong>and</strong> hydrodynamic<br />

<strong>and</strong> optical flow (Ayers,<br />

2004; Blustein & Ayers, 2010).<br />

These layered exteroceptive reflexes can<br />

form taxic responses to point sources<br />

of sound or chemicals (Westphal<br />

et al., 2011). As the point sources<br />

form motivational cues for goal achieving<br />

behavioral sequences, they can<br />

ABSTRACT<br />

The adaptive capabilities of underwater organisms result from layered exteroceptive<br />

reflexes responding to gravity, impediment, <strong>and</strong> hydrodynamic <strong>and</strong> optical<br />

flow. In combination with taxic responses to point sources of sound or chemicals,<br />

these reflexes allow reactive autonomy in the most challenging of environments.<br />

We are developing a new generation of lobster <strong>and</strong> lamprey-based robots that operate<br />

under control by synaptic networks rather than algorithms. The networks,<br />

based on the comm<strong>and</strong> neuron, coordinating neuron, <strong>and</strong> central pattern generator<br />

architecture, code sensor input as labeled lines <strong>and</strong> activate shape memory alloybased<br />

artificial muscles through a simple interface that couples excitation to contraction.<br />

We have completed the lamprey-based robot <strong>and</strong> are adapting this sensor,<br />

board, <strong>and</strong> actuator architecture to a new generation of the lobster-based robot. The<br />

networks are constructed from discrete time map-based neurons <strong>and</strong> synapses <strong>and</strong><br />

are instantiated on the digital signal processing chip. A sensor board integrates inputs<br />

from a short baseline sonar array (for beacon tracking <strong>and</strong> supervisory control),<br />

accelerometer, a compass, antennae, <strong>and</strong> optionally chemosensors. Actuator<br />

control is mediated by pulse-width duty cycle coding generated by the electronic<br />

motor neurons <strong>and</strong> a comparator <strong>and</strong> power field-effect transistor (FET) system<br />

housed on low- <strong>and</strong> high-current driver boards. These circular boards are stacked<br />

in a tubular hull with the processor <strong>and</strong> batteries. This system can readily mimic the<br />

biomechanics of the model organisms by the addition of hydrodynamic control surfaces.<br />

The behavioral set results from chaining sequences of exteroceptive reflexes<br />

released by sensory feedback from the environment.<br />

Keywords: biomimetic, robot, UUV, lobster, lamprey<br />

guide reactive autonomy in the most<br />

challenging of environments. The<br />

task is to capture these performance<br />

advantages in engineered devices.<br />

The Biological Model<br />

We are developing a new generation<br />

of lobster <strong>and</strong> lamprey-based<br />

robots that operate under control by<br />

synaptic networks rather than algorithms.<br />

Previous generations of these<br />

vehicles were controlled by finite<br />

state machines that were organized<br />

around the elements of the corresponding<br />

neurobiological models (Ayers<br />

et al., 2000; Ayers & Witting, 2007).<br />

The neuronal circuits that control<br />

our current generation of vehicles are<br />

based on the comm<strong>and</strong> neuron, coordinating<br />

neuron, central pattern generator<br />

(CCCPG) architecture (Figures 1a<br />

<strong>and</strong> 1b) of innate animal behavior<br />

(Kennedy & Davis, 1977; Stein, 1978;<br />

Pearson, 1993). The networks are organized<br />

into segmental central pattern<br />

generators (CPGs) that control appendages<br />

or axial body musculature in the<br />

July/August 2011 Volume 45 Number 4 147


FIGURE 1<br />

CCCPG architecture with exteroceptive reflex. Labeled circles represent neurons; synapses are<br />

shown as connecting lines with triangular (excitatory) or circular (inhibitory) endpoints. (a) Neuronal<br />

circuit-based controller for a walking lobster robot. The effector organs of each body segment<br />

are controlled by CPGs that contain a neuronal oscillator, a pattern generator <strong>and</strong> sets of<br />

motor neuron pools. The CPGs are coordinated among themselves by a set of coordinating neurons<br />

(CoN) that provide information about the activity status of a governing oscillator to a governed<br />

oscillator. The CPGs are brought into operation by a set of comm<strong>and</strong> neurons (CN) that<br />

initiates their operation <strong>and</strong> controls their average frequency <strong>and</strong> amplitude. (b) Neuronal<br />

circuit-based controller for a swimming lamprey robot. Slight modification of the CCCPG architecture<br />

<strong>and</strong> effectors transforms the system’s motor output from walking to swimming. (c) The<br />

CNs are organized into exteroceptive reflexes that are released by neuronally coded sensor information<br />

(rounded rectangles: heading from a compass, target orientation from SBA) through sensory<br />

interneurons, which mediate in place rotation <strong>and</strong> yaw during locomotion.<br />

animal models <strong>and</strong> the robots (Ayers<br />

et al., 2010). The CPGs are coordinated<br />

among themselves by a category<br />

of neurons called coordinating neurons<br />

that pass status information<br />

from a governing CPG to a governed<br />

CPG that alters its period to remain<br />

coordinated at a particular phase,<br />

depending on the ratio of intrinsic frequencies<br />

of the governing <strong>and</strong> governed<br />

CPGs (Selverston & Ayers,<br />

2006). This temporal resetting occurs<br />

on a cycle-by-cycle basis to entrain the<br />

CPGs in a particular gait in the case of<br />

walking or to ensure propagation of a<br />

wave of flexion down the body during<br />

undulation (Figure 1).<br />

The CPGs are brought into operation<br />

<strong>and</strong> modulated by a category of<br />

neurons called comm<strong>and</strong> neurons<br />

(Kupfermann & Weiss, 1978). Comm<strong>and</strong><br />

neurons generally constitute the<br />

locus at which the decision to evoke a<br />

behavioral act is made <strong>and</strong> project<br />

from the brain through the central nervous<br />

system to bring the segmental<br />

CPGs into operation. They typically<br />

perform this process through the<br />

mechanism of neuromodulation<br />

through second messengers that alter<br />

both the cellular properties <strong>and</strong> synaptic<br />

connectivity within the CPG<br />

(Dickinson, 2006). By this mechanism<br />

or through direct synaptic<br />

modulation, the same CPG can often<br />

produce variations on a behavioral act<br />

in response to different comm<strong>and</strong>s<br />

(Selverston & Ayers, 2006).<br />

The sensors we employ are configured<br />

to encode sensory input as a labeled<br />

line code (Bullock, 1968). In this<br />

form of coding, each sensory neuron is<br />

a unique source of information. The<br />

information consists of (1) the nature<br />

of the sensory stimulus (light, optical<br />

flow, chemicals, bumps, etc.), (2) the<br />

receptive field or position of the stimulus<br />

on the body <strong>and</strong> (3) the magnitude<br />

of the stimulus coded as an<br />

action potential train where the frequency<br />

of the action potentials is<br />

proportional to the logarithm of the<br />

stimulus intensity. We configure<br />

these sensory elements in networks<br />

that filter out features of the environment<br />

through lateral inhibition,<br />

range fractionation <strong>and</strong> motion detection.<br />

These filtered outputs provide<br />

input to the comm<strong>and</strong> neurons to<br />

release behavior.<br />

We have completed the lampreybased<br />

robot <strong>and</strong> are adapting this sensor,<br />

board, <strong>and</strong> actuator architecture to<br />

a new generation of the lobster-based<br />

robot (Figure 2). The lamprey robot<br />

features an electronic nervous system<br />

that we are adapting to the new<br />

lobster-based robot. The key feature<br />

of this architecture is that it is generalizable<br />

between all animal models.<br />

Electronic Nervous Systems<br />

Thediscretetimemap-based<br />

(DTM) neuron <strong>and</strong> synapse equations<br />

(Rulkov, 2002) phenomenologically<br />

modelneuronalactivity.Themodel<br />

has two state variables x <strong>and</strong> y, two<br />

FIGURE 2<br />

Biomimetic robots. (a) Fourth generation lobsterbased<br />

robot. (b) Second generation lampreybased<br />

robot.<br />

148 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


control parameters α <strong>and</strong> σ, <strong>and</strong>a<br />

parameter β for integrated synaptic<br />

input. Variations in α <strong>and</strong> σ can configure<br />

neurons into a silent type, a<br />

spiking type, a bursting type <strong>and</strong> a<br />

chaotic type (Ayers & Rulkov, 2007).<br />

Similar control parameters for the synapse<br />

instruments determine the synaptic<br />

strength, relaxation rate, release<br />

threshold, <strong>and</strong> reversal potential that<br />

determine whether the synapse is excitatory<br />

or inhibitory. The electronic nervous<br />

systems are first prototyped in<br />

the National Instruments LabVIEW<br />

software. Neuron <strong>and</strong> synapse instruments<br />

are configured with different<br />

properties, <strong>and</strong> the modeled<br />

neurons <strong>and</strong> synapses are wired together<br />

in LabVIEW.<br />

Figure 3 demonstrates a simple<br />

CPG circuit configured to illustrate<br />

operation of the four types of neurons<br />

in our CPGs. Here, a comm<strong>and</strong> neuron<br />

(1) initiates an oscillation between<br />

a bursting neuron (2) <strong>and</strong> a spiking<br />

follower (3) using a slow modulatory<br />

synapse. A fourth coordinating neuron<br />

(4) can be activated in bursts to entrain<br />

the bursting pattern evoked by the<br />

comm<strong>and</strong>. In contrast to coordinating<br />

neurons that reset the timing of the oscillation<br />

on a cycle-by-cycle basis by<br />

perturbation, comm<strong>and</strong> neurons initiate<br />

operation of the circuits <strong>and</strong> modulate<br />

their average frequency <strong>and</strong> amplitude<br />

as parameters (Figure 3b–c).<br />

FIGURE 3<br />

DTM network integration. (a) The modeled neuronal circuit: (1) comm<strong>and</strong> neuron, (2) bursting<br />

neuron, (3) spiking neuron, (4) coordinating neuron. (b) Parametric modulation of 2 <strong>and</strong> 3 generates<br />

an antagonistic bursting pattern. Voltage vs. simulation iteration traces are shown for each<br />

component of the network. The trace below neuron 1 shows the current injected to initiate activity.<br />

(c) Perturbation of the bursting pattern in 2 by a coordinating neuron (3) to entrain the evoked<br />

rhythm. The trace below 4 shows the injected current into that neuron. Adapted from Ayers <strong>and</strong><br />

Rulkov (2007).<br />

FIGURE 4<br />

Configuration of robots. (a) Current implementation of board set of the lamprey-based robot. The<br />

sonar stack processes the analog hydrophone signals from the short baseline array (SBA). The<br />

electronic nervous system is instantiated on a Texas Instruments TMS320C6727 chip on the CNS<br />

DSP board. Sensors <strong>and</strong> the SBA processor are housed on the sensor array board. Low- <strong>and</strong> highcurrent<br />

drivers provide the current pulse trains that activate the nitinol actuators. The robot operates<br />

on a 12-V, 4.5-Ah NiMH battery pack. (b) Configuration of the lamprey robot. A hinge in the<br />

pitch module allows the undulator to alter its pitch relative to the hull to dive or climb. (c) Configuration<br />

of the lobster robot. A tubular hull similar to the lamprey robot houses the electronics<br />

<strong>and</strong> batteries. Anterior claw-like <strong>and</strong> posterior abdomen-like hydrodynamic control surfaces provide<br />

a thrust vector into the substrate to increase traction. Externally mounted sensors include<br />

optical flow detectors, hydrodynamic flow sensors on the antennae, <strong>and</strong> sonar transducers for<br />

beacon tracking <strong>and</strong> supervisory input.<br />

Board Architecture<br />

The networks that control the robots<br />

are constructed from DTM neurons<br />

<strong>and</strong> synapses in procedural C <strong>and</strong><br />

are instantiated on a Texas Instruments<br />

digital signal processing (DSP)<br />

chip. A common board architecture<br />

is used to control both the swimming<br />

<strong>and</strong> walking robots (Figure 4a).<br />

The board set consists of four types:<br />

July/August 2011 Volume 45 Number 4 149


(1) The DSP board houses the DSP<br />

chip <strong>and</strong> interconnects to sensor <strong>and</strong><br />

actuator boards. (2) A sensor array<br />

houses a compass, inclinometers, accelerometers<br />

<strong>and</strong> a processor board to<br />

derive azimuth <strong>and</strong> inclination deviation<br />

signals from the short baseline<br />

array stack. (3) A low-current driver<br />

board receives logic signals from the<br />

motor neuron output from the DSP<br />

chip <strong>and</strong> in turn controls (4) a highcurrent<br />

driver that applies current to<br />

heat the individual nitinol actuators.<br />

Artificial muscles constructed from<br />

the shape memory alloy nitinol move<br />

both the walking legs <strong>and</strong> the undulatory<br />

body axis. The nitinol is operated<br />

on a thermal cycle. When cooled by<br />

seawater, it can be deformed into martensite<br />

state that is associated with<br />

about a 5% length increase. When<br />

heated by electrical current, it transforms<br />

into the austenite state <strong>and</strong> contracts<br />

rapidly. Increases in the length<br />

of one muscle are produced by the<br />

contraction of its antagonist. Both<br />

the amplitude <strong>and</strong> velocity of the contractions<br />

can be graded by pulse-width<br />

duty cycle modulation of the drive<br />

pulses. Excitation-contraction coupling<br />

with the motor neurons is mediated<br />

by a comparator circuit on the low<br />

current boards that thresholds the<br />

action potentials to generate a square<br />

wave pulse that controls a power on<br />

the high-current boards to activate<br />

the actuators. Changes in the firing frequency<br />

of the motor neurons provide<br />

the duty cycle modulation.<br />

A common DSP board (Figure 4a)<br />

interfaces the sensor array board to<br />

the current driver boards. A separate<br />

regulator board provides the load to<br />

these boards via a 12 V NiMh battery<br />

pack. Feed-through connectors in the<br />

end caps lead the current conductors<br />

to the actuators. The boards are<br />

stacked in a tubular hull whose length<br />

can be varied to accommodate a variety<br />

of mission packages.<br />

Behavioral Set<br />

This system can readily mimic the<br />

biomechanics of the model organisms<br />

by the addition of hydrodynamic control<br />

surfaces. Turns in the undulatory<br />

robot are mediated by modulation<br />

of the amplitude of the flexions to<br />

thetwosidesasintheanimalmodel<br />

(Ayers, 1989). The direction of propagation<br />

of the flexion waves along the<br />

body axis can be reversed to mediate<br />

backward swimming. Dives <strong>and</strong><br />

climbs can be mediated by alteration<br />

of the pitch of the hull relative to the<br />

undulator (Figure 4b). Dorsal flexion<br />

of the hull generates a low-pressure<br />

area above the hull to mediate a climb<br />

while ventral flexion generates a lowpressure<br />

area below the hull to mediate<br />

diving.<br />

The primary response to hydrodynamic<br />

flow in the lobster is to orient<br />

into the flow, lower the anterior control<br />

surfaces <strong>and</strong> elevate the posterior<br />

control surfaces. As the lobster is only<br />

slightly negatively buoyant, this creates<br />

a thrust vector into the substrate <strong>and</strong><br />

increases the traction of the legs on<br />

the bottom. The three degree of freedom<br />

walking legs of the lobster robot<br />

allow the vehicle to walk in all directions<br />

(Ayers & Witting, 2007). Alterations<br />

in the degree of depression can<br />

regulate the height above the substrate,<br />

while variations along the long body<br />

axis regulates pitch. Biasing the depression<br />

on the two sides can correspondingly<br />

regulate roll to maintain primary<br />

orientation on tilted substrates.<br />

The behavioral set of both robots<br />

is organized around exteroceptive reflexes<br />

(Kennedy & Davis, 1977). An<br />

innate releasing mechanism composed<br />

of sensory neurons <strong>and</strong> interneurons<br />

filters incoming information to extract<br />

relevant features of the environment<br />

such as bumps, tilt, hydrodynamic<br />

<strong>and</strong> optical flow (Figures 1c <strong>and</strong> 5a).<br />

These sensory releasers are coded in interneurons<br />

that, in turn, activate comm<strong>and</strong><br />

systems. The interneurons use<br />

lateral inhibition from low-threshold<br />

to high-threshold elements to produce<br />

range fractionation so that different<br />

ranges of a scalar input are coded by<br />

different sensory neurons, providing<br />

the capability for detailed circuit logic.<br />

An example of such exteroceptive<br />

reflexes are those involved in the mediation<br />

of the yaw plane responses to<br />

hydrodynamic flow that occur during<br />

rheotaxis. Lobsters typically walk<br />

with their antenna projected to the<br />

front (Figure 5a, I). If wave surge<br />

occurs from the side, it bends the<br />

upstream antenna medially <strong>and</strong> the<br />

downstream antenna laterally (Figure<br />

5a, II). Our hypothesis is that<br />

this perturbation activates a rheotaxic<br />

interneuron that activates the backward<br />

walking comm<strong>and</strong> on the upstream<br />

side <strong>and</strong> the forward walking<br />

comm<strong>and</strong> on the downstream side.<br />

This would cause the animal/vehicle<br />

to rotate in place into the flow. While<br />

orienting into flow, the animals project<br />

their antenna laterally, which would<br />

switch control to another bilateral<br />

pair of surge interneurons (Figure 5a,<br />

III). As the most upstream antennae<br />

would be bent more than the more<br />

downstream antenna, <strong>and</strong> these interneurons<br />

project to the contralateral<br />

forward walking comm<strong>and</strong>s, the<br />

animal/vehicle would continue to<br />

yaw into the flow until current to the<br />

two antenna is balanced ensuring<br />

proper orientation into the flow for<br />

maximal hydrodynamic stability. The<br />

hydrodynamic control surfaces can<br />

then ensure proper traction to overcome<br />

the perturbation.<br />

150 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 5<br />

Neural simulation of rheotaxis. (a) Epochs of a lobster’s rheotaxic behavioral response to water<br />

surge shown with the predominant active neural reflex circuit. In the network diagrams, top circles<br />

represent sensory neurons corresponding to high (H), medium (M), or low (L) antennal bending in<br />

the lateral or medial direction. Black ovals represent interneurons that project to bilateral comm<strong>and</strong>s<br />

for forward (F) or backward (B) walking. In I, as a lobster walks forward, bilaterally balanced<br />

low lateral bending of the antennae is observed, which serves to sustain forward locomotion. In II,<br />

left-to-right water surge (blue arrows) causes a high medial bend of the ipsilateral antenna <strong>and</strong> a<br />

high lateral bend of the contralateral antenna eliciting rheotaxis. In III, bilaterally asymmetrical<br />

lateral antennal bending mediates yawing upstream during forward walking. (b) Voltage vs.<br />

time traces for the neurons of the rheotaxis circuit. Dashed lines distinguish the behavioral epochs<br />

shown in a.<br />

alter locomotory outputs of a network,<br />

as we have shown to produce both<br />

swimming <strong>and</strong> walking. Even hybrid<br />

systems can be achieved, such as the<br />

gait of an alligator resulting from a<br />

combination of the lamprey <strong>and</strong> lobster<br />

CPG networks. Sensor packages<br />

can be adapted with little restructuring.<br />

Layered exteroceptive reflex<br />

networks provide capabilities for navigation,<br />

investigation <strong>and</strong> obstacle negotiation<br />

in unpredictable near-shore<br />

marine environments with a minimum<br />

of supervisory control.<br />

Lead Author:<br />

Joseph Ayers<br />

Department of Biology <strong>and</strong><br />

<strong>Marine</strong> Science Center<br />

Northeastern University,<br />

East Point, Nahant MA 01908<br />

Email: lobster@neu.edu<br />

This overall control scheme applies<br />

to a variety of environmental circumstances<br />

<strong>and</strong> perturbations in the yaw,<br />

pitch, <strong>and</strong> roll planes. Many exteroceptive<br />

reflexes form taxic systems.<br />

For example, the three hydrophone<br />

short baseline sonar array (SBA) on<br />

the lamprey robot reports the deviation<br />

of the sonar beacon relative to<br />

the hull orientation in terms of inclination<br />

<strong>and</strong> azimuth (Westphal et al.,<br />

2011). The azimuthal signal modulates<br />

swim comm<strong>and</strong> systems to<br />

cause the vehicle to yaw toward the<br />

beacon (Figure 1c) while the inclination<br />

signal modulates the pitch system<br />

to cause the vehicle to climb/dive<br />

toward the beacon. Taken together<br />

these layered reflexes will cause the<br />

vehicle to home on a sonar beacon. A<br />

similar 2-D SBA is planned for the<br />

lobster robot to control yaw taxis.<br />

The sonar transducers also provide<br />

a capability for supervisory control.<br />

The vehicles will be sent supervisory<br />

comm<strong>and</strong>s that specify a heading <strong>and</strong><br />

odometry information for distance.<br />

The comm<strong>and</strong> will include a propensity<br />

to negotiate or investigate obstacles<br />

depending on the mission. At the<br />

end of the search vector, the vehicle<br />

would annunciate its location to a<br />

long baseline sonar array <strong>and</strong> be sent<br />

a new search vector. By this mechanism<br />

an operator could supervise several<br />

robots simultaneously.<br />

Conclusion<br />

The neuronal mechanisms of<br />

innate behavior can be applied to a<br />

broad variety of biomimetic underwater<br />

robots. Minimal modification<br />

of neuronal components can easily<br />

References<br />

Ayers, J. 1989. Recovery of oscillator function<br />

following spinal regeneration in the Sea<br />

Lamprey. In: Neuronal <strong>and</strong> Cellular Oscillators,<br />

ed. Jacklett, J., 371-406. New York: Marcel<br />

Dekker Inc.<br />

Ayers, J. 2004. Underwater walking. Arthropod<br />

Struct Dev. 33(3):347-60. doi: 10.1016/<br />

j.asd.2004.06.001.<br />

Ayers, J., & Rulkov, N. 2007. Controlling<br />

biomimetic underwater robots with electronic<br />

nervous systems. In: Bio-mechanisms of<br />

Animals in Swimming <strong>and</strong> Flying, eds.<br />

Kato, N., & Kamimura, S., 295-306.<br />

Tokyo: Springer-Verlag.<br />

Ayers, J., Rulkov, N., Knudsen, D., Kim,<br />

Y-B., Volkovskii, A., & Selverston, A. 2010.<br />

Controlling underwater robots with electronic<br />

nervous systems. Appl Bionics Biomech.<br />

7:57-67. doi: 10.1080/11762320903244843.<br />

Ayers, J., Wilbur, C., & Olcott, C. 2000.<br />

Lamprey robots. In: Proceedings of the<br />

International Symposium on Aqua<br />

July/August 2011 Volume 45 Number 4 151


Biomechanisms, eds. Wu, T., & Kato, N.,<br />

Tokai University.<br />

Ayers, J., & Witting, J. 2007. Biomimetic<br />

approaches to the control of underwater<br />

walking machines. Philos T R Soc Lond A.<br />

365:273-95. doi: 10.1098/rsta.2006.1910.<br />

Westphal, A., Rulkov, N., Ayers, J., Brady, D.,<br />

& Hunt, M. 2011. Controlling a lampreybased<br />

robot with an electronic nervous system.<br />

Smart Struct Syst. 7(6):471-84.<br />

Blustein, D., & Ayers, J. 2010. A conserved<br />

network for control of arthropod exteroceptive<br />

optical flow reflexes during locomotion. Lect<br />

Notes Artif Int. 6226:72-81.<br />

Bullock, T.H. 1968. Representation of information<br />

in neurons <strong>and</strong> sites for molecular<br />

participation. P Natl Acad Sci. 60:1058-68.<br />

doi: 10.1073/pnas.60.4.1058.<br />

Dickinson, P.S. 2006. Neuromodulation of<br />

central pattern generators in invertebrates <strong>and</strong><br />

vertebrates. Curr Opin Neurobiol. 16(6):<br />

604-14. doi: 10.1016/j.conb.2006.10.007.<br />

Kennedy, D., & Davis, W.J. 1977. Organization<br />

of invertebrate motor systems. In:<br />

H<strong>and</strong>book of Physiology, eds. Geiger, S.R.,<br />

K<strong>and</strong>el, E.R., Brookhart, J.M., & Mountcastle,<br />

V.B., Section 1, Vol. 1, Pt. 2:1023-87.<br />

Bethesda, MD: American Physiological<br />

<strong>Society</strong>. Section 1, Part 2.<br />

Kupfermann, I., & Weiss, K. 1978. The<br />

comm<strong>and</strong> neuron concept. Behav Brain<br />

Sciences. 1:3-39.<br />

Pearson, K.G. 1993. Common principles of<br />

motor control in vertebrates <strong>and</strong> invertebrates.<br />

Annu Rev Neurosci. 16:265-97. doi: 10.1146/<br />

annurev.ne.16.030193.001405.<br />

Rulkov, N.F. 2002. Modeling of spikingbursting<br />

neural behavior using two-dimensional<br />

map. Phys Rev E. 65:041922. doi: 10.1103/<br />

PhysRevE.65.041922.<br />

Selverston, A., & Ayers, J. 2006. Oscillations<br />

<strong>and</strong> oscillatory behavior in small neural circuits.<br />

Biol Cybern. 95:537-54. doi: 10.1007/<br />

s00422-006-0125-1.<br />

Stein, P.S.G. 1978. Motor systems, with<br />

specific reference to the control of locomotion.<br />

Annu Rev Neurosci. 1:61-81. doi: 10.1146/<br />

annurev.ne.01.030178.000425.<br />

152 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

A Hybrid Class Underwater Vehicle:<br />

Bioinspired Propulsion, Embedded<br />

System, <strong>and</strong> Acoustic Communication<br />

<strong>and</strong> Localization System<br />

AUTHORS<br />

Michael Krieg<br />

Peter Klein<br />

Robert Hodgkinson<br />

Department of Mechanical<br />

<strong>and</strong> Aerospace Engineering,<br />

University of Florida<br />

Kamran Mohseni 1<br />

Departments of Mechanical<br />

<strong>and</strong> Aerospace Engineering<br />

<strong>and</strong> Computer Engineering,<br />

Institute for Cyber Autonomous<br />

Systems, University of Florida<br />

Introduction<br />

Traditionally unmanned underwater<br />

vehicles fall into one of two categories.<br />

One class of vehicles (torpedo<br />

like) are built to travel long distances<br />

with minimal energy <strong>and</strong> are usually<br />

characterized by a long slender body,<br />

a rear propeller for propulsion <strong>and</strong> a<br />

set of fins to provide maneuvering<br />

forces. This type of vehicle is poorly<br />

suited for missions requiring a high degree<br />

of positioning accuracy because<br />

the control surfaces provide little to<br />

no maneuvering force at low forward<br />

velocity. The other class of vehicle<br />

similar to remotely operated vehicles<br />

(ROVs) is designed to operate in<br />

these situations, which do require<br />

1 This work was started while the group was at<br />

the University of Colorado.<br />

ABSTRACT<br />

Inspired by the natural locomotion of jellyfish <strong>and</strong> squid, a series of compact<br />

thrusters series is developed for propulsion <strong>and</strong> maneuvering of underwater vehicles.<br />

These thrusters successively ingest <strong>and</strong> expel jets of water in a controlled manner<br />

at high frequencies to generate propulsive forces. The parameters controlling<br />

the performance of the thrusters are reviewed <strong>and</strong> investigated to achieve higher<br />

thrust levels. The thrusters are compact <strong>and</strong> can be placed completely inside a vehicle<br />

hull providing the desired maneuvering capability without sacrificing a sleek<br />

hydrodynamic shape for efficient cruising. The system design of a prototype hybrid<br />

vehicle, called CephaloBot, utilizing these thrusters, is also presented. A compact<br />

<strong>and</strong> custom-developed embedded system is also designed for the CephaloBot. Key<br />

features of the system include a base set of navigational sensors, an acoustic<br />

system for localization <strong>and</strong> underwater communication, Xbee RF transceiver for<br />

communication above water, <strong>and</strong> a LabVIEW programmed processing board.<br />

Keywords: AUV, thruster, bioinspired, communication<br />

high positioning accuracy, <strong>and</strong> incorporate<br />

several thrusters at various locations<br />

to provide maneuvering forces<br />

in all directions. However, this class<br />

of vehicle typically has a very high<br />

drag coefficient due to the abundance<br />

of external thrusters <strong>and</strong> cannot travel<br />

to remote locations without additional<br />

support.<br />

The abundance of remote marine<br />

research sites requiring high positioning<br />

accuracy for inspection, as well as<br />

the desire to create fully autonomous<br />

vehicle sensor networks, has inspired<br />

significant research in a hybrid class<br />

of vehicles with the efficient cruising<br />

characteristics of the torpedo class<br />

<strong>and</strong> the maneuvering abilities of the<br />

ROV class. Some take a mechanical<br />

approach moving the maneuvering<br />

propellers into tunnels which run<br />

through the hull of the vehicle<br />

(Mclean, 1991; Torsiello, 1994) or<br />

into the fins themselves (Dunbabin<br />

et al., 2005). Others observe that nature’s<br />

swimmers have a healthy balance<br />

of long-distance endurance <strong>and</strong> highaccuracy<br />

low-speed maneuvering. Vehicles<br />

have been designed to use fins<br />

for both high-speed maneuvering as<br />

well as mimic the low-speed flapping<br />

of turtles <strong>and</strong> marine mammals (Licht<br />

et al., 2004; Licht, 2008; Kato,<br />

2011); <strong>and</strong> some use tail fins as a primary<br />

means of propulsion (Barrett<br />

et al., 1999). Our inspiration comes<br />

from the cyclical jet propulsion seen<br />

in jellyfish, scallops, octopus, squid<br />

<strong>and</strong> other cephalopod. Squid jet propulsion<br />

produces the fastest swimming<br />

July/August 2011 Volume 45 Number 4 153


velocities seen in aquatic invertebrates<br />

(O’Dor & Webber, 1991; Anderson<br />

& Grosenbaugh, 2005).<br />

Jetting locomotion begins when<br />

the squid inhales seawater through a<br />

pair of ostia behind the head, filling<br />

the mantle cavity (see Figure 1). The<br />

mantle then contracts forcing seawater<br />

out through the funnel that rolls into a<br />

high-momentum vortex ring <strong>and</strong> imparts<br />

the necessary propulsive force<br />

(Anderson & Grosenbaugh, 2005).<br />

The versatility of the system permits<br />

two distinct gaits, cruising <strong>and</strong> escape<br />

jetting (Bartol et al., 2008). During<br />

cruising, squid swim at nominal<br />

speed with a greater efficiency than escape<br />

jetting, which involves a hyperinflation<br />

of the mantle followed by a<br />

fast powerful contraction to impart<br />

significant acceleration at the cost of<br />

both muscular <strong>and</strong> fluid dynamic<br />

losses. Bartol et al. (2009) report cruising<br />

mode efficiency at 69% (±14%)<br />

averaged over several species <strong>and</strong><br />

swimming speeds <strong>and</strong> 59% (±14%)<br />

for escape jetting. Additionally, propulsive<br />

efficiency was seen to rise as<br />

high as 78% in adult L. brevis swimming<br />

at high velocities <strong>and</strong> averaged<br />

87% (±6.5%) for paralarvae (Bartol<br />

et al., 2008), challenging the notion<br />

that a low-volume high-velocity jet<br />

inherently negates a high propulsive<br />

efficiency.<br />

The locomotion of jellyfish tends<br />

to be very similar to that of squid<br />

with some key differences, primarily<br />

that the refilling phase of jellyfish<br />

FIGURE 1<br />

Diagram of squid layout <strong>and</strong> locomotion.<br />

swimming uses the same bell opening<br />

as the jetting phase. Despite the fact<br />

that squid do not use the funnel during<br />

refilling, the inlet vents are still<br />

on the anterior side of the mantle cavity,<br />

meaning that locomotion for both<br />

organisms is quite different from traditional<br />

pumping mechanisms. Jellyfish<br />

use the cyclic jetting process for<br />

feeding as well as locomotion as is<br />

evidenced by Lagrangian coherent<br />

structures <strong>and</strong> particle tracer analysis<br />

(Lipinski & Mohseni, 2009; Wilson<br />

et al., 2009); in addition, both squid<br />

<strong>and</strong> jellyfish utilize jetting for respiration,<br />

taking advantage of the large<br />

fluid flow rates. Both of these factors<br />

can make it difficult to determine<br />

which swimming behaviors are optimized<br />

for propulsion versus secondary<br />

functions. Similar to the different gaits<br />

seen in squid locomotion, different<br />

species of jellyfish generally fall into<br />

two categories of swimmers based on<br />

the ‘quality’ of vortex ring they produce.<br />

Jellyfish like moon jellyfish<br />

have a very large bell opening, <strong>and</strong><br />

the jetting motion is similar to a paddling<br />

type motion. Box jellyfish <strong>and</strong><br />

otherfasterswimmingjellyfish have<br />

smaller bell openings with nozzle-like<br />

flaps<strong>and</strong>haveamuchmoredistinct<br />

jet. Jellyfish morphology during swimming<br />

has been digitally captured from<br />

experiment, <strong>and</strong> the body motions<br />

were imported into numerical simulations<br />

to predict body forces on the<br />

swimming jellyfish, determining drastically<br />

different swimming efficiencies.<br />

Froude propulsive efficiency of jellyfish<br />

was directly calculated by Sahin<br />

<strong>and</strong> Mohseni (2008, 2009; Sahin et al.,<br />

2009) to be 37% for Aeqorea victoria<br />

<strong>and</strong> 17% for Sarsia tubulosa. It should<br />

be noted that both species of jellyfish<br />

most likely do not use vortex<br />

generation for the sole purpose of<br />

locomotion. Aeqorea victoria uses vortex<br />

generation for feeding <strong>and</strong> Sarsia<br />

tubulosa as an escape mechanism. Empirical<br />

data gathered through digital<br />

particle image velocimetry (DPIV)<br />

measurements of several species shows<br />

similar efficiency characteristics for<br />

the different swimming patterns<br />

(Dabiri et al., 2010).<br />

The general concept of propelling<br />

water craft by ejecting a high-velocity<br />

water jet is centuries old, was hypothesized<br />

by both Bernoulli <strong>and</strong> Benjamin<br />

Franklin, <strong>and</strong> was utilized in a rudimentary<br />

sense in one of the first steam<br />

boat designs by James Rumsey (Allen,<br />

2010). Continuously pumped jets are<br />

usedforpropulsioninmodernwater<br />

craft-like jet skis <strong>and</strong> bow thrusters<br />

of motorboats; however, this type of<br />

jet propulsion is inherently different<br />

from the propulsion of squid <strong>and</strong> jellyfish,<br />

which create distinct vortex rings.<br />

The thrusters of this paper also produce<br />

finite jets, which form arrays of<br />

vortex rings, <strong>and</strong> should be considered<br />

fundamentally different than continuous<br />

jet thrusters.<br />

This paper showcases a complete<br />

hybrid class vehicle that demonstrates<br />

added maneuvering capabilities utilizing<br />

a set of bio-inspired jet thrusters.<br />

The manuscript will focus on three<br />

primary systems of the vehicle: a bioinspired<br />

thruster system <strong>and</strong> fundamentals<br />

of thruster mechanics, an<br />

acoustic system, which serves the<br />

dual purpose of communication <strong>and</strong><br />

localization, <strong>and</strong> a compact embedded<br />

control system. The manuscript is<br />

154 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


organized as follows. The mechanics<br />

ofthethrusteraswellasthethrust<br />

dynamics are described in ‘Vortex<br />

Ring Thrusters’ section. A brief history<br />

of thruster <strong>and</strong> vehicle prototypes is<br />

given in ‘Thruster <strong>and</strong> CephaloBot<br />

Evolution’. The‘Hybrid Vehicle Description’<br />

section gives basic requirements<br />

for a hybrid class vehicle, <strong>and</strong> the<br />

subsystem components of CephaloBot<br />

are described in more details in ‘Communication<br />

Localization System’,<br />

‘Embedded System’, <strong>and</strong>‘Sensors’<br />

sections.<br />

Vortex Ring Thrusters<br />

Ourthrusterinspiredbyjellyfish<br />

<strong>and</strong> squid propulsion consists of an internal<br />

fluid cavity, with a semi-flexible<br />

plunger used to drive fluid motion,<br />

<strong>and</strong> a small circular orifice exposed to<br />

the external fluid. See Figure 2 for a<br />

diagram of the thruster layout. The<br />

cavity of the thruster provides the<br />

same functionality as the squid mantle<br />

or the jellyfish bell, exp<strong>and</strong>ing <strong>and</strong><br />

contracting to cycle water in <strong>and</strong> out<br />

of the circular orifice (functionally<br />

similar to the squid funnel/siphon).<br />

Since the thruster generates propulsion<br />

by creating energetic vortex rings, it<br />

FIGURE 2<br />

Conceptual diagram of the thruster key components.<br />

Jet shown as hypothetical slug of<br />

fluid.<br />

is termed the ‘vortex ring thruster’<br />

(VRT).<br />

Since VRTs are contained internal<br />

to the vehicle (with only a small opening<br />

on the vehicle surface), they do not<br />

significantly affect the vehicle’s forward<br />

drag profile, which means that a<br />

vehicle equipped with a set of VRTs<br />

for low-speed maneuvering <strong>and</strong> a rear<br />

propeller for primary propulsion will<br />

have a sleek aerodynamic shape allowing<br />

fast efficient cruising to a site of<br />

interest but still maintain full maneuverability<br />

(even at zero forward speed)<br />

upon reaching that site of interest. See<br />

Krieg <strong>and</strong> Mohseni (2010) for ‘parallel<br />

parking’ capability of an earlier version<br />

of our vehicle. Additionally, since<br />

the VRT only needs a single opening<br />

(unlike tunnel thrusters or traditional<br />

pumps, which extend from one end<br />

of the hull to the other), it allows for<br />

a greater degree of freedom for internal<br />

system arrangement.<br />

The impulse generated by this type<br />

of device can be modeled as if the jet<br />

acts like a solid slug of fluid with a uniform<br />

velocity across the nozzle opening<br />

(Mohseni, 2004, 2006; Krieg &<br />

Mohseni, 2008). Properties of fluid<br />

slug have been investigated by several<br />

groups (Glezer, 1988; Gharib et al.,<br />

1998; Shariff & Leonard, 1992;<br />

Mohseni & Gharib, 1998; Krieg &<br />

Mohseni, 2008). The total impulse<br />

generated for a single pulsation under<br />

slug assumptions is<br />

I slug ðÞ¼ρπ=4∫ t<br />

t 0 u2 ðτÞD 2 dτ;<br />

ð1Þ<br />

Krueger <strong>and</strong> Gharib (2003) showed<br />

that the impulse created by a cylinder<br />

piston type vortex generator was consistently<br />

higher than the impulse predicted<br />

by the slug model. This added<br />

impulse was attributed to a pressure<br />

gradient at the nozzle exit plane,<br />

referred to as ‘nozzle overpressure.’<br />

Adding I p (the impulse due to overpressure),<br />

we get an equation for the<br />

impulse, in terms of the nozzle pressure,<br />

p (which is a function of time<br />

<strong>and</strong> radial position), <strong>and</strong> the stagnation<br />

pressure, p ∞ .<br />

IðÞ¼I t slug ðÞþI t p ðÞ t<br />

I p t<br />

ðÞ¼∫ t 0∫ A pr; τ<br />

½ ð Þ p ∞ ŠdAdτ ð2Þ<br />

We performed initial testing on a<br />

thruster which periodically ingested<br />

<strong>and</strong> expelled jets with a sinusoidal velocity<br />

program; which was chosen for<br />

simplicity of fabrication. Assuming<br />

that there is no net momentum transfer<br />

during the ingestion phase (fluid<br />

being taken into the cavity starts at<br />

rest outside of the thruster <strong>and</strong> ends<br />

at rest inside the cavity) <strong>and</strong> ignoring<br />

the pressure impulse (which will be<br />

accounted for with a coefficient term<br />

post analysis), then the average thrust<br />

produced over a full cycle can be calculated<br />

in terms of the thruster frequency<br />

to be (Krieg & Mohseni, 2008)<br />

<br />

‐<br />

T ¼ ρπ3 L 2<br />

16 D4 f 2 ð3Þ<br />

D<br />

where u is the piston velocity (mass<br />

flux across thruster opening divided<br />

by nozzle area), D is the nozzle Diameter,<br />

ρ is the fluid density, t is the time<br />

at which the impulse is evaluated, <strong>and</strong><br />

τ is a dummy variable for time initialized<br />

at the beginning of pulsation.<br />

Here f is the frequency of actuation,<br />

<strong>and</strong> the term L=D is the jet stroke<br />

ratio. If the jet maintained its shape as<br />

a solid cylinder the stroke ratio would<br />

be the ratio of length to diameter of<br />

that cylinder (see Figure 2). The stroke<br />

ratio has also been called the formation<br />

July/August 2011 Volume 45 Number 4 155


FIGURE 3<br />

Thruster static testing environment.<br />

time since it is equivalent to the time<br />

since initiation of the jet flow scaled<br />

by jet velocity <strong>and</strong> nozzle diameter,<br />

t* ¼ L=D ¼ ∫ t 0 u ð τ Þdt=D. Theformation<br />

time is closely related to the<br />

jet formation dynamics. When a jet is<br />

expelled into a stationary fluid, the viscous<br />

forces cause the initial portion of<br />

the jet to roll into a tightly wound vortex<br />

ring. As more fluid is expelled, it<br />

feeds the growing vortex ring until<br />

a critical saturation point is reached<br />

<strong>and</strong> the vortex ring can no longer support<br />

the added circulation. At this<br />

point, the vortex ring separates from<br />

the remaining shear flow. The formation<br />

time when the jet has achieved the<br />

same circulation as the final vortex ring<br />

is known as the formation number.<br />

Gharib et al. (1998) demonstrated<br />

that impulsively started vortex rings<br />

have a universal formation number<br />

(≈3.6-4.2) independent of jet velocity<br />

<strong>and</strong> diameter. However, numerical<br />

studies have shown the formation<br />

number to be drastically lower for<br />

jets created with a parabolic velocity<br />

profile (Rosenfeld et al., 1998) or a<br />

2-D jet velocity like those produced<br />

in conical nozzles (Rosenfeld et al.,<br />

2009).<br />

The slug model predicts that the<br />

average thrust is proportional to both<br />

the square of the actuation frequency<br />

<strong>and</strong> the square of the stroke ratio/<br />

formation time. To test this assertion,<br />

the thruster was placed in a static fluid<br />

reservoir <strong>and</strong> suspended from a load<br />

cell, <strong>and</strong> the thrust output was measured<br />

directly. This testing setup is<br />

shown in Figure 3 (Krieg & Mohseni,<br />

2008). It should be noted that if the<br />

vehicle is moving during thruster<br />

pulsation there will inherently be a<br />

non-zero momentum transfer during<br />

ingestion; however, as was previously<br />

mentioned these thrusters are primarily<br />

intended for low-speed maneuvering<br />

(vehicle velocities well below<br />

jetting velocities), so that the momentum<br />

transfer during ingestion will still<br />

be negligible compared to the momentum<br />

transfer of the jetting phase. In<br />

addition, the expelled jet rolls into an<br />

isolated vortex ring which inherently<br />

produces a large momentum transfer<br />

associated with the fluid impulse<br />

of the vortex ring itself. During ingestion,<br />

the small internal cavity severely<br />

limits vortex ring formation as well as<br />

momentum transfer associated with it.<br />

The thruster was tested over a wide<br />

range of stroke ratios <strong>and</strong> actuation frequencies.<br />

As can be seen in Figure 4,<br />

the thrust shows a square proportionality<br />

to frequency, within a certain frequency<br />

range. When producing jets<br />

with low stroke ratio, this range this<br />

is the entire sub-cavitation frequency<br />

range. However, when the jet stroke<br />

ratio goes above the formation number,<br />

the thruster exhibits a parabolic<br />

dependence on the actuation frequency<br />

after a short range of square dependence.<br />

To more clearly show this<br />

trend, we define a scale factor, which<br />

FIGURE 4<br />

Thrust versus frequency. Each set of markers<br />

shows thrust data for a different stroke ratio.<br />

The error bars plotted along with the thrust<br />

relationship represent a single st<strong>and</strong>ard deviation<br />

of the thrust data.<br />

is a measure of the accuracy of the<br />

slug model for various operating conditions<br />

α = T Exp / ‐ T (Krieg & Mohseni,<br />

2008) where ‐ T is the slug model predicted<br />

average thrust from equation (3).<br />

This scale factor is plotted with respect<br />

to frequency for stroke ratios below<br />

the formation number in Figure 5a<br />

<strong>and</strong> for stroke ratios above the formation<br />

number in Figure 5b. Note that<br />

the thruster of this study has a nozzle<br />

which is essentially a flat plate with a<br />

circular orifice in the middle. This<br />

type of nozzle produces a 2-D jet<br />

flow similar to a conical nozzle, meaning<br />

that the jet formation number is<br />

closer to 3 (Rosenfeld et al., 2009).<br />

For a more in depth analysis of the<br />

shift in formation number due to the<br />

2-D aspect of the jet, see Krieg <strong>and</strong><br />

Mohseni (2011).<br />

First, consider the thrust response<br />

of the actuator operating below the<br />

formation number (Figure 5a). In the<br />

low-frequency regime, the scale factor<br />

is higher than that predicted by the<br />

slug model due to the nozzle overpressure,<br />

reaching 1.4 times the predicted<br />

value. Krueger <strong>and</strong> Gharib (2003)<br />

showed that the pressure impulse can<br />

reach as much as 40% of the total<br />

156 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 5<br />

Scale factor (slug model accuracy) versus frequency for stroke ratios below (a) <strong>and</strong> above<br />

(b) the formation number. Error bars shown on data points indicate a st<strong>and</strong>ard deviation of<br />

measured values at that frequency (also taken in the scaled space).<br />

FIGURE 6<br />

Successive frames of jet flow showing the<br />

thruster re-ingesting wake flow.<br />

impulse for low stroke ratio jets corresponding<br />

to a total impulse 1.6 times<br />

the predicted slug model impulse.<br />

However, as the frequency increases,<br />

the total thrust settles on the value predicted<br />

by the slug model, meaning that<br />

in this frequency range the impulse<br />

due to overpressure during expulsion<br />

is equal to the impulse due to “underpressure”<br />

during refilling so that the<br />

net impulse transfer is that predicted<br />

by the slug model. Now consider<br />

the thruster response when operating<br />

above the formation number, shown<br />

in Figure 5b. Again the low-frequency<br />

ranges exhibit an added impulse due<br />

to the nozzle overpressure; albeit to a<br />

lower extent as observed by Krueger <strong>and</strong><br />

Gharib (2003). But the high-frequency<br />

range exhibits an added loss in thrust<br />

with respect to the slug model prediction.<br />

This relative loss is seen to<br />

increase monotonically with both actuation<br />

frequency <strong>and</strong> stroke ratio.<br />

This suggests that another assumption<br />

made in the slug model is no longer<br />

valid when operating above the formation<br />

number. We assume that this loss<br />

in model accuracy is tied into the assumption<br />

made that all fluid being<br />

ingested between pulsations is at rest<br />

outside of the thruster. When a jet is<br />

ejected with a stroke ratio above the<br />

formation number, some of the shear<br />

flow is left behind in the trailing<br />

wake of the leading vortex ring. The<br />

trailing wake has a lower momentum<br />

than the leading vortex ring <strong>and</strong> travels<br />

at a much lower induced velocity but<br />

still has a forward momentum substantially<br />

larger than the surrounding resting<br />

fluid. Therefore, the loss in slug<br />

model accuracy could be explained<br />

by the thruster ingesting some of<br />

the trailing wake during the refilling<br />

phase. Figure 6 shows successive<br />

frames from a video of the thruster’s<br />

forming jet (at a high stroke ratio)<br />

where some of the trailing wake is ingested<br />

back into the thruster.<br />

It should also be noted that the<br />

scale factor results are only presented<br />

above an actuation frequency of 4 Hz,<br />

because the thruster was designed to<br />

operate cyclically rather than generate<br />

individual pulsations. The 2-D nature<br />

of the jet created by this type of<br />

thruster (orifice nozzle) has an added<br />

effectonthenozzleoverpressurenot<br />

seen in the frequency ranges presented.<br />

This effect is fully explained (along<br />

with a more in depth description of<br />

impulse generation) for a single pulsation<br />

with constant jet velocity in Krieg<br />

<strong>and</strong> Mohseni (2011). Despite the<br />

relative magnitude of the overpressure<br />

impulse (with respect to the momentum<br />

impulse, I slug ), it is only observed<br />

in low actuation frequencies, which are<br />

coupled with low thrust output. Often<br />

vehicle mission scenarios will necessitate<br />

a large magnitude thrust output<br />

(higher actuation frequencies), where<br />

the overpressure is cancelled out, <strong>and</strong><br />

the slug model provides an accurate<br />

thrust measurement.<br />

The vortex ring formation phenomenon<br />

plays a key role in the jet<br />

locomotion process in squid as well.<br />

Bartol et al. (2009) observed that squid<br />

have two distinct swimming gaits. In<br />

the efficient cruising gate jets are expelled<br />

below the formation number,<br />

so that the majority of the jet rolls into<br />

the primary vortex ring. Alternatively,<br />

in threatening situations the squid employs<br />

a swimming technique referred<br />

to as escape jetting; which begins with<br />

the hyperinflation of the mantle followed<br />

by a fast contraction expelling a<br />

jet well above the formation number.<br />

July/August 2011 Volume 45 Number 4 157


Presumably this behavior indicates that<br />

this type of jet propulsion is most efficient<br />

when expelling jets below the formation<br />

number <strong>and</strong> that jetting above<br />

the formation number can achieve higher<br />

thrust at the expense of fluid losses.<br />

Therefore, all subsequent vehicle<br />

thrusters have been designed with a<br />

set diameter resulting in a stroke ratio<br />

near the formation number; to achieve<br />

a maximum level of thrust, while still<br />

being accurately described by the slug<br />

model. It should be noted that the formation<br />

number for a jet created on a<br />

moving vehicle will not be the same<br />

as the formation number of the jet in<br />

the static setup. Krueger et al. (2006)<br />

showed that vortex rings formed in<br />

the presence of a uniform background<br />

co-flow have a lower formation number<br />

than vortex rings formed in a resting<br />

fluid. Since a moving vehicle will<br />

inherently induce a co-flow with the<br />

jetting direction, this effect must be<br />

taken into consideration. However,<br />

the reduction in formation number<br />

is proportional to the ratio between<br />

co-flow velocity <strong>and</strong> jet velocity; therefore,<br />

the low-frequency (low jet velocity)<br />

pulsation will experience pinch off<br />

at an earlier formation time, but the<br />

lower-frequency pulsation is also less<br />

effected by the dynamics of cyclic vortex<br />

ring formation.<br />

One advantage of the VRT is that<br />

it produces a desired level of thrust<br />

almost instantaneously (Krieg &<br />

Mohseni, 2010). Propeller style thrusters<br />

suffer from a rise time associated<br />

with reaching the static thrust level<br />

after initiating rotation. This rise time<br />

is inversely proportional to the desired<br />

level of thrust <strong>and</strong> can be on the order<br />

of several seconds for low thrust levels<br />

(Fossen, 1991; Yoerger et al., 1990).<br />

VRTs also have a rise time associated<br />

with reaching the desired level of thrust,<br />

which is inversely proportional to the<br />

level of thrust. However, this rise time<br />

is an order of magnitude smaller for<br />

VRTs. The exact thrust program as a<br />

function of time is sinusoidal, due to<br />

the nature of the thruster; however,<br />

using several thrust data sets to average<br />

out the dynamic component <strong>and</strong> fitting<br />

the average thrust to a basic logarithmic<br />

curve the mean rise time can be observed.<br />

The fitted curves for several operational<br />

frequencies, <strong>and</strong> a stroke ratio<br />

of 4.3 is shown in Figure 7.<br />

FIGURE 7<br />

Mean thrust produced at various frequencies<br />

(static desired thrust level) versus time. Rise<br />

time inversely proportional to thrust level.<br />

Along with a minimal rise time, the<br />

VRT is also immune to a thrust lag seen<br />

in tunnel thrusters (Mclean, 1991),<br />

where thrust continues to be exerted<br />

on the vehicle after the thruster has<br />

been terminated.<br />

Thruster <strong>and</strong> CephaloBot<br />

Evolution<br />

The compact thrusters used in vehicle<br />

testbeds to provide validation of<br />

static testing have taken a wide variety<br />

of forms. The first generation utilized<br />

a solenoid driving mechanism <strong>and</strong> a<br />

flexing diaphragm to expel fluid (Figure<br />

8a). The solenoid driving mechanism<br />

suffered from reduced stroke<br />

at higher actuation frequencies as the<br />

solenoid stroke was load dependent.<br />

All of the subsequent iterations have<br />

used mechanical driving mechanisms<br />

for better flexibility in adjusting the operation<br />

conditions. Gen. 2 (Figure 8b)<br />

used a completely flexible cavity in a<br />

fitted mold, whereas Gen. 3 <strong>and</strong> 4<br />

switched to a semi-flexible cavity<br />

which was reinforced to ensure constant<br />

diameter but allow compression<br />

in height. Gen. 3 (Figure 8c) used a<br />

complicated encompassing cylindrical<br />

tube cam to drive cavity compression,<br />

but mechanical complexities <strong>and</strong> reliability<br />

issues caused us to simplify to<br />

a basic crank shaft design for Gen. 4<br />

(Figure 8d). In the figure, this thruster<br />

is shown with an acrylic casing to allow<br />

the components to be seen. The vehicle<br />

ready thruster uses an aluminum<br />

casing (Clark et al., 2009).<br />

Along with the thrusters themselves<br />

the vehicle testbeds housing<br />

the thrusters have evolved rapidly. Figure<br />

9 shows the evolution of vehicle<br />

testbeds used to demonstrate the feasibility<br />

of maneuvering using VRTs.<br />

Starting with the oldest vehicle at the<br />

bottom <strong>and</strong> successive generations<br />

upwards, the first vehicle only had<br />

fins to provide a st<strong>and</strong>ard for maneuvering<br />

capabilities, 2nd, 3rd <strong>and</strong> 4th<br />

generation vehicles contain 1st, 2nd<br />

<strong>and</strong> 3rd generation thrusters, respectively,<br />

<strong>and</strong> increasing levels of autonomy.<br />

The lessons learned from these<br />

vehicles directly lead to the development<br />

of the most recent hybrid class<br />

vehicle described in the following<br />

section.<br />

Hybrid Vehicle Description<br />

The newest generation of vehicle<br />

is intended to be used in autonomous<br />

sensor network applications. Therefore,<br />

the vehicle must be able to travel<br />

on long range missions collecting data<br />

158 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 8<br />

Successive generations of thrusters (a) utilize solenoid driver (7.5 cm/3-inch diameter), (b) utilize cavity <strong>and</strong> mold (10 cm/4-inch casing diameter),<br />

(c) have encompassing driving mechanism (12.5 cm/5-inch plate diameter) <strong>and</strong> (d) utilize simple crank shaft <strong>and</strong> semi-flexible ducting (10 cm/4-inch<br />

plate diameter).<br />

butmustalsobecapableofautonomously<br />

docking with permanent<br />

support structures. These structures<br />

will be responsible for downloading<br />

the vehicle’s mission data, changing<br />

mission objectives <strong>and</strong> recharging vehicle<br />

batteries.<br />

Autonomous docking with such a<br />

structure is a complicated problem,<br />

requiring not only high-accuracy maneuvering,<br />

but equally high-accuracy<br />

localization <strong>and</strong> attitude determination.<br />

In addition vehicles in this type of environment<br />

need to be capable of communicating<br />

with each other at short<br />

FIGURE 9<br />

The first four generations of our vehicle test<br />

beds. Oldest vehicle at bottom, <strong>and</strong> successive<br />

vehicles placed in ascending order.<br />

FIGURE 10<br />

Fifth generation hybrid vehicle CephaloBot.<br />

distances for coordinated missions. In<br />

general reduction in cost <strong>and</strong> internal<br />

structure are also desirable to allow<br />

for more vehicles in the network with<br />

a wider range of payload options.<br />

The base vehicle (Figure 10) is separated<br />

into three hull sections. The<br />

front <strong>and</strong> back sections house all of<br />

the actuators. Each section has two<br />

VRTs <strong>and</strong> an active buoyancy control<br />

device (BCD). Each thruster has an<br />

overall diameter of 7.6 cm (3 inches),<br />

a nozzle diameter of 1.8 cm (0.6 inch),<br />

has a stroke ratio of 4.5, <strong>and</strong> produces<br />

close to 2N of thrust at 30 Hz before<br />

cavitation starts to occur in the cavity.<br />

The back section also has the rear propeller<br />

motor. The primary batteries<br />

<strong>and</strong> all the electronics except motor<br />

control are housed in the center section.<br />

An additional payload section<br />

may be added between front <strong>and</strong> center<br />

to include additional sensors,<br />

devices, <strong>and</strong>/or batteries. Regulated<br />

power <strong>and</strong> digital communication<br />

lines interface the payload to the center<br />

section. When put together, the<br />

0.15 m (6 inches) diameter, 0.92 m<br />

(36 inches) long vehicle weights a neutrally<br />

buoyant 16 kg (36 lb). The<br />

BCDs can change the buoyancy by<br />

±1% to dive or surface, <strong>and</strong> 2 kg of<br />

internal ballast are adjustable to balance<br />

the pitch of the vehicle.<br />

Communication/<br />

Localization System<br />

Due to the physical properties of<br />

water, RF communication methods<br />

July/August 2011 Volume 45 Number 4 159


are not practical for small unmanned<br />

underwater vehicles. CephaloBot has<br />

a joint acoustic communication <strong>and</strong><br />

localization system which is ideally<br />

suited for underwater sensor network<br />

applications. The communication<br />

technique is based on binary frequency<br />

modulation whereas the localization<br />

methodology is based on a time delay<br />

of arrival technique. The system consists<br />

of a specialized hydrophone array<br />

(fabricated in house as described later<br />

in this section), which interpret acoustic<br />

signals for both information <strong>and</strong><br />

directional content. The receiving hydrophone<br />

array consists of three piezo<br />

electric ceramics spaced in a triangular<br />

arrangement parallel to the vehicle<br />

principle axis plane (Figure 11). Two<br />

piezo electric ceramics are placed on a<br />

line parallel to the pitching axis, <strong>and</strong><br />

the 3rd is extended along the roll axis<br />

from the midpoint of the other two.<br />

The entire array is encased in urethane<br />

rubber with an acoustic impedance<br />

similar to water. The transmitting<br />

node consists of a single high-power<br />

piezo electric ceramic encased in the<br />

same type of urethane rubber as the<br />

FIGURE 11<br />

Definition of principle submarine axes.<br />

hydrophone array. Each vehicle is<br />

equippedwithbothareceiving<strong>and</strong><br />

transmitting node on the underside<br />

of the vehicle (see Figure 12).<br />

FIGURE 12<br />

Transducer nodes on the vehicle. Receiving<br />

node on the right <strong>and</strong> transmitting node on<br />

the left.<br />

All three hydrophones in the receiving<br />

node receive a signal from a<br />

transmitting node (either on another<br />

vehicle or a docking station). The<br />

phase lag (ϕ) between the signal coming<br />

from the hydrophone on the roll<br />

axis <strong>and</strong> the signals coming from the<br />

two hydrophones on the pitching axis<br />

is measured <strong>and</strong> correlated with the<br />

signal frequency ( f ) to determine the<br />

time between when the hydrophones<br />

received the source signal Δt i = ϕ i /f.<br />

The time lag is then multiplied by<br />

the speed of sound in water to get<br />

relative source distances in the vehicle<br />

frame <strong>and</strong> transformed into the inertial<br />

frame to get the azimuth <strong>and</strong> elevation<br />

of the receiving node with respect to<br />

the source node.<br />

The hybrid localization/communication<br />

system is comprised of three<br />

main phases: data sending, data receiving<br />

<strong>and</strong> localization. Data receiving<br />

<strong>and</strong> localization both use the<br />

same incoming acoustic wave, the<br />

first 1000 cycles are dedicated to localization<br />

<strong>and</strong> the remaining cycles are<br />

frequency modulated. The acoustic<br />

wave propagates from a sending transducer<br />

located at some other node <strong>and</strong><br />

propagates towards the vehicle. Once<br />

this wave is received by the hydrophone<br />

array, it is conditioned by filtering<br />

<strong>and</strong> amplification circuitry.<br />

The conditioning circuitry contains a<br />

17-dB gain pre-amplifier followed by<br />

two stages of filtering <strong>and</strong> two stages<br />

of amplification to bring the signal<br />

voltage to 5 V. The electrical voltage<br />

is then passed through exclusive or<br />

gates with one of the other signals (in<br />

the case of hydrophone 0, the signal is<br />

also passed to demodulation circuitry).<br />

During the localization portion of receiving<br />

an onboard microcontroller<br />

reads <strong>and</strong> stores the output of each of<br />

the three exclusive-or gates <strong>and</strong> also<br />

determines which of the three signals<br />

arrives first. The duty cycle of the<br />

exclusive-or gates directly relates to<br />

the phase difference of the signals. The<br />

localization hardware calculates <strong>and</strong><br />

passes the azimuth <strong>and</strong> elevation angles<br />

(which defineaconeofpossiblevehicle<br />

locations with the source node at the<br />

vertex) to the main navigation processor<br />

which uses secondary positioning sensors<br />

(depth sensor <strong>and</strong> electronic compass)<br />

to determine a unique solution to<br />

the position of the receiving node with<br />

respect to the transmitting node.<br />

After the localization portion of<br />

receiving is complete the output of<br />

the exclusive-or gates are ignored <strong>and</strong><br />

the microcontroller reads <strong>and</strong> stores the<br />

DC voltage level from a frequency to<br />

voltage converter which directly represents<br />

the frequency of the incoming<br />

frequency modulated signal.<br />

In order for the localization methodology<br />

to function properly the hydrophones<br />

must be placed no more<br />

than λ/2 apart where the wavelength<br />

λ is equal to the speed of propagation<br />

divided by the frequency (λ =c/f ).<br />

This leads to a maximum spacing of<br />

3 cm for a frequency of 25 kHz <strong>and</strong><br />

1.875 cm for a frequency of 40 kHz.<br />

160 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Miniature hydrophones which are approximately<br />

1 cm in diameter are commercially<br />

available from Reson but are<br />

on the order of $1000 per unit. Halfinch<br />

diameter cylindrical piezo electric<br />

ceramics (SMC14H12111) are available<br />

from Steminc for less than $10/<br />

each. Placement of the three halfinch<br />

diameter piezo electric ceramics<br />

in a triangle pattern yields a maximum<br />

navigational frequency of 25 kHz.<br />

The sending <strong>and</strong> receiving transducers<br />

were made in house using a method<br />

similar to that in Li et al. (2010).<br />

The sending piezoceramic was chosen<br />

primarily based on its resonant frequency<br />

of 22 kHz <strong>and</strong> cylindrical<br />

shape to provide an omni-directional<br />

signal. The receiving hydrophone<br />

array <strong>and</strong> transmitter are shown placed<br />

on the belly of the CephaloBot in<br />

Figure 12.<br />

Overall this customized communication<br />

localization system places a<br />

minimal load on the vehicle. The circuitry<br />

has a very small footprint (Figure<br />

13) compared to typical commercial<br />

acoustic modems. In a two-way communication<br />

mode, the power draw is<br />

less than 1 W, <strong>and</strong> in simple listening<br />

mode, the power draw is less than<br />

0.25 W. The transducers are fabricated<br />

in house so they can be customized to<br />

FIGURE 13<br />

Communication <strong>and</strong> localization hardware.<br />

any shape <strong>and</strong> placed at any location<br />

on the vehicle to improve vehicle drag<br />

characteristics. The entire system is<br />

fabricated for under $300 in materials,<br />

making it a suitable option for sensor<br />

networks requiring several low<br />

cost vehicles.<br />

FIGURE 14<br />

Embedded System<br />

The embedded system was custom<br />

designed for the CephaloBot. The vehicle<br />

will be used by researchers for<br />

various underwater sensor networking<br />

applications <strong>and</strong> multi-vehicle<br />

coordination. In order to enable this<br />

underwater network to interact with<br />

a potential aerial sensor network,<br />

CephaloBot is also equipped with RF<br />

communication capabilities, spare<br />

computation power, <strong>and</strong> the ability<br />

to quickly add new sensors. Each vehicle<br />

must be robust <strong>and</strong> easy to h<strong>and</strong>le<br />

<strong>and</strong> operate. The embedded system is<br />

separated into multiple printed circuit<br />

boards (PCB). A power distribution<br />

board h<strong>and</strong>les voltage regulation <strong>and</strong><br />

battery charging. An interface board<br />

connects the onboard devices <strong>and</strong> sensors<br />

with the power board <strong>and</strong> processing<br />

device. Smaller PCBs are located<br />

throughout the vehicle to provide<br />

specific functionality such as motor<br />

control, user interface, or simply wire<br />

routing. Figure 14 shows the electronics<br />

located in the center section, where<br />

everything except the motor controllers<br />

<strong>and</strong> user interface is located.<br />

Processing<br />

The primary processing device on<br />

the vehicle is a National Instruments<br />

Single-board RIO (sbRIO). It has an<br />

on-board 400 MHz processor running<br />

real-time LabVIEW software, 128 MB<br />

of RAM, 256 MB of flash, <strong>and</strong> a<br />

40-MHz 2 M gate FPGA (field programmable<br />

gate array), also programmed<br />

in LabVIEW, providing<br />

110 digital I/O pins. The combination<br />

of microprocessor <strong>and</strong> FPGA was<br />

proven effective in the 4th generation<br />

vehicle where a Compact RIO was<br />

used. All of the low-level communication,<br />

interface, <strong>and</strong> control tasks are<br />

h<strong>and</strong>led on the FPGA, leaving the<br />

microprocessor open to perform high<br />

level mission control. The core vehicle<br />

software is located on the FPGA. The<br />

primary benefit is that the mission critical<br />

software will always run at full<br />

speed <strong>and</strong> safety checks will be implemented<br />

that a vehicle user cannot easily<br />

override. In this way, even if the<br />

algorithms being tested fail, the vehicle<br />

will remain safe to itself <strong>and</strong> surroundings.<br />

The 400-MHz real-time processor<br />

is little used by the core system <strong>and</strong><br />

therefore provides significant computational<br />

power to the researcher. An<br />

Hybrid vehicle embedded system mounted on battery pack.<br />

July/August 2011 Volume 45 Number 4 161


additional 2 GB of flash storage is<br />

added with a serial data logger <strong>and</strong><br />

can be used for mission review.<br />

Batteries<br />

The power source is a four-cell lithium<br />

polymer pack. The nominal voltage<br />

of 14.8 V ranges from 10 to 16.8 V,<br />

depending on charge level. The lithium<br />

polymer chemistry was selected because<br />

of its high-power density <strong>and</strong><br />

excellent discharge characteristics. A<br />

single four-cell battery pack was chosen<br />

because, during most of its discharge<br />

cycle, it will have a voltage<br />

between 14 <strong>and</strong> 16 V, which minimizes<br />

the voltage change of the highcurrent<br />

devices. The availability of<br />

integrated circuits <strong>and</strong> components<br />

for four-cell batteries was found to be<br />

better than an eight-cell approach,<br />

which would have a maximum voltage<br />

of 33.6 V <strong>and</strong> require larger 35 V tolerant<br />

components. Overall current required<br />

dictates some trace thicknesses<br />

of approximately 8 mm. The capacity<br />

of each cell is 21 Ah, which provides a<br />

total of 310 Wh for the pack. A commercial<br />

frontend battery circuit board<br />

protects the batteries from overcharging,<br />

over discharge, <strong>and</strong> short<br />

circuit. It also balances the cells to increase<br />

overall service life. The circuit<br />

has a very low resistance to have minimal<br />

impact on the efficiency. The<br />

charger built into the power distribution<br />

can charge the batteries in approximately<br />

12 h. The high capacity of the<br />

batteries allows them to be discharged<br />

at a rate lower than 0.5 C, which increases<br />

battery runtime <strong>and</strong> overall<br />

lifetime (Murphy et al., 1990).<br />

Power Distribution<br />

The power distribution is the most<br />

complex custom-designed circuit board<br />

in the vehicle made into a four-layer<br />

10 × 17 cm PCB. It regulates <strong>and</strong><br />

monitors the voltages to power<br />

the rest of the electronics on the<br />

submarine.<br />

The power distribution module<br />

uses a Microchip PIC18F45K22 as<br />

a microcontroller supervisor. This<br />

microcontroller monitors <strong>and</strong> controls<br />

voltages <strong>and</strong> current draws <strong>and</strong> parts or<br />

the whole system can be shut off if<br />

excessive current is drawn or voltages<br />

are not in an acceptable range. The<br />

presence of power sources is also monitored,<br />

<strong>and</strong> the microcontroller correctly<br />

decides which one to use <strong>and</strong><br />

whether a battery requires charging.<br />

The sbRIO can signal the supervisor<br />

to enable or disable certain regulators<br />

on the vehicle to save power (i.e., wireless<br />

bridge is off when submerged). A<br />

function of the microcontroller also<br />

allows the vehicle to enter a “sleep”<br />

mode where everything except the supervisor<br />

microcontroller is turned off<br />

for a pre-determined period of time<br />

which reduces the total power usage<br />

to about 1 W. The microcontroller<br />

has an onboard analog-to-digital converter<br />

<strong>and</strong> a 16-channel multiplexer<br />

is used to read all of the required<br />

voltages <strong>and</strong> currents. Voltages are<br />

measured using a resistor-divider network<br />

to reduce the voltage into the<br />

multiplexer, <strong>and</strong> therefore the microcontroller<br />

to between 0 <strong>and</strong> 5 V with<br />

some error margin in case the voltage<br />

rises to more than intended. Allegro<br />

ACS714 Hall effect current sensors<br />

are used to provide very low loss method<br />

current measurements (0.06 W at 5 A).<br />

Sensors<br />

CephaloBot incorporates a minimum<br />

set of sensors needed to maintain<br />

a heading <strong>and</strong> depth underwater.<br />

Acoustics provide the vehicle with<br />

a relative position to a static pinger.<br />

Intervehicle communication prevents<br />

collisions between vehicles <strong>and</strong> the<br />

walls of the pool. The vehicle is robust<br />

enough to withst<strong>and</strong> a collision if it<br />

does occur. When the vehicles are<br />

deployed to an ocean environment,<br />

payload sensors may be added heterogeneously<br />

to the vehicles <strong>and</strong> shared so<br />

that each submarine has all required<br />

data to successfully navigate its environment.<br />

To simplify the design,<br />

CephaloBot uses an all-in-one IMU<br />

solution from VectorNav. The device<br />

has an onboard three-axis accelerometer,<br />

gyroscope, <strong>and</strong> magnetometer. It<br />

performs Kalman filtering <strong>and</strong> outputs<br />

quaternions, Euler angles, <strong>and</strong> the raw<br />

sensor data to the sbRIO FPGA. The<br />

onboard filtering eliminates the need<br />

to develop or perform the computations<br />

on the sbRIO. Stated accuracies<br />

are less than 2°, <strong>and</strong> because a magnetometer<br />

provides an absolute reference,<br />

this accuracy will not degrade<br />

with time. A Honeywell pressure sensor<br />

provides fine resolution (0.01 m)<br />

measurements to 10-m depths. The<br />

device outputs an analog voltage. The<br />

test pool for the vehicle is 5-m deep,<br />

<strong>and</strong> so a higher resolution device as<br />

chosen over one that may be used to<br />

a deeper depth.<br />

Conclusion<br />

The CephaloBot provides an ideal<br />

low cost option for underwater sensor<br />

networking <strong>and</strong> hybrid vehicle applications.<br />

The vehicle has maneuvering<br />

capabilities at zero forward velocity necessary<br />

for docking <strong>and</strong> high-resolution<br />

sensing. This capability is provided by<br />

an array of novel squid <strong>and</strong> jellyfish<br />

inspired thrusters. The thrusters are<br />

located internal to the hull with only<br />

asmallorifice exposed to the outer<br />

flow minimizing the effect on forward<br />

drag <strong>and</strong> allowing for efficient<br />

162 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


high-speed transit. Additionally these<br />

thrusters only require the single<br />

opening, allowing for greater system<br />

freedom internal to the vehicle inbetween<br />

thrusters.<br />

The embedded controller system<br />

designed for the vehicle has a compact<br />

modular design allowing for a wide<br />

variety of possible mission objectives.<br />

The microcontroller, which is operated<br />

on an easily adaptable LabVIEW<br />

platform, includes several open connections<br />

for future mission operations,<br />

on top of the base level vehicle operation<br />

input/outputs. The vehicle also has<br />

a wide variety of communication options<br />

including a low cost <strong>and</strong> in house<br />

developed acoustic communication/<br />

localization system for communication<br />

between underwater vehicles <strong>and</strong><br />

support structures. The vehicle has an<br />

RF system for communication with<br />

aerial vehicles while on the surface,<br />

<strong>and</strong> a WIFI bridge for communication<br />

with testers <strong>and</strong> data loggers while in<br />

controlled laboratory environments.<br />

Acknowledgments<br />

The authors would like to thank<br />

S. Lawrence-Simon, Tyler Thomas,<br />

Ryan Delgizzi, Dan Ambrosio, Colin<br />

Miller, Mikhail Kosna, <strong>and</strong> Matt<br />

Rhode for their hours of working on<br />

design <strong>and</strong> fabrication of the vehicle.<br />

We would also like to thank the Office<br />

of Naval Research (code 34) for funding<br />

this research project.<br />

Corresponding Author:<br />

Kamran Mohseni<br />

231 MAE-A, Department of<br />

Mechanical <strong>and</strong> Aerospace<br />

Engineering<br />

University of Florida, Gainesville, FL<br />

Email: mohseni@ufl.edu<br />

References<br />

Allen, D.G. 2010. James Rumsey: Steamboat<br />

inventor or pioneer in steam navigation.<br />

Magazine of the Jefferson County Historical<br />

<strong>Society</strong>. 96:48-54.<br />

Anderson, E., & Grosenbaugh, M. 2005.<br />

Jet flow in steadily swimming adult squid.<br />

J Exp Biol. 208:1125-46. doi: 10.1242/<br />

jeb.01507.<br />

Barrett, D., Triantafyllou, M.S., Yue, D.K.P.,<br />

Grosenbaugh, M.A., & Wolfgang, M.J. 1999.<br />

Drag reduction in fish-like locomotion.<br />

J Fluid Mech. 392:183-212. doi: 10.1017/<br />

S0022112099005455.<br />

Bartol, I.K., Krueger, P.S., Stewart, W.J., &<br />

Thompson, J.T. 2009. Hydrodynamics of<br />

pulsed jetting in juvenile <strong>and</strong> adult brief squid<br />

Lolliguncula brevis: Evidence of multiple jet<br />

‘modes’ <strong>and</strong> their implications for propulsive<br />

efficiency. J Exp Biol. 212:1889-903.<br />

doi: 10.1242/jeb.027771.<br />

Bartol, I.K., Krueger, P.S., Thompson, J.T.,<br />

& Stewart, W.J. 2008. Swimming dynamics<br />

<strong>and</strong> propulsive efficiency of squids throughout<br />

ontogeny. Integr Comp Biol. 48(6):720-33.<br />

Clark, T., Klein, P., Lake, G., Lawrence-<br />

Simon, S., Moore, J., Rhea-Carver, B., …<br />

Wu, A. 2009. Kraken: Kinematically roving<br />

autonomously kontrolled electro-nautic.<br />

In: Aerospace Sciences Meeting. Orl<strong>and</strong>o, FL:<br />

AIAA.<br />

Dabiri, J.O., Colin, S.P., Katija, K., &<br />

Costello, J.H. 2010. A wake based correlate of<br />

swimming performance <strong>and</strong> foraging behavior<br />

in seven co-occurring jellyfish species. J Exp<br />

Biol. 213:1217-75. doi: 10.1242/jeb.034660.<br />

Dunbabin, M., Roberts, J., Usher, K.,<br />

Winstanley, G., & Corke, P. 2005. A hybrid<br />

auv design for shallow water reef navigation.<br />

In: Proceedings of the International Conference<br />

on Robotics <strong>and</strong> Automation. pp. 2105-10.<br />

Barcelona, Spain: IEEE. doi: 10.1109/<br />

ROBOT.2005.1570424.<br />

Fossen, T.I. 1991. Nonlinear modeling <strong>and</strong><br />

control of underwater vehicles. Ph.D. thesis,<br />

Norwegian Inst. Technol., Trondheim,<br />

Norway.<br />

Gharib, M., Rambod, E., & Shariff, K. 1998.<br />

A universal time scale for vortex ring formation.<br />

J Fluid Mech. 360:121-40. doi: 10.1017/<br />

S0022112097008410.<br />

Glezer, A. 1988. The formation of vortex rings.<br />

Phys Fluids. 31(12):3532-42. doi: 10.1063/<br />

1.866920.<br />

Kato, N. 2011. Swimming <strong>and</strong> walking of<br />

an amphibious robot with fin actuators.<br />

Mar Technol Soc J. 45(4):181-97.<br />

Krieg, M., & Mohseni, K. 2008. Thrust<br />

characterization of pulsatile vortex ring<br />

generators for locomotion of underwater<br />

robots. IEEE J Oceanic Eng. 33(2):123-32.<br />

doi: 10.1109/JOE.2008.920171.<br />

Krieg, M., & Mohseni, K. 2010. Dynamic<br />

modeling <strong>and</strong> control of biologically inspired<br />

vortex ring thrusters for underwater robot<br />

locomotion. IEEE T Robot. 26(3):542-54.<br />

doi: 10.1109/TRO.2010.2046069.<br />

Krieg, M., & Mohseni, K. 2011. Thrust<br />

enhancement from radial velocity in squid<br />

inspired thrusters. OCEANS 2011, in press.<br />

Kona, HI: MTS/IEEE.<br />

Krueger, P.S., Dabiri, J., & Gharib, M. 2006.<br />

The formation number of vortex rings formed<br />

in a uniform background co-flow. J Fluid<br />

Mech. 556(1):147-66. doi: 10.1017/<br />

S0022112006009347.<br />

Krueger, P.S., & Gharib, M. 2003. The<br />

significance of vortex ring formation to the<br />

impulse <strong>and</strong> thrust of a starting jet. Phys<br />

Fluids. 15(5):1271-81. doi: 10.1063/<br />

1.1564600.<br />

Li, Y., Kastner, R., Faunce, B., Domond, K.,<br />

Kimball, D., Schurgers, C., & Benson, B.<br />

2010. Design of a low-cost underwater<br />

acoustic modem. IEEE ESL. 2:58-61.<br />

doi: 10.1109/LES.2010.2050191.<br />

Licht, S. 2008. Biomimetic oscillating foil<br />

propulsion to enhance underwater vehicle<br />

agility <strong>and</strong> maneuverability. Ph.D. thesis,<br />

Massachusetts Institute of <strong>Technology</strong>,<br />

Cambridge, Massachusetts.<br />

Licht, S., Polidoro, V., Flores, M., Hover, F.,<br />

& Triantafyllou, M. 2004. Design <strong>and</strong><br />

projected performance of a flapping foil<br />

July/August 2011 Volume 45 Number 4 163


auv. IEEE J Oceanic Eng. 29(3):786-68.<br />

doi: 10.1109/JOE.2004.833126.<br />

Lipinski, D., & Mohseni, K. 2009. A numerical<br />

investigation of flow structures <strong>and</strong><br />

fluid transport with applications to feeding<br />

for the hydromedusae Aequorea victoria <strong>and</strong><br />

Sarsiatubulosa. J Exp Biol. 212:2436-47.<br />

doi: 10.1242/jeb.026740.<br />

Mclean, M. 1991. Dynamic performance of<br />

small diameter tunnel thrusters. Ph.D. thesis,<br />

Naval Postgraduate School, Monterey,<br />

California.<br />

Murphy, T., Cason-Smith, D., James, S., &<br />

Smith, P. 1990. Characterization of aa size<br />

lithium rechargeable cells. In: Power Sources<br />

Symposium, Proceedings of the 34th International.<br />

pp. 176-80. Cherry Hill, NJ: IEEE.<br />

Mohseni, K. 2004. Zero-mass pulsatile jets<br />

for unmanned underwater vehicle maneuvering.<br />

In: 3rd “Unmanned Unlimited” Technical<br />

Conference, Workshop <strong>and</strong> Exhibit.<br />

pp. 2004-3686. Chicago, IL: AIAA.<br />

Mohseni, K. 2006. Pulsatile vortex generators<br />

for low-speed maneuvering of small underwater<br />

vehicles. Ocean Eng. 33(16):2209-23.<br />

Mohseni, K., & Gharib, M. 1998. A model<br />

for universal time scale of vortex ring formation.<br />

Phys Fluids. 10:2436-8. doi: 10.1063/1.869785.<br />

O’Dor, R.K., & Webber, D. 1991. Invertebrate<br />

athletes: Trade-offs between transport<br />

efficiency <strong>and</strong> power density in cephalopod<br />

evolution. J Exp Biol. 160:93-112.<br />

Sahin, M., & Mohseni, K. 2009. An arbitrary<br />

Lagrangian-Eulerian formulation for the numerical<br />

simulation of flow patterns generated<br />

by the hydromedusa Aequorea Victoria.<br />

J Comp Physiol. 228:4588-605. doi: 10.1016/<br />

j.jcp.2009.03.027.<br />

Sahin, M., Mohseni, K., & Colins, S. 2009.<br />

The numerical comparison of flow patterns<br />

<strong>and</strong> propulsive performances for the hydromedusae<br />

Sarsia tubulosa <strong>and</strong> Aequorea Victoria.<br />

J Exp Biol. 212:2656-67. doi: 10.1242/<br />

jeb.025536.<br />

Shariff, K., & Leonard, A. 1992. Vortex<br />

rings. Annu Rev Fluid Mech. 34:235-79.<br />

doi: 10.1146/annurev.fl.24.010192.001315.<br />

Torsiello, K. 1994. Acoustic positioning of<br />

the NPS autonomous underwater vehicle<br />

(AUV II) during hover conditions. Master’s<br />

thesis, Naval Postgraduate School, Monterey,<br />

California.<br />

Wilson, M.M., Peng, J., Dabiri, J.O., &<br />

Eldridge, J.D. 2009. Lagrangian coherent<br />

structures in low Reynolds number swimming.<br />

J Phys-Condens Mat. 21. Art. 204105.<br />

doi: 10.1088/0953-8984/21/20/204105.<br />

Yoerger, D., Cooke, J., & Slotine, J.J. 1990.<br />

The influence of thruster dynamics on underwater<br />

vehicle behavior <strong>and</strong> their incorporation<br />

into control system design. IEEE J Oceanic<br />

Eng. 15(3):167-78. doi: 10.1109/48.107145.<br />

Rosenfeld, M., Katija, K., & Dabiri, J.O.<br />

2009. Circulation generation <strong>and</strong> vortex ring<br />

formation by conical nozzles. J Fluids Eng.<br />

131, 091204:1-8.<br />

Rosenfeld, M., Rambod, E., & Gharib, M.<br />

1998. Circulation <strong>and</strong> formation number<br />

of laminar vortex rings. J Fluid Mech.<br />

376:297-318.<br />

Sahin, M., & Mohseni, K. 2008. The numerical<br />

simulation of flow patterns generated<br />

by the hydromedusa Aequorea victoria using<br />

an arbitrary Lagrangian- Eulerian formulation.<br />

In: 38th Fluid Dynamics Conference <strong>and</strong><br />

Exhibit. pp. 2008-3715. Seattle: AIAA.<br />

164 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Modeling of Artificial Aurelia aurita<br />

Bell Deformation<br />

AUTHORS<br />

Keyur B. Joshi<br />

Alex Villanueva<br />

Colin F. Smith<br />

Shashank Priya<br />

Center for Energy Harvesting<br />

Materials <strong>and</strong> Systems,<br />

Center for Intelligent Material<br />

Systems <strong>and</strong> Structure,<br />

Virginia Polytechnic Institute<br />

<strong>and</strong> State University<br />

ABSTRACT<br />

Recently, there has been significant interest in developing underwater vehicles<br />

inspired by jellyfish. One of these notable efforts includes the artificial Aurelia aurita<br />

(Robojelly). The artificial A. aurita is able to swim with similar proficiency to the<br />

A. aurita species of jellyfish even though its deformation profile does not completely<br />

match the natural animal. In order to overcome this problem, we provide a systematic<br />

finite element model (FEM) to simulate the transient behavior of the artificial<br />

A. aurita vehicle utilizing bio-inspired shape memory alloy composite (BISMAC) actuators.<br />

The finite element simulation model accurately captures the hyperelastic<br />

behavior of EcoFlex (Shore hardness-0010) room temperature vulcanizing silicone<br />

by invoking a three-parameter Mooney-Rivlin model. Furthermore, the FEM incorporates<br />

experimental temperature transformation curves of shape memory alloy<br />

wires by introducing negative thermal coefficient of expansion <strong>and</strong> considers the<br />

effect of gravity <strong>and</strong> fluid buoyancy forces to accurately predict the transient deformation<br />

of the vehicle. The actual power cycle used to drive artificial A. aurita vehicle<br />

was used in the model. The overall profile error between FEM <strong>and</strong> the vehicle profile<br />

is mainly due to the difference in initial relaxed profiles.<br />

Keywords: autonomous undersea vehicle, Aurelia aurita, BISMAC, finite element<br />

analysis, transient dynamics<br />

Introduction<br />

J<br />

ellyfish have been in existence<br />

for millions of years <strong>and</strong> are the earliestknownmetazoansthatusemuscles<br />

for swimming (Valentine, 2004).<br />

They are found at various ocean<br />

depths <strong>and</strong> possess the ability to survive<br />

under hostile ocean environment.<br />

They exhibit colonial behavior <strong>and</strong><br />

have the ability to maintain certain<br />

depth <strong>and</strong> certain distance from the<br />

ocean shore (Albert, 2009). Jellyfish<br />

have relatively simple biological form<br />

<strong>and</strong> muscle architecture, lacking advanced<br />

sensors <strong>and</strong> a complex neural<br />

network possessed by many oceanic<br />

creatures (Chapman, 1974; Gladfelter,<br />

1972, 1973). but they are still able<br />

to survive <strong>and</strong> adapt in hostile environments.<br />

It maintains territorial existence<br />

by swimming on minimal energy<br />

intake. These abilities have created<br />

tremendous interest in the scientific<br />

community to discover their structureproperty-performance<br />

relationships<br />

<strong>and</strong> apply the learning towards creating<br />

a jellyfish-inspired swimming vehicle<br />

to perform various surveillance<br />

<strong>and</strong> monitoring tasks.<br />

There have been various efforts<br />

in literature on developing jellyfishinspired<br />

robots by using smart materialbased<br />

actuators. Inspired by the jetter<br />

class of jellyfish that swim by creating<br />

a jet of water by forcing it out of the<br />

bell, Villanueva et al. (2009) developed<br />

the JETSUM. This prototype used<br />

shape memory alloy (SMA) wires <strong>and</strong><br />

created an actuating stroke of bell<br />

segments attached to a passive neutrally<br />

buoyant bell structure. Yang<br />

et al. (2007) used flappers with control<br />

surfaces made of ionomeric polymer<br />

metal composites (IPMC) to<br />

control directionality of the vehicle.<br />

Tadesse et al. (2010a) used polypyrrole–<br />

polyvinylidene difluoride composites<br />

to achieve bending actuation to create<br />

a jellyfish robot. Larger jellyfish typically<br />

use “rowing” locomotion (Colin<br />

& Costello, 2002) <strong>and</strong> are characterized<br />

by formation of counter rotating<br />

starting <strong>and</strong> stopping vortex rings. Interaction<br />

of these vortex rings reduces<br />

energy lost in the wake <strong>and</strong> lends<br />

rowerstheirsuperiorswimmingefficiencies<br />

relative to jetters (Colin &<br />

Costello, 2002). Thus, with regard to<br />

energy efficiency, rowers provide a better<br />

platform for larger vehicles. Yeom<br />

<strong>and</strong> Oh (2009) proposed an entire<br />

jellyfish made from IPMC actuators<br />

with segments cut such that, on<br />

contraction, all the segments close to<br />

form a contracted bell shape. Recently,<br />

asignificant breakthrough was made<br />

by Villanueva et al. (2010b), who proposed<br />

the high-energy density bioinspired<br />

shape memory alloy composite<br />

July/August 2011 Volume 45 Number 4 165


(BISMAC) actuator that opened the<br />

possibility of converting high-force generation<br />

capability of SMA wires into<br />

high displacements. Using BISMAC,<br />

the design <strong>and</strong> implementation of<br />

biomimetic rowers became feasible.<br />

In this study, we investigate the<br />

bell deformation of BISMAC-based<br />

jellyfish robots using a finite element<br />

model (FEM, conducted using<br />

ANSYS) <strong>and</strong> identify the correlation<br />

with the natural species. In order to<br />

do so, the first major challenge was a<br />

precise implementation of SMA in<br />

FEM due to their giant aspect ratio<br />

<strong>and</strong> hysteretic temperature transformation.Wewereabletosuccessfully<br />

demonstrate the deformation of SMA<br />

using ANSYS by optimizing the meshing<br />

technique, identifying the variability<br />

in thermal coefficient of expansion,<br />

<strong>and</strong> separating the total deformation<br />

cycle into individual heating <strong>and</strong> cooling<br />

curves. Building upon this success,<br />

we implemented the SMA in BISMAC<br />

structure <strong>and</strong> investigated its mechanics<br />

to optimize the BISMAC configuration<br />

for mimicking the jellyfish<br />

profile. The FEM model provides the<br />

underst<strong>and</strong>ing of mechanism for bending<br />

strain amplification in BISMAC<br />

actuators <strong>and</strong> clearly delineates the<br />

effect of structural <strong>and</strong> thermal variables.UsingtheFEMresults,we<br />

were able to identify the inaccuracies<br />

that can occur due to variability in<br />

prototyping of BISMAC. Furthermore,<br />

our results provide important<br />

insight towards the development of<br />

feedback controller based on resistance<br />

changes. Next, we introduce the design<br />

of bell geometry in FEM for artificial<br />

Aurelia aurita (later referred to as,<br />

A. aurita for convenience in this work)<br />

using radial arrangement of BISMAC<br />

actuators. The objective was to develop<br />

the proper joint geometry that allows<br />

BISMACs to provide maximum deformation.<br />

Next, we describe the method for fabricating artificial A. aurita <strong>and</strong><br />

experimental characterization. Lastly, using the FEM simulations, we present<br />

a comparative analysis between the biological A. aurita (swimming profile)<br />

<strong>and</strong> artificial A. aurita <strong>and</strong> FE simulation.<br />

BISMAC Actuator<br />

The BISMAC actuator is a composite of an incompressible flexible metal strip<br />

<strong>and</strong> SMA wires separated by distance d <strong>and</strong> embedded in silicone rubber. Thermal<br />

transformation from martensite phase into austenite upon Joule heating induces<br />

the contraction of SMA wires, which is resisted by the incompressible metal<br />

strip. This introduces a tensile force f in the SMA that opposes contraction <strong>and</strong><br />

reduces the strain in SMA wire by a small amount.<br />

Figure 1 helps in explaining the mechanics of BISMAC bending deformation<br />

(Figure 1(a)). Under Joule heating, as SMA transforms from martensite<br />

to austenite phase <strong>and</strong> tends to contract, but due to the structure of the BISMAC,<br />

the metal strip resists this contraction <strong>and</strong> induces tensile force f in the SMA wire<br />

<strong>and</strong> compressive force f of the same magnitude in the metal strip. This force<br />

couple being distance d apart generates effective moment in the BISMAC actuator<br />

causing it to bend. Figure 1(b) shows the deformed BISMAC geometry. The<br />

mechanics of BISMAC has been discussed in detail by Smith et al. (2011). By<br />

using constitutive relation for SMA, we can express the generated stress as<br />

σ σ 0 ¼ E SMA ðɛ ɛ 0 ÞþΩðζ ζ 0 ÞþΘðT T 0 Þ ð1Þ<br />

where σ is the stress, ɛ is the strain, E SMA is the effective Young’s modulus of<br />

SMA, Ω is the transformation coefficient, ζ is the martensite fraction, Θ is the<br />

thermoelastic coefficient, <strong>and</strong> T is the temperature. Subscript zero denotes the<br />

initial state. Since transformation here is from fully martensite to fully austenite<br />

phase, ζ = 0, <strong>and</strong> ζ 0 = 1. Pure thermoelastic expansion is negligible, <strong>and</strong> initial<br />

stress <strong>and</strong> strain states are zero. Using E SMA = E austenite , since SMA is completely<br />

in austenite phase, Eq. (1) transforms into<br />

σ ¼ E austenite ɛ Ω ð2Þ<br />

FIGURE 1<br />

(a) Schematic of the force <strong>and</strong> moment in BISMAC. (b) Geometry of beam curvature.<br />

166 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


If there was no BISMAC structure to resist the SMA contraction, σ = 0 <strong>and</strong> ɛ = ɛ l<br />

transformation strain, providing<br />

Ω ¼ E austenite ɛ l<br />

If there is a uniform tensile force applied onto the SMA wire while transformation<br />

takes place,<br />

σ ¼ E austenite ɛ þ E austenite ɛ l<br />

Since SMA wires are thin with negligible bending stiffness, it is safe to assume<br />

uniform distribution of axial stress. Thus, axial stress can be written as<br />

σ ¼<br />

f<br />

A SMA<br />

¼ ɛ f E austenite ⇒ ɛ f ¼ ɛ þ ɛ l<br />

where f is the force generated by the SMA wire, A SMA is the total area of SMA<br />

cross section, <strong>and</strong> ɛ f is the force induced tensile strain. If SMA length after contraction<br />

is reduced from L to L′, from kinematics consideration, we can write<br />

L 0<br />

L ¼ 1 þ ɛ ¼ R<br />

d<br />

R<br />

where R is the radius of curvature of the metal strip <strong>and</strong> d is the distance between<br />

SMA wire <strong>and</strong> the metal strip. Solving Eq.(5) <strong>and</strong> (6),<br />

1 þ ɛ f ɛ l ¼ 1<br />

R ¼<br />

d<br />

R<br />

d<br />

ɛ l ɛ f<br />

ð8Þ<br />

Using Euler beam theory,<br />

M<br />

I ¼ E R ⇒ fd I ¼ E ɛ l<br />

d<br />

ɛ f E austenite A SMA d<br />

I<br />

ɛ f ¼<br />

¼ E ɛ l<br />

d<br />

EI ɛ l<br />

E austenite A SMA d 2 þ EI<br />

ɛ f<br />

<br />

ɛ f<br />

<br />

⇒ ɛ f ¼ EI ɛ l ɛ f<br />

<br />

E austenite A SMA d 2<br />

ð3Þ<br />

ð4Þ<br />

ð5Þ<br />

ð6Þ<br />

ð7Þ<br />

ð9Þ<br />

ð10Þ<br />

ð11Þ<br />

where EI is the total bending stiffness of the composite beam. For the particular<br />

configuration of the BISMAC used by Smith et al. (2011), Figure 2(a) shows the<br />

relationship between the radius of<br />

curvature for the BISMAC <strong>and</strong> the<br />

distance between SMA wires <strong>and</strong> flexible<br />

metal strip. Figure 2(b) reveals<br />

that as the distance d decreases, the<br />

tensile force induced in SMA wires<br />

increases dramatically <strong>and</strong> attains a limiting<br />

value of 80 g force for 100-μmthick<br />

SMA wire (BioMetal Fiber,<br />

TOKI Corporation) beyond which<br />

high stresses cause austenite phase to<br />

transform into stress-induced martensite<br />

resulting in loss of transformation<br />

strain <strong>and</strong> thus loss of performance,<br />

which is not accounted by Eqs. (1)-(11).<br />

BISMAC Customization<br />

<strong>and</strong> Bell Geometry<br />

A. aurita curvature profiles in its<br />

relaxed <strong>and</strong> contracted states are<br />

showninFigure3(a)(Dabirietal.,<br />

2005). Before any actuation, the bell<br />

is fully exp<strong>and</strong>ed <strong>and</strong> said to be in<br />

the relaxed position. Fully contracted<br />

state refers to the state corresponding<br />

to complete contraction of subumbrellar<br />

muscles <strong>and</strong> minimum bell volume.<br />

After actuation, the bell passively<br />

regains its original (relaxed) position.<br />

In earlier work (Villanueva et al.,<br />

2010b), we have presented the methodology<br />

used to customize BISMAC<br />

configurations to mimic the natural<br />

A. aurita curvature profile. For a<br />

curved beam, we can write<br />

M ¼ EI ðÞ s<br />

dðΔθÞ<br />

ds<br />

ð12Þ<br />

where M is the moment, EI(s) isthe<br />

local bending stiffness at location s,<br />

Δθ(s) =θ(s) − θ 0 (s) is the change in<br />

slope at location s, <strong>and</strong>s is location<br />

on curved profile length. Since the<br />

force generated by SMA wires is uniform,weensureaconstantmoment<br />

M by maintaining the fixed distance<br />

July/August 2011 Volume 45 Number 4 167


FIGURE 2<br />

(a) Radius of curvature vs. distance d. (b) Tensile force in the SMA wires vs. distance d.<br />

FIGURE 3<br />

(a) A. aurita profile in relaxed <strong>and</strong> contracted conditions. (b) Curvature comparison of BISMAC<br />

muscle with A. aurita after customization. (c) Schematic of the BISMAC placement in the artificial<br />

A. aurita.<br />

between SMA wires <strong>and</strong> the metal<br />

strip. The bending stiffness was varied<br />

(i) to match jellyfish exumbrella <strong>and</strong><br />

subumbrella profiles in the relaxed<br />

state <strong>and</strong> (ii) to manipulate the bending<br />

stiffness EI(s) such that upon SMA<br />

actuation, dðΔθÞ<br />

ds<br />

matches that of real<br />

A. aurita in order to ensure that we<br />

have a good match between natural<br />

<strong>and</strong> artificial vehicles in the contracted<br />

state. The resultant BISMAC achieved<br />

close similarity with A. aurita profile as<br />

illustrated in Figure 3(b). The inflexion<br />

point in contracted A. aurita profile<br />

at ∼2.5 cm represents the fact that<br />

at the center bell thickens <strong>and</strong> the profile<br />

becomes a little convex near the<br />

center <strong>and</strong> changes to concave at inflexion<br />

point. Towards the end of the<br />

profile, the BISMAC shows reduction<br />

in curvature due to passive material at<br />

the tip end for protection from water.<br />

Similar reductions in contracted<br />

A. aurita curvature profile towards<br />

the bell margin lead to discovery of<br />

the passive flap in A. aurita near the<br />

bell margin. The passive flap was<br />

shown to improve the swimming performance<br />

of the artificial A. aurita significantly<br />

(Villanueva et al., 2010a).<br />

Figure 3(c) shows a schematic of artificial<br />

A. aurita consisting of eight<br />

BISMACs radially distributed around<br />

the bell, which is made of soft silicone.<br />

Artificial A. aurita has shown bell deformation<br />

<strong>and</strong> kinematics as well as a<br />

swimming performance comparable<br />

to that of natural animal (Villanueva<br />

et al., 2010a). Since most of the available<br />

engineering materials have lower<br />

compliance compared to that of natural<br />

animal, the artificial A. aurita design<br />

includes joint structures between<br />

BISMAC actuators to localize the material<br />

folding upon contraction. These<br />

joint structures modify the original<br />

axisymmetric A. aurita bell shape <strong>and</strong><br />

increases similarity to the Cyanea capillata<br />

bell shape. The joints are wedgelike<br />

cavities <strong>and</strong> are found on natural<br />

jellyfish. The joint structure is described<br />

by Smith <strong>and</strong> Priya (2010)<br />

<strong>and</strong> is copied from the Polyorchis montereyensis.<br />

Since the cross-sectional<br />

shape of the joints varies with bell<br />

height, we take the equation below to<br />

represent the profile at midbell. The<br />

joint structure can be described as a<br />

piecewise function representing two<br />

symmetric sides of a single function,<br />

mirrored about the y-axis,<br />

y ¼ δjk ð<br />

δk<br />

xÞ<br />

jx<br />

ð13Þ<br />

This function has been formed to describe<br />

the joint shape across a 2-D<br />

plane with height j, half-widthk,<br />

<strong>and</strong> curvature δ as parameters. In this<br />

way, the joint structure from a wide<br />

variety of species can be described<br />

with one basic equation by changing<br />

168 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 4<br />

(a) Original axisymmetric A. aurita bell (b). Schematic of joint geometry (Smith & Priya, 2010).<br />

(c) Top <strong>and</strong> side view showing location in bell for which joint calculations were made. (d) Final<br />

artificial A. aurita bell shape.<br />

−2.75. The final shape was evolved<br />

by sweeping circular section at bell<br />

tip into the curve at mid-bell height<br />

joint section for a smooth blending.<br />

Figure 4 shows the evolution of<br />

original axisymmetric A. aurita bell<br />

shape into the final bell design of artificial<br />

A. aurita. Figure 5 shows the details<br />

of the unigraphics model that was<br />

used in finite element (FE) simulation.<br />

the constants. Artificial A. aurita’s<br />

joint structure was produced by using<br />

a program that could manipulate the<br />

constants in Mathematica. The values<br />

to match the shape of joints in P. montereyensis<br />

were found to be height =<br />

2.15, half-width = 3, <strong>and</strong> curvature =<br />

FIGURE 5<br />

Unigraphics model of the artificial A. aurita.<br />

Experimental Setup<br />

Artificial A. aurita consists of a<br />

central mount that houses the electrical<br />

circuitry <strong>and</strong> clamps the BISMAC<br />

actuators together (see Figure 6). The<br />

radius of this hub was 25% of the bell<br />

diameter <strong>and</strong> covers a region where<br />

minimal deformation is expected to<br />

occur in both the natural <strong>and</strong> artificial<br />

A. aurita. Since the distance between<br />

the SMA wires <strong>and</strong> the metal strip is<br />

very crucial parameter in BISMAC design,<br />

the SMA wires <strong>and</strong> the metal<br />

strip are slide in position in small<br />

acrylic supports, designed to maintain<br />

this distance. The supports are then<br />

placed in the mold <strong>and</strong> silicone is<br />

poured in to settle (Villanueva et al.,<br />

2010a). The actual vehicle has a<br />

small portion (3-4 mm) of metal strip<br />

<strong>and</strong> SMA wires protruding out of the<br />

bell geometry. This is neglected in this<br />

work for model simplification. The<br />

simplification is justified by negligible<br />

bending moment contribution from<br />

the protruded part towards the bell<br />

deformation. The experimental data<br />

used for the deformation comparison<br />

were acquired by using the same artificial<br />

A. aurita <strong>and</strong> experimental setup as<br />

FIGURE 6<br />

Artificial A. aurita with uniform bell <strong>and</strong> flap,<br />

in the relaxed configuration.<br />

July/August 2011 Volume 45 Number 4 169


described in previous work (Villanueva<br />

et al., 2010a). Artificial A. aurita was<br />

submerged underwater <strong>and</strong> clamped<br />

down by supports pressing on the top<br />

<strong>and</strong> bottom of the bell. The contacting<br />

area between supports <strong>and</strong> bell covered<br />

the region of the internal central<br />

mount. This region is meant to undergo<br />

negligible deformation since<br />

BISMAC actuators do not directly deform<br />

the bell at that location. The bell<br />

was contracted by heating the SMA<br />

wires using a rapid heating control algorithm<br />

(Villanueva & Priya, 2010a).<br />

This controller uses SMA resistance<br />

feedback to monitor the bell state of<br />

deformation. It sends high-current<br />

pulses for rapid contraction <strong>and</strong> low<br />

current to maintain deformation<br />

allowing fast contraction while minimizing<br />

power consumption. The controller<br />

was developed in LabView<br />

(National Instruments), <strong>and</strong> the measurements<br />

were made using a NI<br />

cDAQ 9172 with NI-9215 <strong>and</strong><br />

NI-9263 analog input <strong>and</strong> output<br />

cards, respectively. A NF HAS 4052<br />

power amplifier was used to amplify<br />

the DAQ output <strong>and</strong> actuate the<br />

robot. The bell profile was recorded<br />

during the first actuation cycle. The<br />

deformation was captured using an<br />

IN250 high-speed camera from Fastec<br />

Imaging. The bell deformation was<br />

tracked by placing reflective beads<br />

along the profile <strong>and</strong> by processing<br />

the images manually using ImageJ.<br />

This process included an error on the<br />

order of ±1 cm.<br />

FEM Setup <strong>and</strong> Solution<br />

Material Properties<br />

To ensure accuracy of the model in<br />

adequately representing the behavior<br />

of various materials, we determined<br />

the properties experimentally. There<br />

are three materials critical to the simulation:<br />

silicone matrix, flexible but incompressible metal strip <strong>and</strong> SMA wires<br />

(BioMetal Fiber).<br />

Silicone Rubber<br />

Room temperature vulcanizing (RTV) silicone is a c<strong>and</strong>idate material for bell<br />

mesoglea. In addition to forming the main jellyfish body, it is also responsible for<br />

maintaining the required distance between spring steel <strong>and</strong> the SMA wires. We<br />

have tested several silicone rubbers with different shore hardness to evaluate their<br />

mechanical properties. Ecoflex TM (Smooth-On) with initial tensile Young’s modulus<br />

of the order of 10,580 Pa was selected to construct the Artificial A. aurita<br />

because it offered much less resistance to the BISMAC deflection. Figure 7(a)<br />

shows the tensile test, on a dog bone–shaped sample with gauge length of 7 mm<br />

<strong>and</strong> cross-sectional area of 2.45 × 2.8 mm 2 for several cycles at room temperature<br />

at 1 mm/s displacement rate up to 20 mm extension to ensure silicone properties<br />

do not change with multiple loading cycles. Hysteresis is evident in the figure, <strong>and</strong><br />

we used the average of the two curves to generate our model. Figure 7(b) represents<br />

completely defined stress-strain behavior of silicone including compression<br />

test data carried out on 16- mm-thick 25.4-mm diameter cylindrical specimen.<br />

We generated a three-parameter Mooney-Rivlin model for silicone from test<br />

data using ANSYS’s curve fitting tool, as a more conventional two-parameter<br />

Mooney-Rivlin model failed to provide a good fit to the experimental data.<br />

The Mooney-Rivlin model for Ecoflex TM is given by Eq. (14),<br />

W ¼ c 10 ðI 1 3Þþc 01 ðI 2 3Þþc 11 ðI 1 3ÞðI 2 3Þþ 1 d ðJ 1Þ2<br />

ð14Þ<br />

where W is the strain energy function, I 1 , I 2 , I 3 are the stretch invariants, J is the<br />

determinant of deformation gradient tensor, <strong>and</strong> c 10 = 2307.1 Pa, c 01 = −223.76 Pa,<br />

c 11 =142.83Pa,d =0Pa −1 (compressibility parameter). Silicone thermal conductivity<br />

was measured to be 0.22 W/m K. Silicone properties used in the<br />

FEM are tabulated in Table 1.<br />

SMA Wire<br />

We selected 100-μm diameter BioMetal fibers due to their superior performance<br />

(Tadesse et al., 2010b). The temperature transformation of the wires was<br />

measured experimentally as shown in Figure 8(a). Figure 8(b) represents the<br />

FIGURE 7<br />

(a) Tensile test for several cycles showing hysteresis. (b) Stress-strain curve used to model<br />

Ecoflex TM .<br />

170 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


TABLE 1<br />

Silicone properties used in the FEM.<br />

Property<br />

Value<br />

Thermal Conductivity<br />

0.22 W/m K<br />

Elastic modulus<br />

Mooney-Rivlin model<br />

Poisson’s ratio 0.49<br />

Density 982 kg/m 3<br />

Specific heat<br />

300 J/kg K<br />

TABLE 2<br />

SMA properties used in the FEM.<br />

Property Martensite Austenite<br />

Thermal Conductivity 8 W/m K 18 W/m K<br />

Elastic modulus 28 MPa 75 MPa<br />

Poisson’s ratio 0.33 0.33<br />

Density 6450 kg/m 3 6450 kg/m 3<br />

Specific heat 837.36 J/kg K 837.36 J/kg K<br />

stress-strain relationship as measured<br />

by tensile test to confirm the manufacturer’s<br />

claims (Toki Corporation) that<br />

the wires can easily take 400 MPa<br />

stress.<br />

Other properties that were used in<br />

the simulation are tabulated in Table 2.<br />

For temperatures other than transformation<br />

temperatures, the properties<br />

were interpolated according to martensite<br />

fraction in the SMA. Martensite<br />

fraction was calculated based on the<br />

temperature-strain curve.<br />

In order to model the temperature<br />

transformation hysteresis of SMA<br />

wires, we defined two separate material<br />

curves, one for the heating profile <strong>and</strong><br />

other for the cooling profile. Accordingly,<br />

the elastic modulus EX <strong>and</strong> thermal<br />

conductivity KXX also followed<br />

two separate profiles as depicted in Figures<br />

9(a) <strong>and</strong> 9(b). Figure 9(c) shows<br />

the artificially defined negative thermal<br />

coefficient of expansion to achieve the<br />

transformation strains with increase in<br />

temperature during heating <strong>and</strong> cooling,<br />

respectively.<br />

Metal Strip<br />

For choosing the metal strip, two<br />

criteria should be met: (i) incompressibility<br />

for BISMAC mechanics to work<br />

well <strong>and</strong> (ii) flexibility (low bending<br />

stiffness) to obtain maximum<br />

deformation with a small actuation<br />

moment. We selected st<strong>and</strong>ard spring<br />

steel (low carbon steel) as the suitable<br />

metal strip material with properties<br />

tabulated in Table 3.<br />

Transient Heat Transfer Model<br />

Meshing, Boundary Conditions,<br />

<strong>and</strong> Loads<br />

We chose ANSYS as our simulation<br />

package. To simulate transient<br />

heat transfer, we built the model<br />

by meshing the SMA, metal strip<br />

<strong>and</strong> one element thick silicone layer<br />

around them with SOLID70 (8-node<br />

brick element). The rest of the silicone<br />

matrix was meshed with SOLID87<br />

(10-node tetrahedrons) <strong>and</strong> SOLID90<br />

(20-node brick elements) was used as<br />

transition element between SOLID70<br />

FIGURE 8<br />

(a) BioMetal Fiber temperature transformation curve (A s = Austenite start temperature, A f = Austenite finish temperature, M s = Martensite start<br />

temperature, M f = Martensite finish temperature). (b) Stress-strain relationship of Martensite phase.<br />

July/August 2011 Volume 45 Number 4 171


FIGURE 9<br />

(a) Variation of SMA Young’s modulus with temperature. (b) Variation of SMA thermal conductivity<br />

with temperature. (c) Variation of thermal coefficient of expansion with temperature.<br />

<strong>and</strong> SOLID87 element meshes as shown<br />

in Figure 10(a). All these elements have<br />

TEMP degree of freedom.<br />

We took advantage of the circular<br />

symmetry <strong>and</strong> modeled only 1/8th of<br />

the bell segment with no loss of physics.<br />

To reduce the meshing efforts<br />

we chose to mesh only half of the<br />

1/8th segment of the bell, reflected<br />

the mesh about the plane of symmetry<br />

<strong>and</strong> finally merged the nodes to create<br />

the complete model. The boundary<br />

FIGURE 10<br />

(a) Mesh detail for transient thermal model. (b) Boundary conditions.<br />

conditions are depicted in Figure 10(b).<br />

Due to symmetry, the circular symmetric<br />

sides (Figure 10(b), violet colored<br />

faces) of the model do not have<br />

any temperature gradient; thus, they<br />

are modeled as insulated boundaries.<br />

Exumbrella <strong>and</strong> subumbrella surfaces<br />

(Figure 10(b), yellow colored faces)<br />

are in contact with surrounding water<br />

<strong>and</strong> conveys heat into the fluid, <strong>and</strong><br />

thus, were modeled as convective<br />

boundaries. The heat transfer coefficient<br />

of both these surfaces were<br />

taken to be 20 W/m 2 K (Baker, 1972),<br />

<strong>and</strong> the ambient temperature was<br />

fixed at 25°C. The thermal loading<br />

resulting from electrical heating was<br />

applied using Villanueva et al.’s resistance<br />

feedback control algorithm<br />

(Villanueva & Priya, 2010a) to reduce<br />

power requirement. The internal heating<br />

load was applied on the SMA wire<br />

as shown in Figure 11, which is consistently<br />

1/8th of total power consumed<br />

by the vehicle’s eight segments. It consists<br />

of rapid heating pulses of high<br />

TABLE 3<br />

Metal strip properties used in the FEM.<br />

Property<br />

Value<br />

Thermal Conductivity 47 W/m K<br />

Elastic modulus 210 GPa<br />

Poisson’s ratio 0.3<br />

Density 7860 kg/m 3<br />

Specific heat<br />

510 J/kg K<br />

172 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


current followed by a constant current<br />

regime <strong>and</strong> finally reducing the magnitude<br />

to idling minimum current governed<br />

by need for resistance feedback<br />

measurement. The three initial spikes<br />

seen in the curve correspond to the<br />

high-current impulse sent for rapid heating.<br />

Multiple pulses are usually needed<br />

during the first few actuation cycles<br />

since the low current is not enough to<br />

maintain the deformation. The material<br />

surrounding the SMA warms up <strong>and</strong><br />

eventually the low current input is<br />

enough to maintain deformation.<br />

Solution: Transient Thermal Analysis<br />

We first find the solution with<br />

SMA wires having material properties<br />

corresponding to heating curve. To<br />

capture initial sharp temperature rise<br />

accurately over the first 0.12 s, small<br />

time steps of 0.01 s were used to provide<br />

sufficient time resolution. The<br />

low current period from 0.12 to 0.70 s<br />

was solved in 100 uniform time steps.<br />

After solving two more transient current<br />

time steps at 0.71 <strong>and</strong> 0.72 s in<br />

one time step each, we switched the<br />

SMA wire material to the one with<br />

properties corresponding to cooling<br />

curve. Temperatures of SMA wires<br />

corresponding to t =0.72sbeingfar<br />

beyond A f ensures properties of both<br />

the curves are same at this point. Finally,<br />

relaxation phase of A. aurita<br />

jellyfish vehicle from 0.72 to 2.00 s<br />

was solved in 200 uniform time steps.<br />

FIGURE 11<br />

Variation of internal heating load on SMA wire with time.<br />

FIGURE 12<br />

Transient Structural<br />

Deformation Model<br />

Meshing, Boundary Conditions,<br />

<strong>and</strong> Loads<br />

To simulate transient structural<br />

deformation, we reused the mesh<br />

from transient heat transfer model<br />

for rapid model building. SOLID70<br />

(8-node brick element) of the SMA,<br />

metal strip <strong>and</strong> one element thick silicone<br />

layer around them were transformed<br />

into SOLID185 (8-node<br />

brick element) with structural degree<br />

of freedom UX, UY, UZ. Similarly,<br />

SOLID87 (10-node tetrahedrons)<br />

was transformed into SOLID187<br />

(10-node tetrahedrons) elements <strong>and</strong><br />

transition elements SOLID90 (20-<br />

node brick elements) were transformed<br />

into SOLID186 (20-node<br />

brick elements) as shown in Figure<br />

12(a). Figure 12(b) displays the<br />

displacement boundary condition for<br />

the model. The boundary condition<br />

was applied in cylindrical coordinate<br />

system defined using jellyfish vehicle<br />

centralaxisasthez-axis of the cylindrical<br />

coordinate system (CSYS = 5).<br />

(a) Transient structural deformation model mesh. (b) Displacement boundary conditions. (c) Temperature<br />

<strong>and</strong> acceleration boundary conditions.<br />

July/August 2011 Volume 45 Number 4 173


Both sides of 1/8th segment being circular symmetric were constrained from<br />

moving in hoop direction (UY = 0). The line on the axis of the vehicle was<br />

also constrained from moving in radial direction (UX = 0). Also, to constrain<br />

any rigid body translation, we chose to constrain one node on rigid central hub<br />

lying on the axis of the vehicle from moving in axial direction (UZ = 0).<br />

We accounted for gravitation <strong>and</strong> buoyancy force of the water by applying<br />

inertial acceleration corresponding to reduced gravitation under buoyancy from<br />

Eq. (15),<br />

<br />

ACEL ‐<br />

Y ¼ ΣN mat¼1 V matρ mat ρ water Σ N mat¼1 V <br />

mat<br />

Σ N mat¼1 V 9:81 m<br />

matρ mat<br />

s 2 ð15Þ<br />

where ACEL_Y is the inertial acceleration for the vehicle, mat is the material index,<br />

ρ is the density, <strong>and</strong> V is the volume. Earth’s gravitation acceleration was taken to<br />

be 9.81 m/s 2 . Temperatures were applied as body forces <strong>and</strong> were read from previously<br />

solved transient heat transfer model result. To account for the fluid drag<br />

we have introduced damping by defining BETAD = 0.03. This creates a damping<br />

matrix [C] =BETAD[K ], where [K ] = stiffness matrix of the FEM.<br />

The overall FE system equation can be written as<br />

½M<br />

ŠfÜ<br />

gþ ½CŠ<br />

:<br />

U þ ½K<br />

ŠfU<br />

g ¼ fFg<br />

ð16Þ<br />

Here, [M], [C]<strong>and</strong> [K ] are displacement dependent on the mass matrix, the<br />

damping matrix <strong>and</strong> the stiffness matrix, respectively, {U} is the displacement<br />

vector <strong>and</strong> {F } is effective force vector. The system defined by Eq. (16) is nonlinear<br />

due to variable [M ], [C], <strong>and</strong> [K ] matrices.<br />

Solution: Transient Structural Deformation Analysis<br />

As in transient heat transfer model, we start the modeling with SMA wires having<br />

properties corresponding to heating curve (martensite to austenite temperature<br />

transformation) for contraction phase of the cycle <strong>and</strong> during relaxation we<br />

use material corresponding to cooling curve (austenite to martensite temperature<br />

transformation). In simulation, this was achieved by defining two separate SMA<br />

material property curves corresponding to (martensite (M) to austenite (A) <strong>and</strong> austenite<br />

to martensite) <strong>and</strong> switching from material with M → A curve to material with<br />

A → M curve after completion of contraction phase. As SMA wire is well above austenite<br />

finish temperature A f , for which both the material curves (cooling <strong>and</strong> heating)<br />

have identical properties there’s no abrupt jump between these curves. We used the<br />

time intervals as used in transient heat transfer model.<br />

Results <strong>and</strong> Discussion<br />

Transient Heat Transfer Analysis<br />

Figure 13 summarizes all major results of transient heat transfer analysis. Figure<br />

13(a) shows overall temperature distribution at the end of heating phase. It<br />

suggests that practical region of interest is concentrated near SMA wire, <strong>and</strong> for<br />

most part, it does not change along SMA wire length. Near the edge of the bell,<br />

however, model predicts a little more temperature rise in the SMA wire compared<br />

to rest of the area due to very thin silicone<br />

layer as same amount of generated<br />

heat is absorbed by lesser silicone<br />

available. Since rate of heat conduction<br />

into silicone is higher compared to the<br />

rate of heat being convected away at<br />

the subumbrella surface, silicone temperature<br />

rises faster <strong>and</strong> temperature<br />

gradient between SMA wire <strong>and</strong> silicone<br />

reduces, inhibiting heat conduction<br />

from SMA wire to subumbrella<br />

surface. Figure 13(b) shows temperature<br />

history at a typical location at<br />

SMA center <strong>and</strong> in silicone 1 element<br />

away from SMA wire surface. It shows<br />

typical response of first order system to<br />

a given excitation. Initial high-current<br />

pulses till t = 0.12 s result in fast temperature<br />

rise at SMA center which<br />

keep increasing at logarithmic rate<br />

from t = 0.12 s to t = 0.70 s. Temperature<br />

at SMA center decreases exponentially<br />

from t = 0.70 s to t = 2.00 s<br />

but fails to return to starting temperature<br />

of 25°C at the end of the cycle<br />

(T =45°C at t = 2.00 s) suggesting<br />

net heat accumulation in the model.<br />

It also suggests that we may be adding<br />

heat unnecessarily at higher rate<br />

during low current constant heating<br />

cycle as the temperature at the end of<br />

high current pulsed heating is well<br />

beyond austenite finish temperature<br />

A f . This additional heating also aggravates<br />

heat accumulation problem that<br />

eventually results in partial loss of performance<br />

of the actuator as insufficient<br />

cooling results into incomplete transformation<br />

into martensite phase. Temperature<br />

in silicone 1 element away<br />

from SMA surface shows quite similar<br />

trend but has reduced temperature rise<br />

peaks during high-current pulse cycle<br />

as low conductivity dampens out the<br />

sharp peaks. Cooling cycle starts with<br />

significant temperature gradient between<br />

SMA center <strong>and</strong> the aforementioned<br />

silicone location but diminishes<br />

174 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 13<br />

(a) Overall temperature distribution at t = 0.70 s. (b) Temperature time history at SMA wire<br />

center <strong>and</strong> in silicone 1 element away from SMA surface. (c) Typical temperature distribution<br />

along SMA wire length.<br />

exponentially fast as is evident from Figure<br />

13(b). Figure 13(c) compares temperature<br />

distribution near SMA length<br />

at typical location for key time points<br />

t = 0.12 s (end of high-current pulse<br />

heating), t = 0.70 s (end of constant<br />

low current heating), <strong>and</strong> t = 2.00 s<br />

(end of cooling cycle).<br />

Transient Structural<br />

Deformation Analysis<br />

Overall Bell Deformation<br />

Since, gravity <strong>and</strong> buoyancy forces<br />

are accounted for, in the two equilibrium<br />

positions (relaxed <strong>and</strong> contracted),<br />

where fluid pressure differential does<br />

not exist across subumbrella <strong>and</strong><br />

exumbrella due to the FEM <strong>and</strong> the<br />

artificial A. aurita being held at the<br />

bell center; we can predict deformation<br />

at these positions accurately<br />

at this location without worrying<br />

about approximations involved in<br />

fluid drag. Figure 14(a) shows various<br />

views of FE simulation result<br />

for contracted state at t =0.12s.<br />

Original undeformed model is shown<br />

with only black edges superimposed<br />

on deformed model for comparison.<br />

Figure 14(a.1) represents bottom<br />

view of the contracted artificial<br />

A. aurita bell model, which qualitatively<br />

matches the contracted bell<br />

segments of C. capillata species (Figure<br />

14(d)) <strong>and</strong> experimental artificial<br />

FIGURE 14<br />

(a) FE result for artificial A. aurita bell contraction: (a.1) bottom view, (a.2) isometric view from bottom, (a.3) cross-sectional view at BISMAC<br />

location, <strong>and</strong> (a.4) side view. (b) Biological A. aurita contraction. (c) Artificial A. aurita experimental contraction. (d) C. capillata bell segment<br />

contraction bottom view. (e) Artificial A. aurita experimental contraction bottom view (Villanueva, submitted).<br />

July/August 2011 Volume 45 Number 4 175


A. aurita contraction (Figure 14(e))<br />

that we have intended to model in<br />

our FE simulation. Note that due to<br />

manufacturing imperfections not all<br />

the bell segments deform the same<br />

amount (Figure 14(e)). The comparison<br />

of contracted profile with that of<br />

C. capillata is justified as our artificial<br />

A. aurita possess joint structure <strong>and</strong><br />

radial muscles that were inspired by<br />

C. capillata possessing similar radial<br />

muscle arrangement (Gladfelter,<br />

1973). Figure 14(a.2) depicts isometric<br />

view of deformed model for<br />

better visualization. Figure 14(a.3)<br />

shows cross-sectional view taken along<br />

BISMAC center line. Figure 14(a.4)<br />

exhibits side view of the contracted<br />

FEM also matching well with natural<br />

A. aurita shown in Figure 14(b) <strong>and</strong><br />

experimental artificial A. aurita (Figure<br />

14(c)). It is evident that the bell<br />

curves inwards at the BISMAC locations<br />

radially <strong>and</strong> axially well <strong>and</strong><br />

confirms that the joint design provided<br />

by Smith <strong>and</strong> Priya (2010) is<br />

effective in assisting the bell deformation<br />

at the BISMAC locations.<br />

This behavior was also confirmed in<br />

experimental A. aurita deformation<br />

(Figure 14(c)).<br />

Comparison of Deformation<br />

at the BISMAC Location<br />

<strong>and</strong> at the Joint Location<br />

Figures 15(a)-15(f) reveal radial<br />

displacements in the bell at the<br />

BISMAC cross-section <strong>and</strong> in the joint<br />

cross-section, respectively, at the key<br />

time points (t = 0.12 s, t = 0.70 s <strong>and</strong><br />

t = 2.0 s); each group uses a common<br />

color legend for ease of comparison.<br />

It is obvious that the deformation is<br />

negligible near the centre of the jellyfish<br />

bell <strong>and</strong> maximum at the tip of<br />

the bell. At joint location, the bell is<br />

practically undeformed.<br />

Figure 16(b) plots the time history<br />

of radial <strong>and</strong> axial tip displacement at<br />

BISMAC location, which conforms<br />

to second order system response<br />

to step excitation. In fact, the highpower<br />

pulse heating is done at much<br />

higher frequency than the bell is able<br />

to respond, thus inertia of the bell<br />

acts as low-pass filter for the pulsed<br />

heating excitation. But as discussed<br />

in transient heat transfer results, SMA<br />

goes through complete martensite to<br />

austenite transformation in this period,<br />

contracting due to phase transformation<br />

strain <strong>and</strong> achieves maximum<br />

deformation of U radial =1.31mm<br />

<strong>and</strong> U axial =7.87mmatt =0.12s.<br />

At the maximum deformation point,<br />

the inertial forces <strong>and</strong> elastic restoring<br />

forces are not in equilibrium. The bell<br />

has a slight overshoot above equilibrium<br />

position due to inertia. During<br />

moderate constant current excitation<br />

from t = 0.12s to t =0.70sastemperature<br />

keeps rising, without any<br />

additional benefit toSMAtransformation<br />

strain the structure relaxes<br />

under stored strain energy during<br />

rapid contraction <strong>and</strong> after a small oscillation<br />

around equilibrium condition<br />

finally settles to U radial = 10.87 mm<br />

<strong>and</strong> U axial = 6.04 mm. During relaxation<br />

phase from t = 0.70 s to t = 2.0 s,<br />

SMA goes through austenite to<br />

martensite transition as temperature<br />

drops rapidly <strong>and</strong> returns to original<br />

configuration after little oscillations as<br />

FIGURE 15<br />

(a-c) Deformed artificial A. aurita bell at BISMAC location at different time t = 0.12 s, 0.70 s, <strong>and</strong> 2.0 s (d-f) Deformed artificial A. aurita bell at joint<br />

location at different time t = 0.12 s, 0.70 s <strong>and</strong> 2.0 s.<br />

176 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 16<br />

(a) Comparison of the FE results with artificial A. aurita bell deformation <strong>and</strong> natural A. aurita.<br />

(b) Time history of A. aurita tip displacement at BISMAC location.<br />

FIGURE 17<br />

is evident from Figures 16(b), 15(a)-<br />

15(b), <strong>and</strong> 17(a).<br />

Figure 16(a) compares the relaxed<br />

<strong>and</strong> contracted position of the<br />

simulated A. aurita bell with natural<br />

A. aurita <strong>and</strong> robotic A. aurita. The<br />

natural A. aurita with its muscle contracting<br />

in excess of 40% clearly<br />

achieves the highest deformation,<br />

not being fully matched by either<br />

artificial A. aurita or FE simulation.<br />

Experimental deformation on artificial<br />

jellyfish was measured through<br />

image processing thus only exumbrella<br />

points are available for comparison.<br />

Experimental model, however,<br />

has changed its form during manufacturing,<br />

which was designed to match<br />

FEM in relaxed state. This results in<br />

apparent mismatch in FE prediction<br />

from the experimental deformation.<br />

Inspite of this mismatch, the overall<br />

deformations are comparable. Figure<br />

16(b) shows comparison of time<br />

response of the transient tip displacements<br />

between FE simulation <strong>and</strong> artificial<br />

<strong>and</strong> natural A. aurita. It should<br />

be noted that the natural A. aurita<br />

Trace of tip displacement at BISMAC location by (a) FE simulation <strong>and</strong> (b) artificial A. aurita<br />

experiment <strong>and</strong> (c) natural A. aurita.<br />

swims freely in water moving forward,<br />

while artificial (experimental) A. aurita<br />

<strong>and</strong> the model are held at the bell center<br />

<strong>and</strong> does not move it water. Also,<br />

for the particular cycle for which data<br />

is obtained has cycle time of 1.7 s instead<br />

of 2 s. Natural A. aurita contracts<br />

<strong>and</strong> relaxes very smoothly <strong>and</strong><br />

does not have any steady equilibrium<br />

position. Artificial A. aurita achieves<br />

about twice as much radial displacement<br />

as predicted by simulation but<br />

follows similar trend of sharp rise<br />

(fall), overshooting above the equilibrium<br />

position beyond t =0.12s,relaxing<br />

to equilibrium position during<br />

t = 0.12 s to t = 0.70 s due to stored<br />

strain energy in the bell during contraction<br />

cycle. In relaxation cycle, as<br />

SMA cools down <strong>and</strong> undergoes austenite<br />

to martensite transition, the<br />

bell returns back to its original location<br />

slightly overshooting beyond<br />

equilibrium position before returning<br />

back to relaxed configuration. FE simulation<br />

captures this entire physics<br />

perfectly. Artificial A. aurita returns<br />

to relaxed position slower than predicted<br />

by model, suggesting that<br />

heat gets conducted away from SMA<br />

more slowly than we predict. However,<br />

artificial A. aurita axial displacement<br />

being too small; gets affected by<br />

finite resolution of image processing<br />

<strong>and</strong> shows inconclusive trend. FE<br />

simulation predicts axial tip displacement<br />

time history similar to that of<br />

radial tip displacement but reduced<br />

in magnitude.<br />

Figure 17(a) exhibits the trace of<br />

the bell tip at BISMAC location <strong>and</strong><br />

suggests that bell tip does not follow<br />

the same path while relaxing to its original<br />

configuration than it did during<br />

contraction. It also illustrates the<br />

overshoot around the equilibrium positions<br />

(relaxed <strong>and</strong> contracted). Tip<br />

displacement for the artificial A. aurita<br />

July/August 2011 Volume 45 Number 4 177


(Figure 17(b)) shows different trace patterns during contraction <strong>and</strong> relaxation as<br />

predicted by the model. Since artificial A. aurita begins with different relaxation<br />

configuration, this is expected. Interaction with surrounding fluid would cause<br />

the tip displacement trace to change which are not accommodated in the<br />

model. Oscillations in the trace of the experimental results are partly due to tracking<br />

error. The magnitude of the oscillations is close to the predicted error from the<br />

tracking method. Figure 17(c) represents trace of tip displacement of natural<br />

A. aurita scaled to the same bell diameter. Natural A. aurita is freely moving in<br />

water. The direction of contraction <strong>and</strong> relaxation is quite similar to the one predicted<br />

by the model (Figure 17(a)); however, relative location of the path during<br />

contraction <strong>and</strong> relaxation is reversed. We believe that this is caused by the fluid<br />

forces <strong>and</strong> the freely forward motion present in natural A. aurita.<br />

For a more quantitative comparison, we calculated the curvature of the profiles<br />

by calculating radius of circle passing through three consecutive points,<br />

RP ð 2 Þ ¼<br />

ρðP 2 Þ ¼ 1<br />

RP ð 2 Þ<br />

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

ðx 1 x 2 Þ 2 þðy 1 y 2 Þ 2 ðx 2 x 3 Þ 2 þðy 2 y 3 Þ 2 ðx 3 x 1 Þ 2 þðy 3 y 1 Þ 2<br />

2ðx 2 y 1 x 1 y 2 þ x 3 y 2 x 2 y 3 þ x 1 y 3 x 3 y 1 Þ<br />

ð17Þ<br />

ð18Þ<br />

where R(P 2 ) is the radius at point P 2 (x 2, y 2 ), having previous point P 1 (x 1 , y 1 )<br />

<strong>and</strong> next P 3 (x 3 , y 3 ). Except first <strong>and</strong> last point, radius <strong>and</strong> curvature of all points<br />

can be calculated by using Eq. (17) <strong>and</strong> (18). Equal interval of all the points is<br />

very essential for this method, thus additional or fewer points were generated<br />

from available FE nodes, experimental traces <strong>and</strong> A. aurita profile.<br />

Figures 18(a) <strong>and</strong> 18(b) compare the curvature of the profile along the<br />

length of the curved exumbrella surface in relaxed <strong>and</strong> contracted condition between<br />

FE simulation, experimental <strong>and</strong> natural A. aurita. In relaxed condition,<br />

FEM curvature matches that of A. aurita. Experimentally as evident from Figure<br />

16(a) <strong>and</strong> Figure 18(a) it does not match A. aurita profile in the relaxed state.<br />

Figure 18(b) emphasizes that curvatures of FEM, experimental as well as natural<br />

animal, increases as the fish contracts. Artificial A. aurita does not have the same<br />

relaxed state as the natural animal, but in contracted state it follows natural<br />

A. aurita exumbrella deformation. It outperforms the curvature of the natural<br />

A. aurita by achieving a maximum curvature of 67 m −1 at about 80% length.<br />

FEM <strong>and</strong> natural A. aurita achieve maximum curvature of 50 m −1 at 67% <strong>and</strong><br />

45 m −1 at 85%, respectively, <strong>and</strong> show sharp increase at the tip, suggesting dynamic<br />

overshoot of the passive flap that is hypothesized to help swimming performance.<br />

Figure 18(c) displays profile errors between various profiles <strong>and</strong><br />

conditions,<br />

ɛ FER ðÞ¼ s<br />

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

ðÞ s x ExpBR ðÞ s<br />

2<br />

þ yFEBR ðÞ s y ExpBR ðÞ s<br />

2<br />

x FEBR<br />

ð19Þ<br />

The profile error ɛ FER is defined as the<br />

distance between corresponding<br />

points (x(s), y(s)) (subscript BR corresponds<br />

to the BISMAC location <strong>and</strong><br />

relaxed condition) of FEM <strong>and</strong> the artificial<br />

A. aurita vehicle relaxed profiles<br />

at location s. This was calculated by<br />

Eq. (19), <strong>and</strong> we observe that profile<br />

errors between the FE simulation <strong>and</strong><br />

artificial A. aurita experiment in relaxed<br />

<strong>and</strong> contracted states, ɛ FER <strong>and</strong><br />

ɛ FEC , have similar trends (Figure 18(c)).<br />

This indicates that the difference between<br />

FE simulation <strong>and</strong> artificial<br />

A. aurita is mainly due to the initial difference<br />

in relaxed profiles. We hypothesize<br />

that if artificial A. aurita were<br />

re-designed to match the natual<br />

A. aurita’s relaxed state, the simulation<br />

would match it more closely. The profile<br />

error does not increase beyond<br />

5.8mm which corresponds to 22%<br />

of the initial profile error in relaxed<br />

configuration.<br />

The model developed in this study<br />

gives better underst<strong>and</strong>ing of the temperature<br />

rise in the SMA wires <strong>and</strong><br />

transient heat distribution in <strong>and</strong><br />

around it during entire cycle. It reveals<br />

that in spite of complex geometry, heat<br />

distribution is practically uniform<br />

along SMA length <strong>and</strong> can be effectively<br />

captured by 1st order unsteady heat<br />

conduction equation. It explains degradation<br />

of performance of artificial<br />

A. aurita, over number of cycle due<br />

to heat accumulation resulting from<br />

insufficient cooling <strong>and</strong> suggests<br />

opportunity of making the bioinspired<br />

A. aurita vehicle more energy<br />

efficient by reducing oversupply of<br />

heat to SMA wire <strong>and</strong> effectively controlling<br />

SMA temperature. The model<br />

predicts transient deformation behavior<br />

of the artificial A. aurita qualitatively<br />

at BISMAC <strong>and</strong> fold locations,<br />

confirming effectiveness of the join<br />

design. It reveals that the transient tip<br />

178 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 18<br />

Curvature comparison at BISMAC <strong>and</strong> fold location between ANSYS, experiment, <strong>and</strong> biological<br />

A. aurita (a) relaxed condition, (b) contracted condition, (c) profile errors at BISMAC location-<br />

FER: between FE <strong>and</strong> experiment relaxed profile, FEC:FE <strong>and</strong> experiment contracted profile, AEC:<br />

A. aurita <strong>and</strong> experimental contracted profile, FAC:FE <strong>and</strong> A. aurita contracted profile.<br />

displacement response can be modeled<br />

as a second-order system; however, due<br />

to mismatch in initial profile, it predicts<br />

the radial tip displacement response<br />

∼44% of the experimental curve. The<br />

model captures these physical aspects<br />

of bell deformation accurately <strong>and</strong> is a<br />

useful tool to evaluate effectiveness of<br />

design changes on performance of the<br />

vehicle without building one.<br />

Conclusion<br />

We introduce a customization procedure<br />

for BISMAC actuators to<br />

be used as radial muscles in artificial<br />

A. aurita such that we can match<br />

the relaxed <strong>and</strong> contracted profiles of<br />

A. aurita at the BISMAC locations.<br />

By accurately modeling the hyperelastic<br />

behavior of EcoFlex (Shore<br />

00-10) RTV silicone <strong>and</strong> using experimentally<br />

obtained temperature-strain<br />

transformation curves for SMA wires,<br />

we provide a high-fidelity model that<br />

captures most essential physics of<br />

artificial A. aurita deformation. The<br />

model suggests a unique approach<br />

to overcome ANSYS limitation in<br />

modeling SMA temperature transformation<br />

by using negative thermal coefficient<br />

of expansion <strong>and</strong> two separate<br />

material curves for heating <strong>and</strong> cooling.<br />

This approach is generic enough<br />

to be used in numerical modeling<br />

of the SMA temperature-dependent<br />

transformation in any design. The<br />

transient heat transfer model provides<br />

better underst<strong>and</strong>ing of temperature<br />

rise in SMA wire <strong>and</strong> heat distribution<br />

around it during an entire contractionrelaxation<br />

cycle. This information is<br />

crucial in designing <strong>and</strong> evaluating<br />

SMA heating algorithm <strong>and</strong> cannot<br />

be obtained from the prototype directly.<br />

The model also reveals that in<br />

spite of the complex geometry, the<br />

temperature distribution along the<br />

length of SMA wires is uniform <strong>and</strong><br />

we could use a simple first order heat<br />

transfer model to study temperature<br />

distribution in <strong>and</strong> around SMA<br />

wires. It also exposes the reason for<br />

degradation of the performance of artificial<br />

A. aurita over a number of cycles<br />

due to heat accumulation <strong>and</strong> reveals<br />

an opportunity to make the artificial<br />

A. aurita more energy-efficient by reducing<br />

the amount of heat supplied<br />

above the austenite’s finish temperature<br />

A f that does not contribute toward<br />

bell deformation. Transient structural<br />

model accounts for the gravity <strong>and</strong><br />

the buoyancy forces <strong>and</strong> thus in equilibrium<br />

conditions (fully contracted<br />

<strong>and</strong> fully relaxed states), where fluid<br />

pressure differentials across subumbrella<br />

<strong>and</strong> exumbrella do not<br />

exist, we can predict the bell deformations<br />

fairly accurately. Transient<br />

structural analysis captures all essential<br />

behavior of artificial A. aurita at<br />

BISMAC <strong>and</strong> fold locations <strong>and</strong> confirms<br />

effectiveness of the joint design,<br />

though an exact match was not<br />

achieved due to initial profile mismatch.<br />

The model reveals that the<br />

transient tip displacement response<br />

can be modeled as a second-order system;<br />

however, due to initial profile<br />

mismatch; the model predicts the radial<br />

tip displacement response at<br />

∼44% of the experimental curve. Profile<br />

error analysis suggests that initial<br />

error in simulation <strong>and</strong> artificial<br />

A. aurita relaxed profiles is a major<br />

cause for the mismatch. The profile<br />

error increases by a small amount<br />

with an average of 0.00015 m (0.58%<br />

of ɛ FER ) <strong>and</strong> reaches a maximum of<br />

0.0058 m (22% of ɛ FER ). The model<br />

captures these physical aspects of bell<br />

deformation physics accurately <strong>and</strong><br />

isausefultooltoevaluateeffectiveness<br />

of design changes on performance<br />

of the vehicle without the need for<br />

July/August 2011 Volume 45 Number 4 179


time-consuming physical construction.<br />

The modeling methodology is<br />

generic <strong>and</strong> could be used to model<br />

other bio-inspired robots using similar<br />

construction technique.<br />

Acknowledgment<br />

This research is sponsored by the<br />

Office of Naval Research through contract<br />

N00014-08-1-0654.<br />

Lead Authors:<br />

Keyur B. Joshi <strong>and</strong> Shashank Priya<br />

Center for Energy Harvesting<br />

Materials <strong>and</strong> Systems<br />

Center for Intelligent Material<br />

Systems <strong>and</strong> Structure<br />

Virginia Polytechnic Institute<br />

<strong>and</strong> State University<br />

310 Durham Hall, Blacksburg,<br />

VA 24061<br />

Email: key4josh@vt.edu;<br />

spriya@mse.vt.edu<br />

References<br />

Albert, D.J. 2009. Aurelia labiata (Scyphozoa)<br />

jellyfish in Roscoe Bay: Their spatial distribution<br />

varies with population size <strong>and</strong> their<br />

behaviour changes with water depth. J Sea<br />

Res. 61(3):140-3. doi: 10.1016/j.seares.<br />

2008.11.001.<br />

Baker, E. 1972. Liquid cooling of microelectronic<br />

devices by free <strong>and</strong> forced convection.<br />

Microelectron Reliab. 11(2):213-22.<br />

doi: 10.1016/0026-2714(72)90704-4.<br />

Chapman, D.M. 1974. Cnidarian histology.<br />

In Coelenterate Biology, eds. Muscatine, L.,<br />

& Lenhoff, H.M., 2-92. New York:<br />

Academic Press.<br />

Colin, S.P., & Costello, J.H. 2002. Morphology,<br />

swimming performance <strong>and</strong> propulsive<br />

mode of six co-occurring hydromedusae.<br />

J Exp Biol. 205:427.<br />

Dabiri, J.P., Colin, S.P., & Costello, J.H.<br />

2005. Flow patterns generated by oblate<br />

medusan jellyfish: Field measurements <strong>and</strong><br />

laboratory analyses. J Exp Biol. 208(7):1257.<br />

doi: 10.1242/jeb.01519.<br />

Gladfelter, W.G. 1972. Structure <strong>and</strong> function<br />

of the locomotory system of Polyorchis montereyensis<br />

(Cnidaria, Hydrozoa). Helgol<strong>and</strong> Wiss<br />

Meer. 23(1):38. doi: 10.1007/BF01616310.<br />

Gladfelter, W.G. 1973. Structure <strong>and</strong> function<br />

of the locomotory system of the Scyphomedusa<br />

Cyanea capillata. Mar Biol. 14(2):150.<br />

doi: 10.1007/BF00373214.<br />

ImageJ. Image Processing <strong>and</strong> Analysis in<br />

Java. http://rsbweb.nih.gov/ij/. (accessed 12<br />

July 2011).<br />

National Instruments. LabView. http://<br />

www.ni.com/labview/. (accessed 12 July 2011).<br />

Smith, C., Villanueva, A., Joshi, K., Tadesse,<br />

Y., & Priya, S. 2011. Working principle of<br />

bio-inspired shape memory alloy composite<br />

actuators. Smart Mater Struct. 20(1):012001.<br />

doi: 10.1088/0964-1726/20/1/012001.<br />

Smith, C.F., & Priya, S. 2010. Bio-inspired<br />

Unmanned Undersea Vehicle. Behavior <strong>and</strong><br />

Mechanics of Multifunctional Materials <strong>and</strong><br />

Composites 2010, March 8, 2010 - March<br />

11, 2010, The <strong>Society</strong> of Photo-Optical<br />

Instrumentation Engineers (SPIE); American<br />

<strong>Society</strong> of Mechanical Engineers. San Diego,<br />

CA: SPIE.<br />

Tadesse, Y., Brennan, J., Smith, C., Long,<br />

T.E., & Priya, S. 2010a. Synthesis <strong>and</strong><br />

characterization of polypyrrole composite<br />

actuator for jellyfish unmanned undersea<br />

vehicle. 764222-11. San Diego, CA: SPIE.<br />

Tadesse, Y., Thayer, N., & Priya, S. 2010b.<br />

Tailoring the response time of shape memory<br />

alloy wires through active cooling <strong>and</strong> prestress.<br />

J Intel Mat Syst Str. 21(1):19-40.<br />

doi: 10.1177/1045389X09352814.<br />

Toki Corp. Biometal Fiber. http://www.toki.<br />

co.jp/biometal/english/q<strong>and</strong>a.php. (accessed<br />

12 July 2011).<br />

Valentine, J.W. 2004. On the Origin of<br />

Phyla. Chicago: University of Chicago Press.<br />

Villanueva, A., Bresser, S., Chung, S.,<br />

Tadesse, Y., & Priya, S. 2009. Jellyfish<br />

inspired underwater unmanned vehicle.<br />

Proc SPIE. 7287:72871G. doi: 10.1117/<br />

12.815754.<br />

Villanueva, A., & Priya, S. 2010. BISMAC<br />

control using SMA resistance feedback.<br />

In: SPIE Conference Proceedings, SPIE<br />

NDE Conference. 76421Z-12. San Diego, CA:<br />

SPIE.<br />

Villanueva, A., Priya, S., Anna, C., & Smith,<br />

C. 2010a. Robojelly bell kinematics <strong>and</strong><br />

resistance feedback control. In: IEEE<br />

International Conference on Robotics <strong>and</strong><br />

<strong>Biomimetics</strong> (ROBIO). 1124-9. Tianjin,<br />

China: IEEE (Institute of Electrical <strong>and</strong><br />

Electronics Engineers).<br />

Villanueva, A., Smith, C., & Priya, S.<br />

submitted. Under review, Paper Ref no.<br />

“BB/378339/PAP/255146”. Biomimetic robotic<br />

jellyfish (Robojelly) using shape memory<br />

alloy. IOP Bioinspir Biomim.<br />

Villanueva, A.A., Joshi, K.B., Blottman, J.B.,<br />

& Priya, S. 2010b. A bio-inspired shape<br />

memory alloy composite (BISMAC) actuator.<br />

Smart Mater Struct. 19(2):025013.<br />

doi: 10.1088/0964-1726/19/2/025013.<br />

Yang, Y., Ye, X., & Guo, S. 2007. A new<br />

type of jellyfish-like microrobot. In: IEEE<br />

International Conference on Integration<br />

<strong>Technology</strong>. 673-8. Rourkela, India: IEEE<br />

(Institute of Electrical <strong>and</strong> Electronics<br />

Engineers). doi: 10.1109/ICITECHNOLOGY.<br />

2007.4290404.<br />

Yeom, S.-W., & Oh, I.-K. 2009. A biomimetic<br />

jellyfish robot based on ionic polymer<br />

metal composite actuators. Smart Mater<br />

Struct. 18(8):085002. doi: 10.1088/0964-<br />

1726/18/8/085002.<br />

180 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


PAPER<br />

Swimming <strong>and</strong> Walking of an Amphibious<br />

Robot With Fin Actuators<br />

AUTHOR<br />

Naomi Kato<br />

Graduate School of Engineering,<br />

Osaka University<br />

Introduction<br />

I<br />

t has recently been clarified that<br />

natural coastline areas <strong>and</strong> tidal flats<br />

play an important role in preserving<br />

ocean environments. To protect<br />

coastal environments, regular monitoring<br />

of these areas is important.<br />

Monitoring has previously been<br />

done by humans on foot or using<br />

boats, but this work can be dangerous<br />

because of breaking waves <strong>and</strong> rip<br />

currents. Monitoring on foot is limited<br />

because it cannot be carried out<br />

in deep waters, while monitoring by<br />

boat is limited to the areas accessible<br />

by water. Automatic monitoring by<br />

an amphibious robot is therefore expected<br />

to eliminate the safety threats<br />

to human monitors <strong>and</strong> improve operational<br />

efficiency. However, if an<br />

amphibious robot moves by gaining<br />

traction with screws <strong>and</strong> caterpillars,<br />

it will not be able to move about in<br />

areas such as marshes <strong>and</strong> will damage<br />

the environments of the areas in<br />

which it moves. An environmentally<br />

friendly amphibious robot is thus<br />

needed.<br />

Several studies have reported the<br />

development of amphibious robots.<br />

An amphibious snake-like robot, the<br />

ACM-R5 (Yamada et al., 2005), can<br />

operate both on ground <strong>and</strong> in water<br />

by undulating its long body. The<br />

ACM-R5 uses special paddles <strong>and</strong><br />

ABSTRACT<br />

With the goal of automatic monitoring of environments along natural coastal<br />

areas <strong>and</strong> tidal flats, researchers designed <strong>and</strong> developed an amphibious robot<br />

equipped with fin actuators called “RT-I” that mimics the locomotion of both a tortoise<br />

<strong>and</strong> a sea turtle. Experiments were carried out using a forearm with 4 degrees<br />

of freedom, which can reproduce the walking motions of tortoises <strong>and</strong> sea turtles on<br />

s<strong>and</strong>, to evaluate the walking performances of a robotic tortoise <strong>and</strong> a robotic sea<br />

turtle. It was clarified that the arm for a robotic tortoise is more suitable for use on<br />

soil compared with the arm for a robotic sea turtle. The advantages of both sea<br />

turtles <strong>and</strong> tortoises were adopted in a robotic turtle, namely, the lift-based swimming<br />

mode sea turtles use <strong>and</strong> the quadrupedal locomotion tortoises use. The present<br />

amphibious robot consists of four main components: (i) leg units, (ii) a control<br />

unit pressure hull, (iii) a buoyancy adjusting device, <strong>and</strong> (iv) a fairing cover. To realize<br />

not only swimming motion with the combination of flapping, rowing, <strong>and</strong><br />

feathering, but also tortoise-like walking motion, three motors were set up at the<br />

acromioclavicular joint using a differential gear mechanism to independently produce<br />

the three types of motion, <strong>and</strong> one motor was set up to produce elbow joint<br />

motion. A buoyancy-adjusting device was installed to realize walking on l<strong>and</strong> <strong>and</strong> in<br />

water as well as swimming in shallow water. The swimming <strong>and</strong> walking performances<br />

of the amphibious robot in water were evaluated by measuring the forward<br />

swimming speed, backward swimming speed, speed of turning, <strong>and</strong> speed of descending<br />

vertically as the indexes of the maneuverability of the robotic turtle, <strong>and</strong><br />

the walking speed <strong>and</strong> propulsive efficiency with the crawl gait for various walking<br />

patterns in still water <strong>and</strong> in waves.<br />

Keywords: locomotion, tortoise, sea turtle, experiment<br />

wheelsmountedarounditsbodyto<br />

propel itself through water <strong>and</strong> over<br />

ground in a snake-like fashion, generating<br />

propulsive force that allows it to<br />

glide freely in the tangential direction.<br />

The biomimetic amphibious soft cord<br />

robot (Wakimoto et al., 2006), which<br />

is made of Mckibben actuators <strong>and</strong><br />

plastic plates, can move both on the<br />

ground <strong>and</strong> in water, undulating its<br />

long body. A spinal cord model <strong>and</strong><br />

its implementation in an amphibious<br />

salam<strong>and</strong>er robot (Ijspeert et al.,<br />

2007) were studied to demonstrate<br />

how a primitive neural circuit for<br />

swimming can be extended using<br />

phylogenetically more recent limb oscillatory<br />

centers to explain the ability<br />

of salam<strong>and</strong>ers to switch between<br />

swimming <strong>and</strong> walking. The AQUA<br />

(Dudek et al., 2007), an amphibious<br />

robot, can swim <strong>and</strong> walk along the<br />

shore <strong>and</strong> on the bottom of the ocean<br />

by moving its fins. The AQUA uses<br />

six paddles, which act as control surfaces<br />

during swimming <strong>and</strong> as legs<br />

while walking. An amphibious walking<br />

robot was developed by Tanaka<br />

<strong>and</strong> Shirai (2006) to perform a shoreline<br />

survey. It has six legs, each of<br />

July/August 2011 Volume 45 Number 4 181


which has three joints. It was successful<br />

in obtaining the distribution of<br />

ground levels from l<strong>and</strong> to shallow<br />

water. An amphibious robotic turtle<br />

was built by Low et al. (2007) to imitate<br />

the locomotion of Cheloniidae,<br />

both in water <strong>and</strong> on l<strong>and</strong>, to perform<br />

various operations. The crawling <strong>and</strong><br />

lift-based swimming gaits were analyzed<br />

<strong>and</strong> implemented in the prototype.<br />

However, all of these robots are<br />

not operated in practice to monitor<br />

the coastal environment on l<strong>and</strong> <strong>and</strong><br />

in water by using the multiple functions<br />

of walking <strong>and</strong> swimming.<br />

For the field operation of an amphibious<br />

robot, a rigid fuselage with<br />

an adequate payload is necessary so<br />

that a control system <strong>and</strong> sensors for<br />

monitoring the environment can be<br />

installed. We have been studying a biomimetic<br />

underwater robot equipped<br />

with mechanical pectoral fins from<br />

the viewpoint of high maneuverability<br />

under disturbances such as waves <strong>and</strong><br />

water currents (Kato et al., 2006;<br />

Suzuki & Kato, 2005; Kato & Liu,<br />

2003). Taking the field operation<br />

<strong>and</strong> application of our experiences on<br />

the biomimetic underwater robot<br />

into account, this study focuses on an<br />

amphibious robotic turtle, which not<br />

only can swim in the sea but also<br />

walk on the l<strong>and</strong> to perform environmental<br />

monitoring of natural coast<br />

<strong>and</strong> tidal flat areas.<br />

Turtles that can walk <strong>and</strong> swim are<br />

generally categorized intoseaturtles<br />

<strong>and</strong> tortoises. Sea turtles have good<br />

swimming ability but poor walking<br />

ability because they drag their bodies<br />

on the l<strong>and</strong>, which causes friction<br />

against the s<strong>and</strong>. Tortoises, on the<br />

other h<strong>and</strong>, cannot swim smoothly,<br />

but they can walk better than sea<br />

turtles can because of their quadrupedal<br />

locomotion capability. In this<br />

study, we attempted to adopt the advantages<br />

of these two turtles into a robotic<br />

turtle.<br />

This paper presents (1) a description<br />

of the walking performance of<br />

an arm with 4 degrees of freedom<br />

(DOF) that can reproduce the walking<br />

motionsofseaturtles<strong>and</strong>tortoises<br />

from the viewpoints of mobility <strong>and</strong><br />

terrain trafficability, (2) details of the<br />

design <strong>and</strong> development of an amphibious<br />

robot with fin actuators,<br />

<strong>and</strong> (3) an evaluation of the walking<br />

<strong>and</strong> swimming performance of the<br />

robot in a laboratory environment.<br />

Here, mobility is defined as the walking<br />

performance of a vehicle depending<br />

on motor torque <strong>and</strong> vehicle<br />

configuration. Trafficability is the<br />

soil-bearing capacity of a vehicle.<br />

Locomotion of Tortoises<br />

<strong>and</strong> Sea Turtles<br />

Terrestrial Locomotion<br />

Walker (1971) studied the walking<br />

of a tortoise, Chrysemys picta, using<br />

cinephotography <strong>and</strong> X-rays <strong>and</strong><br />

clarified the relationship between the<br />

structure of the skeleton <strong>and</strong> the movements<br />

of the fore <strong>and</strong> hind limbs. The<br />

structure of the skeleton of the forelimb<br />

is the same as that of the hind<br />

limb. The forelimb consists of an<br />

acromioclavicular joint with 3 DOF,<br />

ahumerus,anelbowwith1DOF,a<br />

FIGURE 1<br />

Top view <strong>and</strong> side view of tortoise Chinemys reevesii.<br />

forearm, a wrist with 2 DOF, <strong>and</strong> a<br />

h<strong>and</strong>. Wyneken (1997) explained<br />

that in the locomotion of sea turtles<br />

on l<strong>and</strong>, clutching movements are<br />

seen in adult Chelonia mydas, Natator<br />

depressus, <strong>and</strong>Dermochelys coriacea,<br />

while adults of other cheloniid species<br />

employ quadrupedal gaits in which diagonally<br />

opposite feet move as a pair.<br />

Sea turtles support themselves on the<br />

carpus <strong>and</strong> the anterior edge of the<br />

h<strong>and</strong> rather than on the palmar<br />

surface.<br />

To design a robotic turtle, we analyzed<br />

quantitative information on the<br />

movement of the forelimb joints of a<br />

captured tortoise, Chinemys reevesii,<br />

with the following dimensions: length<br />

of shell × length of forelimb × length<br />

of hindlimb = 230 mm × 40 mm ×<br />

40 mm. The markers <strong>and</strong> body-fixed<br />

coordinates (x, y, z) on the tortoise<br />

were set up as shown in Figure 1.<br />

The origin of the body-fixed coordinates<br />

was set on the acromioclavicular<br />

joint. First, movies of the walking<br />

motions of the tortoise were taken<br />

using two CCD cameras. The movies<br />

from the top view <strong>and</strong> side view were<br />

taken at the same time. Second, the<br />

marked points were tracked using<br />

software that computed the twodimensional<br />

coordinates of the points.<br />

Third, the three-dimensional coordinates<br />

of the elbow <strong>and</strong> wrist on<br />

the body-fixed coordinates were<br />

182 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Trajectories of the elbow <strong>and</strong> the wrist in x-y plane.<br />

computed. Figures 2 <strong>and</strong> 3 show the<br />

two-dimensional tracks in the x-y<br />

plane <strong>and</strong> x-z plane of the motion of<br />

the wrist <strong>and</strong> the elbow, respectively,<br />

during walking in the case of a walking<br />

speed of 0.13 m/s, a period of walking<br />

of 1.3 s, <strong>and</strong> a stance phase of 0.78.<br />

Here, the stance phase denotes the fraction<br />

of time during which the forelimb<br />

is set on the ground during the walking<br />

period.<br />

If the forelimb is considered to be<br />

an arm, the joint angles of the forelimb<br />

FIGURE 3<br />

Trajectories of the elbow <strong>and</strong> the wrist in x-z plane.<br />

derived from inverse kinematics of the<br />

arm can be obtained. Here, an arm is<br />

assumed to consist of an acromioclavicular<br />

joint with 3 DOF, a humerus, an<br />

elbow with 1 DOF, a forearm, <strong>and</strong> a<br />

wrist. The (x, y, z) coordinates are<br />

fixed at the acromioclavicular joint, as<br />

shown in Figure 4. The rowing motion<br />

is defined as rotational motion around<br />

the z axis. The (x′, y′, z′) coordinates<br />

are defined as coordinates rotated<br />

around the z axis by the rowing motion.<br />

The feathering motion is defined<br />

as rotational motion around the y′ axis.<br />

The (x″, y″, z″) coordinates are defined<br />

as the coordinates rotated around the<br />

y′ axis by the feathering motion. The<br />

flapping motion is defined as rotational<br />

motion around the x″ axis.<br />

The coordinate system of the arm<br />

was set up as shown in Figure 4,<br />

where n, s, <strong>and</strong>a, denotingunitvectors<br />

fixed on the forearm, were set parallel<br />

to x, y, <strong>and</strong> z, respectively, when<br />

all of the joint angles were zero. The<br />

joint angles θ 1 , θ 2 , θ 3 , θ 4 are defined<br />

as the angle of rowing motion, angle<br />

of feathering motion, angle of flapping<br />

motion, <strong>and</strong> angle of bending of forearm,<br />

respectively. Lengths of the arms<br />

l 1 , l 2 are defined as the length of the<br />

humerus <strong>and</strong> the length of the forearm,<br />

respectively.<br />

Figure 5 shows time variations of<br />

thejointangles.Wecanseethatthe<br />

feathering angle varies from 5° to 49°<br />

during the power stroke from 0 to<br />

0.6 s, <strong>and</strong> the bending angle of the<br />

forelimb varies between −75° <strong>and</strong><br />

−115° during the power stroke,<br />

which indicates that the forelimb produces<br />

forward thrust by kicking the<br />

ground <strong>and</strong> positioning the forearm almost<br />

vertically on the ground. During<br />

the recovery stroke from 0.6 to 1.3 s,<br />

the bending angle of the forelimb<br />

reaches −20°, which indicates that the<br />

forearm is raised from the ground. The<br />

flapping angle does not vary much<br />

during the entire stroke.<br />

Aquatic Locomotion<br />

Sea turtles swim in water using<br />

their forelimbs to provide thrust. The<br />

synchronous sweeping of the flippers<br />

introduces a stable heading direction.<br />

Wyneken (1997) explains lift-based<br />

mechanisms of thrust production in<br />

which the locomotor apparatus of a<br />

sea turtle acts as a wing to generate<br />

July/August 2011 Volume 45 Number 4 183


FIGURE 4<br />

Coordinate system <strong>and</strong> definitions.<br />

minimize surface area as they are<br />

brought forward.<br />

Experiment on the<br />

Walking Performance<br />

of an Arm<br />

We constructed an arm that could<br />

reproduce the walking motion of a tortoise<br />

<strong>and</strong> a sea turtle to analyze the<br />

walking performance from the viewpoints<br />

of mobility <strong>and</strong> trafficability.<br />

lift forces during large portions of the<br />

powerstroke. Isobe et al. (2010) clarified<br />

that the feathering motion of<br />

fore flippers of a sea turtle influences<br />

the thrust production by measuring<br />

the 3-D motion of fore flippers of a<br />

sea turtle in a water circulating tank<br />

<strong>and</strong> conducting numerical simulations<br />

FIGURE 5<br />

Time variations of joint angles.<br />

based on quasi-steady wing element<br />

theory.<br />

Wyneken (1997) showed dragbased<br />

propulsion by a swimming semiaquatic<br />

turtle. Diagonally opposite<br />

limbs are protracted, then retracted together,<br />

<strong>and</strong> act as paddles. The distal<br />

elements of the limbs are flexed to<br />

Arm<br />

The robotic arm we developed<br />

makes the motions of rowing, feathering,<br />

<strong>and</strong> flapping, <strong>and</strong> the forearm can<br />

also be bent. The arm moves on rails to<br />

make these motions on s<strong>and</strong>. The angles<br />

of the motions of rowing, feathering,<br />

flapping, <strong>and</strong> bending were<br />

measured by four potentiometers, the<br />

distance of the movement along the<br />

rails was measured by a potentiometer,<br />

<strong>and</strong> the forces that were applied to the<br />

h<strong>and</strong> were measured by a six-axes force<br />

sensor, as shown in Figure 6. The humerus<br />

<strong>and</strong> forearm are each 150 mm<br />

long. The rowing angle, feathering<br />

angle, flapping angle, <strong>and</strong> bending<br />

angle of the forearm vary within the<br />

following ranges, respectively: ±70°,<br />

FIGURE 6<br />

Picture of manipulator.<br />

184 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


±70°, ±50°, <strong>and</strong> 0° to −110°. The rowing motion, feathering motion, flapping motion, <strong>and</strong> bending motion were produced by<br />

four motors independently. Because frictional force works between the rails <strong>and</strong> the arm, a force corresponding to the static<br />

frictional force was applied to the arm along the rail using a pulley <strong>and</strong> a weight. In the experiments on the walking performance<br />

of a sea turtle, an amount of force was subtracted from the static frictional force to simulate the friction a sea turtle<br />

generates on s<strong>and</strong>. The arm was connected to the weight by a string. Toyoura s<strong>and</strong>, which has almost constant particle<br />

diameter <strong>and</strong> known physical parameters, was used in the experiments.<br />

Figure 7 shows the h<strong>and</strong> shapes. The shape of the end of the h<strong>and</strong> of the model simulating a tortoise (T ) is a 70 mm ×<br />

70 mm square. That of the model simulating a sea turtle (S) is a 150 mm × 125 mm rectangle.<br />

Kinetic Relations Between External Force <strong>and</strong> Joint Torque<br />

To discuss the propulsive efficiency in terms of walking performance of the arm, we need to know the torque of each joint<br />

of the arm. We then consider the kinetic relations between external force <strong>and</strong> joint torque.<br />

First, we refer to the Jacobian matrix in the kinetics of the arm. We define θ(θ 1 , θ 2 , θ 3 , θ 4 ) T as the rotational angles of<br />

joints. The relations between the rotational velocities of the joints, θ : θ : 1 ; θ : 2 ; θ : 3 ; θ : T 4 , translational velocity of the end of the<br />

forearm (see Figure 6), ( P : : : :<br />

r ðx;<br />

y; zÞ<br />

T ), <strong>and</strong> rotational velocities of the end of the forearm, ( Φ : r ð ϕ : x ; ϕ : y ; ϕ : z Þ T ), are described as<br />

follows:<br />

: <br />

P : r<br />

¼ J θ<br />

: Φ r<br />

ð1Þ<br />

<br />

J ¼ s <br />

1 × ðP r P 1 Þ s 2 × ðP r P 2 Þ s 3 × ðP r P 3 Þ s 4 × ðP r P 4 Þ<br />

s 1 s 2 s 3 s 4<br />

ð2Þ<br />

where J, P i ,<strong>and</strong>s i denote the Jacobian matrix, position vector, <strong>and</strong> vector of rotational direction, respectively (refer to<br />

Figure 8).<br />

Next, we will explain the kinetic relations between the external force, F(F x , F y , F z ) T , moment, M(M x , M y ,M z ) T , <strong>and</strong><br />

joint torque, Q(q 1 , q 2 , q 3 , q 4 ). The moment, M i , that is applied to each joint is<br />

M i ¼ ðP r P i Þ×Fþ M ð3Þ<br />

<strong>and</strong> thus the torque of each joint is<br />

q i ¼ s i ⋅ M i<br />

¼ s i ⋅ðP r<br />

P i<br />

Þ× F þ s i ⋅M<br />

¼ fs i ×P ð r P i Þg⋅F þ s i ⋅M<br />

ð4Þ<br />

<strong>and</strong> the relation between eternal forces, moments, <strong>and</strong> joint torque is represented using a transposed Jacobian matrix as<br />

follows:<br />

2 3 2<br />

3<br />

q 1 s 1 × ðP r P 1 Þ s 1 <br />

q<br />

Q ¼ 6 2<br />

7<br />

4 q 3<br />

5 ¼ s 2 × ðP r P 2 Þ s 2<br />

6<br />

7<br />

4 s 3 × ðP r P 3 Þ s 3<br />

5 F <br />

¼ J T F<br />

M M<br />

q 4 s 4 × ðP r P 4 Þ s 4<br />

ð5Þ<br />

July/August 2011 Volume 45 Number 4 185


FIGURE 7<br />

H<strong>and</strong> models.<br />

In the present study, we discuss<br />

propulsive efficiency from the viewpoints<br />

of mobility <strong>and</strong> trafficability.<br />

We define propulsive efficiency,η, as<br />

follows:<br />

FIGURE 9<br />

Trajectories of wrist <strong>and</strong> elbow of type (a).<br />

motion of a robotic tortoise were configured<br />

with reference to the experimental<br />

motion analysis of a tortoise<br />

in locomotion of tortoises <strong>and</strong> sea<br />

turtles. Type (a) motion includes<br />

movement of the arm vertically during<br />

the power stroke. It has a symmetric<br />

trajectory in the anteroposterior direction<br />

(see Figure 9). Type (b) motion is<br />

based on the pulling motion during<br />

the power stroke. It has a larger anterior<br />

trajectory component within the<br />

whole trajectory, which is analogous<br />

to the observed trajectory of the tortoise.<br />

Type (c) motion is based on<br />

the kicking motion during the power<br />

stroke. It has a larger posterior trajectory<br />

component within the whole trajectory.<br />

Figures 10, 11, <strong>and</strong> 12 show<br />

the time variations of the joint angles<br />

of type (a), type (b), <strong>and</strong> type (c), respectively.<br />

The range of movement of<br />

η ¼<br />

∫ T F x vdt<br />

4 :<br />

∑ ∫ T q i θi dt<br />

i¼1<br />

ð6Þ<br />

Here, v denotes the translational speed<br />

along the x axis. The numerator is the<br />

value of the work that the arm applies<br />

to the s<strong>and</strong>. The denominator is the<br />

value of the total input work by the<br />

joints.<br />

Comparison of Walking<br />

Performance of an Arm<br />

Between a Robotic Tortoise<br />

<strong>and</strong> a Robotic Sea Turtle<br />

The walking performance of an<br />

arm for a robotic tortoise using the<br />

h<strong>and</strong> models (T ) was measured. The<br />

following three types of walking<br />

FIGURE 8<br />

Vectors of coordinates.<br />

186 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 10<br />

Time variations of joint angles of type (a).<br />

FIGURE 12<br />

Time variations of joint angles of type (c).<br />

FIGURE 11<br />

Time variations of joint angles of type (b).<br />

the wrist along the x axis <strong>and</strong> that along<br />

the z axis were taken as 162 <strong>and</strong><br />

70 mm, respectively. The range of<br />

movement of the wrist along the<br />

y axis was taken as 103 mm. The period<br />

of motion was changed from 7.5<br />

to 15.0 s with an interval of 2.5 s.<br />

The sinkage of the end of the forearm<br />

below the surface of the s<strong>and</strong> was also<br />

varied; values of 10, 15, <strong>and</strong> 20 mm<br />

were used. Although the sinkage of a<br />

robotic turtle changes with time according<br />

to the weight of the body,<br />

the type of arm motion <strong>and</strong> the condition<br />

of the soil, it was treated as an independent<br />

parameter affecting the<br />

walking performance of the arm in<br />

this study so we could evaluate it together<br />

with other parameters from<br />

the viewpoint of fundamental walking<br />

performance. It was controlled by<br />

a feedback controller located in the<br />

arm controller using the measured angles<br />

of the motions of rowing, feathering,<br />

flapping, <strong>and</strong> bending by four<br />

potentiometers, after the initial attitude<br />

of the arm was set up in relation<br />

to the surface of the s<strong>and</strong>.<br />

Figure 13 shows the averaged propulsive<br />

forces, Fx, in the direction of<br />

the x axis during one period for the<br />

three types of arm motion with the<br />

motion period of 10 s. Figure 14<br />

shows the averaged vertical forces, Fz,<br />

in the direction of the z axis during one<br />

period for the three types of arm motion<br />

with the motion period of 10 s.<br />

The fact that the averaged vertical<br />

forces, Fz, for type (b) motion are negative<br />

can be attributed to the posture of<br />

the forearm in the first half of the time<br />

period. Namely, the flat plate at the<br />

end of the forearm digs into the s<strong>and</strong><br />

by positioning the forearm anteriorly<br />

in the firsthalfofthetimeperiod.<br />

This action leads it to carry s<strong>and</strong> on<br />

the flat plate, which causes a negative<br />

vertical force during one cycle. On<br />

the other h<strong>and</strong>, the forearm in type<br />

(a) motion is positioned almost vertically<br />

on the s<strong>and</strong> during one period,<br />

<strong>and</strong> the forearm in type (c) motion is<br />

positioned posteriorly so as to kick<br />

s<strong>and</strong>. Although it is difficult to discriminate<br />

among the three motion<br />

types in terms of the propulsive efficiencies<br />

beyond the sinkage of 15 mm,<br />

type (a) shows the largest values for<br />

the averaged propulsive force. The<br />

July/August 2011 Volume 45 Number 4 187


FIGURE 13<br />

Averaged propulsive force Fx in the direction of the x-axis during one<br />

period for the three types of arm motions against sinkage of the end<br />

of the forearm below the surface of the s<strong>and</strong> using the h<strong>and</strong> model (T).<br />

FIGURE 15<br />

Time variations of joint angles of type (d).<br />

same relation among the three motion<br />

types can be applied to the averaged<br />

vertical forces. Type (a) arm motion is<br />

suitable not only from the viewpoint<br />

of mobility, but also trafficability.<br />

Because sea turtles use the anterior<br />

edge of the h<strong>and</strong> for locomotion on<br />

l<strong>and</strong>, we used type (d) of walking<br />

FIGURE 14<br />

motion, in which the flipper moves inclined<br />

at an angle of 45°, in the design<br />

of the robotic sea turtle. Figure 15<br />

shows the time variations of the joint<br />

angles of type (d) with the motion period<br />

of 10 s. The sinkage of the bottom<br />

of the flipper below the surface of the<br />

s<strong>and</strong> was varied among three different<br />

Averaged vertical force Fz in the direction of the z-axis during one period for the three types of arm<br />

motions against sinkage of the end of the forearm below the surface of the s<strong>and</strong> using the h<strong>and</strong><br />

model (T).<br />

levels:10mm,15mm,<strong>and</strong>20mm.<br />

The walking motion of a sea turtle generates<br />

friction on the s<strong>and</strong> because the<br />

turtle drags its body on the l<strong>and</strong> as it<br />

walks. To model this situation, friction<br />

on the carriage of the manipulator was<br />

added. We compared the walking performance<br />

of these two types of motion<br />

under the conditions including friction<br />

on the carriage of the manipulator<br />

set at the following levels: 2.94, 4.90,<br />

<strong>and</strong> 6.86 N.<br />

Figure 16 shows the averaged propulsive<br />

force, Fx, in the direction of the<br />

x axis during one period against sinkage<br />

of the end of the forearm below<br />

the surface of the s<strong>and</strong> <strong>and</strong> friction<br />

on the carriage of the manipulator<br />

(the key to the symbols expresses friction<br />

on the carriage of the manipulator).<br />

As the friction on the rail increases, the<br />

averaged propulsive force, Fx, also increases.<br />

Figure 17 shows the average<br />

vertical force during one period against<br />

sinkage. As the sinkage increases, the<br />

maximum vertical force, Fz, also<br />

increases. The dependency of the<br />

maximum vertical force, Fz, on the<br />

188 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 16<br />

Averaged propulsive force Fx in the direction of x-axis during one period<br />

for type (d) arm motion against sinkage of the end of the forearm<br />

below the surface of s<strong>and</strong> using the h<strong>and</strong> model (S).<br />

FIGURE 17<br />

Averaged vertical force Fz in the direction of z-axis during one period<br />

for type (d) of arm motion against sinkage of the end of the forearm<br />

from the surface of s<strong>and</strong> using the h<strong>and</strong> model (S).<br />

friction on the rail is clear. When we<br />

consider the walking motion of a robotic<br />

sea turtle on the beach, friction<br />

between the s<strong>and</strong> <strong>and</strong> the body depends<br />

on the vertical force subtracting<br />

the vertically upward force from the<br />

weight of the robotic sea turtle.<br />

When comparing type (a) walking<br />

motion of a robotic tortoise, which<br />

showed the best performance among<br />

the three types, with type (d) walking<br />

motion of a robotic sea turtle, it can be<br />

seen that the propulsive efficiency of<br />

type (d) is influenced by friction <strong>and</strong><br />

the averaged vertical force acting on<br />

the h<strong>and</strong> plate in the case of the arm<br />

for a robotic tortoise is larger than in<br />

the case of the arm for a robotic sea turtle.<br />

An arm for a robotic tortoise is suitable<br />

for moving heavy payloads over a<br />

variety of terrains. To discuss the trafficability<br />

of the arm for a sea turtle, we<br />

define M as the mass of the body, g as<br />

gravitational acceleration, k as the frictional<br />

coefficient (Setouchi & Shinjo,<br />

2001) between the soil <strong>and</strong> the ventral<br />

surface of the robotic sea turtle, Fxm as<br />

the mean value of propulsive force, <strong>and</strong><br />

Fzm as the mean value of vertical force<br />

against the soil. Assuming that a pair of<br />

fore flippers <strong>and</strong> a pair of rear flippers<br />

exert thrust forces simultaneously, we<br />

obtain the following equation of static<br />

equilibrium for walking on soil:<br />

ðM⋅g − 4⋅FzmÞ⋅k ¼ 4⋅Fxm ð7Þ<br />

IfweassumeFxm=Fzmforthe<br />

type (d) arm motion based on Figures<br />

16 <strong>and</strong> 17, we obtain the following<br />

relation:<br />

k<br />

Fzm ¼ M⋅g⋅ < M⋅g=4 ð8Þ<br />

41þ ð kÞ If we set the length of the humerus as l 1<br />

<strong>and</strong> the length between the elbow joint<br />

<strong>and</strong> center of the h<strong>and</strong> plate as l 2 using<br />

the arm structure shown in Figures 6<br />

<strong>and</strong> 8, the torque around the x axis<br />

<strong>and</strong> that around the z axis are Fzm∙<br />

(l 1 + l 2 ). For a robotic tortoise using a<br />

crawling gait with slow speed, we obtain<br />

Fzm = M · g/4 during the time when<br />

the 4 feet are placed on ground. If we<br />

assume that the robotic turtle uses<br />

type (a) arm motion <strong>and</strong> that Fxm is<br />

nearly equal to Fzm/25 based on Figures<br />

13 <strong>and</strong> 14, we obtain the torque<br />

around the x axis of Fzm∙l 1 <strong>and</strong> a<br />

torque around the z axis of Fzm · l 1 /25.<br />

The total torque on the arm of the<br />

sea turtle robot, Q S , is expressed as<br />

follows:<br />

k<br />

Q S ¼ 2⋅M⋅g⋅<br />

41þ ð kÞ ⋅ ð l 1 þ l 2 Þ: ð9Þ<br />

The total torque on the arm of the<br />

tortoise robot, Q L , is expressed as<br />

follows:<br />

Q L ¼ 26=25⋅M⋅g=4⋅l 1<br />

The ratio of Q S to Q L γ is<br />

γ ¼ 25 <br />

13 ⋅ k<br />

1 þ k ⋅ 1 þ l2 <br />

l1<br />

ð10Þ<br />

ð11Þ<br />

Because the friction ratio of s<strong>and</strong><br />

(Setouchi & Shinjo, 2001) is in the<br />

July/August 2011 Volume 45 Number 4 189


ange of 0.4-0.6 <strong>and</strong> l 2 /l 1 is larger than<br />

1.0 for a sea turtle, we find that γ tends<br />

to approach 1.0 if k becomes smaller,<br />

<strong>and</strong> that γ tends to become larger<br />

than 1.0 if k becomes larger. Because<br />

smaller torques acting around the<br />

joints are needed from the viewpoint<br />

of the mechanical design of a<br />

turtle robot, the arm for the robotic<br />

tortoise is suitable on soil from the<br />

viewpoint of trafficability.<br />

Design <strong>and</strong> Development<br />

of an Amphibious Robot<br />

With Fin Actuators<br />

A previously designed mechanical<br />

pectoral fin (Kato & Liu, 2003) has a<br />

drag-based swimming mode <strong>and</strong> a liftbased<br />

swimming mode. The former is<br />

characterized by the rowing action<br />

forming a high angle to the horizontal<br />

axis of the body, while the latter is<br />

characterized by the flapping action<br />

forming a small angle to the horizontal<br />

axis. It was revealed through the optimization<br />

of motion of the mechanical<br />

pectoral fin that the lift-based swimming<br />

mode rather than the dragbased<br />

swimming mode is suitable for<br />

generation of propulsive force in uniform<br />

flow, while the drag-based swimming<br />

mode rather than the lift-based<br />

swimming mode is suitable for generation<br />

of propulsive force in still water.<br />

To realize both the drag-based swimming<br />

mode <strong>and</strong> the lift-based swimming<br />

mode, a combination of rowing<br />

motion, flapping motion, <strong>and</strong> feathering<br />

motion is needed. The hydrodynamic<br />

characteristics of the drag-based<br />

swimming mode <strong>and</strong> the lift-based<br />

swimming mode are discussed in details<br />

elsewhere (Suzuki et al., 2007).<br />

Walking locomotion using fin actuators<br />

is of two types, imitating the motionofatortoise<strong>and</strong>thatofasea<br />

turtle. Sea turtle-like walking motion<br />

has the following characteristics:<br />

It is dynamically stable.<br />

Terrain condition strongly affects<br />

the attitude of the body.<br />

Friction between the soil <strong>and</strong> the<br />

body necessitates additional thrust<br />

force.<br />

Tortoise-like walking motion has the<br />

following characteristics:<br />

It is dynamically more unstable<br />

compared with the sea turtle-like<br />

walking motion.<br />

The attitude of the body has more<br />

DOF compared with the sea turtlelike<br />

walking motion.<br />

The body weight is vertically applied<br />

to the foot.<br />

In this study, we adopted the<br />

tortoise-like walking motion, which<br />

has better trafficability on soil than<br />

the sea turtle-like walking motion, as<br />

discussed in the previous section.<br />

FIGURE 18<br />

Components of Robotic Turtle (RT-I) <strong>and</strong> fin actuator with 4 DOF.<br />

The amphibious robot we designed<br />

consists of the following four main<br />

components (see Figure 18):<br />

(1) leg units,<br />

(2) a control unit in a pressure hull,<br />

(3) a buoyancy adjusting device in a<br />

pressure hull, <strong>and</strong><br />

(4) a fairing cover<br />

Each pressure hull was designed to resist<br />

the pressure at the water depth of<br />

10 m. We used “Solidworks” for 3-D<br />

CAD software to design the hardware<br />

<strong>and</strong> “LabVIEW” to develop<br />

the control software. Four DOFs are<br />

needed to realize not only three<br />

types of swimming motion but also<br />

the tortoise-like walking motion.<br />

Therefore, three motors were set up<br />

at the acromioclavicular joint using a<br />

differential gear mechanism to independently<br />

produce flapping motion,<br />

rowing motion, <strong>and</strong> feathering motion,<br />

<strong>and</strong> one motor was set up at<br />

190 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


the elbow joint to produce bending<br />

motion of the forearm (see Figure 18).<br />

Motors <strong>and</strong> reduction gears were<br />

selected according to the simulation<br />

results on walking. An open dynamic<br />

engine of the kinetics calculation library<br />

with open sources was used for<br />

the simulation. A fuselage with the dimensions<br />

(W × H × L = 0.50 × 0.20 ×<br />

0.80 m) was used. The length of the<br />

humerus was set at 0.25 m, <strong>and</strong> the<br />

length of the forearm was set at<br />

0.15 m. There are various walking<br />

gaits for quadrupedal locomotion.<br />

We selected a crawling gait for low<br />

speed where the legs are lifted up one<br />

by one. The fin actuators of this<br />

robot have two functions, walking <strong>and</strong><br />

swimming, so we had to design the<br />

shape of the fin withbothwalking<br />

<strong>and</strong> swimming in mind. For swimming,<br />

it is desirable to have a large<br />

fin area. However, the fin mayinterfere<br />

with the base of the leg unit<br />

during walking. The fin shapeisdesigned<br />

to minimize the interference.<br />

Figure 19 shows the form of the fin<br />

with the root chord of 0.195 m, the<br />

span of 0.15 m, <strong>and</strong> the maximum<br />

thickness of 0.32 m.<br />

Figure 20 shows the outline of<br />

the control unit’s electric circuit. The<br />

control unit consists of a CPU,<br />

motor drivers, an azimuth sensor,<br />

FIGURE 19<br />

Form of fin.<br />

FIGURE 20<br />

Electric circuit of the control unit.<br />

three-axes rate gyros, a pressure sensor,<br />

a GPS, <strong>and</strong> related minor parts. Batteries<br />

are used separately for motors<br />

<strong>and</strong> the CPU with sensors. Nickel<br />

hydride batteries with the capacity of<br />

13.2 V, 8.0 Ah are used for motors<br />

by arranging a series circuit of two<br />

<strong>and</strong> a parallel circuit of two. It is also<br />

possible to provide the control unit<br />

with electric power from outside.<br />

The amphibious robot has to realize<br />

walking in which the buoyancy is<br />

less than the weight <strong>and</strong> swimming<br />

in which the buoyancy is equal to or<br />

greater than the weight, because the<br />

mission of the robot includes travel<br />

in shallow water with breaking waves.<br />

To realize these capabilities, the robot<br />

must be equipped with a buoyancy<br />

adjusting device. The buoyancy adjusting<br />

device with the buoyancy capacity<br />

of ±0.35 kg was designed using<br />

a pair of pistons arranged symmetrically<br />

in the longitudinal direction so<br />

as not to have an effect on the attitude<br />

of the robot.<br />

It is desirable to cover the body<br />

of the robot with streamlined fairing.<br />

The general-purpose computer fluid<br />

dynamics software “FLUENT” was<br />

used to estimate the hydrodynamic<br />

drag on the fairing cover <strong>and</strong> to design<br />

a fairing cover with low hydrodynamic<br />

drag. The cover was made of glass<br />

fiber-reinforced plastics.<br />

Figure 21 shows a photograph of<br />

the amphibious robot equipped with<br />

fin actuators, named “RT-I.” The<br />

principal dimensions are shown in<br />

Table 1.<br />

July/August 2011 Volume 45 Number 4 191


FIGURE 21<br />

Photograph of the RT-I amphibious robot<br />

equipped with fin actuators.<br />

TABLE 1<br />

Specification of RT-I.<br />

Total mass [kg] (without buoyancy control device <strong>and</strong> battery) 90<br />

Depth of pressure resistant [m] 10<br />

Walking speed in the water [m/s] (experimental value) 0.025<br />

Swimming speed [m/s] (experimental value) 0.168<br />

Dimension [m] Body width 0.73<br />

Body height 0.55<br />

Body length 1.68<br />

Forearm length 0.26<br />

Humerus length 0.2<br />

Swimming <strong>and</strong> Walking<br />

Performance by the<br />

Robotic Turtle RT-I<br />

Swimming Performance<br />

A towing tank test was carried out<br />

to measure drag forces acting on the<br />

body of the robotic turtle in the towing<br />

tank of Osaka University. The forces<br />

applied to the body were measured<br />

by the force sensor at various towing<br />

speeds. From this experiment, the relationship<br />

between the drag forces <strong>and</strong><br />

the fluid velocities was obtained to estimate<br />

the thrust forces produced by<br />

the fins. The swimming speeds in<br />

free swimming condition of the robotic<br />

turtle were measured in the<br />

same towing tank for the estimation<br />

of the swimming performance. The<br />

lift-based swimming mode was<br />

adopted in the experiments. It uses<br />

the flapping motion <strong>and</strong> the feathering<br />

motion in horizontal plane of the<br />

robot; on the other h<strong>and</strong>, it uses the<br />

rowing motion <strong>and</strong> the feathering motion<br />

in vertical motion of the robot.<br />

The speeds of the towing carriage<br />

were determined by operating it parallel<br />

to the robot for several values of<br />

amplitudes of flapping motion <strong>and</strong><br />

feathering motion. From the experimental<br />

result of the swimming speeds,<br />

the thrust forces were estimated by<br />

using the relationship between drag<br />

forces <strong>and</strong> swimming speeds. Figure<br />

22 shows the thrust forces for the<br />

amplitudes of feathering motion of<br />

30° <strong>and</strong> 45° <strong>and</strong> the amplitudes of flapping<br />

motion of 20° <strong>and</strong> 30°. From this<br />

figure, we can see that the thrust force<br />

increases as the amplitude of the<br />

flapping motion increases. Figure 23<br />

shows the thrust forces for amplitudes<br />

of a flapping motion of 30° versus amplitudes<br />

of the feathering motion.<br />

From this figure, we can see that the<br />

thrust force becomes the largest at the<br />

FIGURE 22<br />

amplitude of the feathering motion of<br />

30°. Judging from these results, the<br />

largest thrust force is made at the amplitudes<br />

of flapping motion <strong>and</strong> feathering<br />

motion of 30°.<br />

The swimming speeds of forward<br />

motion, backward motion, turning<br />

motion, <strong>and</strong> vertically descending motion<br />

were measured as performances of<br />

the maneuverability of the robotic turtle.<br />

Table 2 shows the swimming<br />

speeds of these motions with fixed amplitudes<br />

of joint angle. Unlike the forward<br />

motion, the swimming speed in<br />

backward motion with the amplitudes<br />

of feathering motion of 45° is larger<br />

than the case of 30°.<br />

Thrust forces in swimming motion against amplitudes of flapping motion <strong>and</strong> feathering motion.<br />

192 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 23<br />

Thrust forces in swimming motion against amplitudes of feathering motion.<br />

Walking Performance<br />

in Still Water<br />

TABLE 2<br />

The swimming speeds for various swimming motions.<br />

The crawl gait was adopted as the<br />

walking gait of the robotic turtle because<br />

it has a high level of stability.<br />

During the walking motion using the<br />

crawl gait, the robot moves each arm<br />

individually. This means the robot<br />

supports the body with at least three<br />

arms at all times. In operation of the<br />

robot in sea water, it is supposed that<br />

the seabed is not flat <strong>and</strong> there exist<br />

disturbances such as waves <strong>and</strong> water<br />

currents. Therefore, the robotic turtle<br />

needs more stability under such conditions.<br />

For that reason, we introduced a<br />

modified walking motion by considering<br />

the movement of the center of<br />

gravity in the normal crawl gait, as<br />

showninFigure24.Themovement<br />

of the center of gravity makes the stability<br />

margin larger, which is defined as<br />

the distance from the side of the support<br />

polygon to the projection of the<br />

center of gravity on l<strong>and</strong>, as shown in<br />

Figure 25.<br />

Two types of walking patterns were<br />

made. The first walking pattern consists<br />

of four steps to move an arm, for<br />

a total of 16 steps to move four arms<br />

during a cycle. The trajectory of the<br />

end of an arm in this pattern forms a<br />

rectangle. The second walking pattern<br />

consists of three steps to move an arm,<br />

for a total of 12 steps to move four<br />

Swimming Pattern<br />

Flapping (Rowing)<br />

Amplitude (°)<br />

Feathering<br />

Amplitude (°)<br />

Swimming<br />

Speed<br />

Forward motion 30 30 0.168<br />

Forward motion 30 45 0.15<br />

Backward motion 30 30 0.088 m/s<br />

Backward motion 30 45 0.102 m/s<br />

Turning motion 30 30 9.33°/s<br />

Turning motion 30 45 10.30°/s<br />

Descending motion 30 (rowing) 45 0.05 m/s<br />

arms during a cycle (Figure 26). The<br />

trajectory of the end of the arm of<br />

the second walking pattern forms a<br />

right triangle. In this case, it is expected<br />

that walking will be faster because of<br />

the reduction in the number of walking<br />

steps <strong>and</strong> that the energy consumption<br />

will be smaller because of the<br />

reduction of total motor speed.<br />

The walking speed <strong>and</strong> the motor<br />

currents for each walking pattern<br />

were measured while the robot walked<br />

on the bottom of the pool (L × B × H =<br />

4.5 m × 2.0 m × 0.8 m) (see Figure 27)<br />

to evaluate the walking performances<br />

in still water. Here the length of the<br />

stride of each walking pattern was set<br />

at 0.2 m. The joint torque was derived<br />

from the value of the motor current<br />

measured by the experiment, the<br />

torque constant, the reduction ratio,<br />

<strong>and</strong> the transmission efficiency, as<br />

showninEq.(12).Theconsumption<br />

energy was derived from the<br />

value of the derived joint torque, as<br />

follows:<br />

q ij<br />

¼ I ij ⋅K t ⋅i⋅η′<br />

ð12Þ<br />

Here, I ij denotes the motor current of<br />

the joint P ij ,K t is the torque constant,<br />

i is the reduction ratio, <strong>and</strong> η′ is the<br />

transmission efficiency. Figure 28<br />

shows a comparison of walking speeds<br />

between measurement <strong>and</strong> simulation<br />

by the walking simulator for the two<br />

walking patterns. From this figure,<br />

we can see that the walking speed for<br />

the second walking pattern is greater.<br />

This is attributed to the small period<br />

of the second walking pattern for one<br />

cycle because of the reduction in walking<br />

steps. We can see that the experimental<br />

value of the walking speed for<br />

each walking pattern is smaller than<br />

the calculated value. This is because<br />

the end of the arm slips on the bottom<br />

July/August 2011 Volume 45 Number 4 193


FIGURE 24<br />

Schematic view of crawl gait <strong>and</strong> modified crawl gait.<br />

FIGURE 27<br />

Picture of Robotic Turtle in walking in a pool.<br />

FIGURE 25<br />

Definition of stability margin.<br />

FIGURE 26<br />

Trajectories of the arm end of the first <strong>and</strong> second walking patterns.<br />

of the pool when the robot walks in<br />

still water.<br />

Walking Performance in Waves<br />

We estimated the walking performance<br />

of the robotic turtle in waves<br />

through a series of experiments. Figure<br />

29 shows a schematic view of the<br />

experiment for estimating the walking<br />

performance in waves. Two types of<br />

experiments were carried out in the<br />

towing tank of Osaka University.<br />

The first type measured the walking<br />

performance by changing the walking<br />

patterns in waves <strong>and</strong> in still water.<br />

The second type measured the walking<br />

performances of the second walking<br />

pattern in waves by changing the<br />

wave conditions, namely, the wave<br />

height, the wavelength, <strong>and</strong> the<br />

depth of the virtual ground. For the<br />

first type of experiment, the walking<br />

speeds of the first <strong>and</strong> second normal<br />

walking patterns were measured for<br />

thestridelengths(L)of0.2m<strong>and</strong><br />

0.3 m under the wave height of 0.1 m,<br />

the wavelength of 2 m, <strong>and</strong> the water<br />

depth of 0.9 m. The walking speeds<br />

of the first <strong>and</strong> second modified walking<br />

patterns were also measured. Figure<br />

30 shows a comparison of the<br />

walking speeds for various walking patterns<br />

in waves <strong>and</strong> in still water. From<br />

this figure, we can see that the walking<br />

speeds for all cases decreased in waves<br />

compared to the performance in still<br />

water. The walking speed reaches its<br />

maximum value in the case of the<br />

length of the stride of 0.3 m <strong>and</strong> the<br />

second walking pattern, regardless of<br />

crawl gait. Figure 31 shows a comparison<br />

of the consumption energy for<br />

various walking patterns in still water<br />

<strong>and</strong> in waves. From this figure, we<br />

can see that the consumption energy<br />

of the first walking pattern in waves<br />

is larger than that in still water in the<br />

caseofthelengthofthestrideof<br />

194 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 28<br />

Comparison of walking speed in water.<br />

FIGURE 29<br />

Schematic view of experiment for walking performance in waves.<br />

FIGURE 30<br />

Comparison of walking speeds for various walking patterns in still water <strong>and</strong> in waves.<br />

0.3 m, although that of the second<br />

walking pattern in waves is smaller<br />

than that in still water. Regarding<br />

that the walking speeds of the first<br />

<strong>and</strong> second walking patters in waves<br />

become smaller than those in still<br />

water, we find that the propulsive efficiency<br />

of the first walking pattern in<br />

waves becomes worse than that of the<br />

second walking pattern in waves in<br />

the case of the length of the stride of<br />

0.3 m. This may be attributed to the<br />

interaction between the hydrodynamic<br />

forces acting on the arms <strong>and</strong> the paths<br />

of the arms, namely, larger energy is<br />

consumed in the first walking pattern<br />

during one cycle of the arm motion<br />

in waves by the hydrodynamic forces<br />

than in the second walking.<br />

Figure 32 shows the variation of<br />

walking speed against the wave heights<br />

from 0 m to 0.15 m with the wavelength<br />

of 2 m, the water depths of<br />

0.6m<strong>and</strong>0.9m,<strong>and</strong>normal<strong>and</strong><br />

crawl modified gaits for the second<br />

walking pattern with the length of<br />

the stride of 0.3 m. From this figure,<br />

we can see that the walking speed is decreased<br />

as the wave height increases<br />

<strong>and</strong> that the walking speed becomes<br />

smaller <strong>and</strong> the rates of the decrease become<br />

larger as the water depth decreases<br />

from 0.9 m to 0.6 m. Figure 33 shows<br />

the variation of consumption energy<br />

against the wave heights under the<br />

same condition for Figure 30. From<br />

this figure,wecanseethattheconsumption<br />

energy is almost constant<br />

against the wave height, although the<br />

consumption energies are smaller<br />

than those in still water, <strong>and</strong> that the<br />

propulsive efficiency in waves become<br />

worse if the water depth decreases because<br />

the rate of decrease of walking<br />

speed for d = 0.6m is larger than for<br />

d = 0.9 m. This is attributed to larger<br />

water current induced by waves in<br />

shallow water, which produces larger<br />

July/August 2011 Volume 45 Number 4 195


FIGURE 31<br />

Comparison of consumption energy for various walking patterns in still water <strong>and</strong> in waves.<br />

FIGURE 32<br />

Variation of walking speed against wave height.<br />

hydrodynamic drag force on the<br />

vehicle.<br />

Summary<br />

In this paper we discussed the<br />

mechanisms of swimming <strong>and</strong> walking<br />

of turtles <strong>and</strong> examined how those<br />

mechanisms could be applied to an<br />

amphibious robot, with the goal of<br />

designing a robot that can perform<br />

automatic monitoring of environments<br />

along natural coastal areas <strong>and</strong><br />

tidal flats. Using a manipulator with<br />

4 DOF, we discussed the characteristics<br />

of walking of sea turtles <strong>and</strong> tortoises<br />

from the viewpoints of mobility<br />

<strong>and</strong> trafficability. The advantages of<br />

sea turtles in swimming <strong>and</strong> tortoises<br />

in walking were adopted in the robotic<br />

turtle. The experiments in a water tank<br />

revealed that the mobility <strong>and</strong> the trafficability<br />

of the amphibious robot at<br />

the bottom of the water were greatly<br />

affected by waves in shallow water.<br />

The author will investigate the tracking<br />

performance of the amphibious<br />

robot on tidal flats, moving from<br />

l<strong>and</strong> to sea <strong>and</strong> vice versa.<br />

FIGURE 33<br />

Variation of consumption energy against wave height.<br />

Acknowledgments<br />

This research was funded for<br />

3 years beginning in 2008 by the Ministry<br />

of Education, Culture, Sports <strong>and</strong><br />

<strong>Technology</strong>, Japan (grant 20246130)<br />

under the project title “Establishment<br />

of basic technology of a biomimetic<br />

underwater vehicle <strong>and</strong> its applications.”<br />

Author:<br />

Naomi Kato<br />

Graduate School of<br />

Engineering, Osaka University<br />

Yamadaoka 2-1, Suita,<br />

Osaka 565-0871, Japan<br />

Email: kato@naoe.eng.osaka-u.ac.jp<br />

196 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


References<br />

Dudek, G., Giguere, P., Prahacs, C.,<br />

Saunderson, S., Sattar, J., Torres-Mendez,<br />

L.-A., … Georgiades, C. 2007. Aqua: An<br />

amphibious autonomous robot. IEEE<br />

Comput Mag. 40(1):46-53.<br />

Ijspeert, A.J., Crespi, A., Ryczko, D., &<br />

Cabelguen, J-M. 2007. From swimming to<br />

walking with a salam<strong>and</strong>er robot driven by<br />

a spinal cord model. Science. 315(5817):<br />

1416-20. doi: 10.1126/science.1138353.<br />

Isobe, Y., Kato, N., Suzuku, H., Senga, H.,<br />

Matsuda, Y., Kawamura, K., … Oshika, T.<br />

2010. Motion analysis of sea turtle with<br />

prosthetic flippers. In: The 20th (2010) Int.<br />

Offshore <strong>and</strong> Polar Eng. Conf. pp. 335-42.<br />

Beijing: International <strong>Society</strong> of Polar <strong>and</strong><br />

Offshore Eng.<br />

Kato, N., Ando, Y., Shigetomi, T., &<br />

Katayama, T. 2006. Biology-inspired precision<br />

maneuvering of underwater vehicles (Part 4).<br />

Int J Offshore Polar. 16(3):195-201.<br />

Kato, N., & Liu, H. 2003. Optimization<br />

of motion of a mechanical pectoral fin.<br />

JSME Int J C. 47(4):1356-62. doi: 10.1299/<br />

jsmec.46.1356.<br />

amphibious walking robot—The design<br />

concepts <strong>and</strong> the 1st field experiment.<br />

In: paper presented at the International<br />

Symposium on Automation <strong>and</strong> Robotics<br />

in Construction 2006 (ISARC 2006). 46-51.<br />

Tokyo: IEEE.<br />

Wakimoto, S., Suzumori, K., & K<strong>and</strong>a, T.<br />

2006. A bio-mimetic amphibious soft cord<br />

robot. JSME Int J C. 72(714):171-7.<br />

Walker, W.F., Jr. 1971. A structual <strong>and</strong><br />

functional analysis of walking in the turtle<br />

Chrysemys picta marginata. J Morphol.<br />

134:195-213. doi: 10.1002/jmor.1051340205.<br />

Wyneken, J. 1997. Sea turtle locomotion:<br />

Mechanisms, behavior, <strong>and</strong> energetics. In:<br />

The Biology of Sea Turtles, eds. Lutz, P.L.,<br />

& Musick, J.A., 165-98. New York: CRC<br />

Press.<br />

Yamada, H., Chigisaki, S., Mori, M., Takita,<br />

K., Ogami, K., & Hirose, S. 2005. Development<br />

of amphibious snake-like robot ACM-R5.<br />

In: paper presented at the 36th International<br />

Symposium on Robotics(ISR 2005). Tokyo:<br />

Japan Robot Association.<br />

Low, K.H., Zhou, C., Ong, T.W., & Yu, J.<br />

2007. Modular design <strong>and</strong> initial gait study<br />

of an amphibious robotic turtle. In: paper<br />

presented at the International Conference<br />

on Robotics <strong>and</strong> <strong>Biomimetics</strong>, 2007. 535-40.<br />

Sanya, China: IEEE.<br />

Setouchi, H., & Shinjo, T. 2001. The effect<br />

of water on frictional characteristics between<br />

calcareous s<strong>and</strong>s <strong>and</strong> a steel plate. In: Report<br />

of Faculty of Agriculture. University of the<br />

Ryukyus. 48:103-11.<br />

Suzuki, H., & Kato, N. 2005. A numerical<br />

study on unsteady flow around a mechanical<br />

pectoral fin. Int J Offshore Polar. 15(3):161-7.<br />

Suzuki, H., Kato, N., & Suzumori, K. 2007.<br />

Load characteristics of mechanical pectoral<br />

fin. Exp Fluids. 44(5):759-71. doi: 10.1007/<br />

s00348-007-0432-x.<br />

Tanaka, T., & Shirai, T. 2006. Development<br />

of automated shoreline surveying system using<br />

July/August 2011 Volume 45 Number 4 197


PAPER<br />

<strong>Marine</strong> Applications of the Biomimetic<br />

Humpback Whale Flipper<br />

AUTHORS<br />

Frank E. Fish<br />

West Chester University<br />

Paul W. Weber<br />

Applied Research Associates, Inc.<br />

Mark M. Murray<br />

United States Naval Academy<br />

Laurens E. Howle<br />

Duke University<br />

Introduction<br />

Life began with water. The highdensity<br />

<strong>and</strong> viscous nature of water<br />

has imposed a strong evolutionary selection<br />

pressure on the design of animals<br />

that move through this aqueous<br />

medium. Over the course of millions<br />

of years, different phylogenetic lines<br />

of animals have, in effect, experimented<br />

with various combinations<br />

of morphologies <strong>and</strong> behaviors to enhance<br />

locomotor performance. The<br />

great diversity of body shapes, surface<br />

textures, <strong>and</strong> propulsive mechanisms<br />

exhibited by aquatic animals has produced<br />

a variety of biomechanical solutions<br />

for the reduction of drag, increase<br />

in thrust production <strong>and</strong> efficiency,<br />

maintenance of stability, <strong>and</strong> enhancement<br />

of maneuverability. By emulating<br />

these biological characteristics in<br />

those instances where animal performance<br />

is superior to manufactured devices,theperformanceofengineered<br />

marine systems may be improved<br />

through the field of biomimetics.<br />

The cetaceans (whales, dolphins,<br />

porpoises) have been the focus of inspiration<br />

for technological development in<br />

ABSTRACT<br />

The biomimetic approach seeks technological advancement through a transfer<br />

of technology from natural technologies to engineered systems. The morphology of<br />

the wing-like flipper of the humpback whale has potential for marine applications.<br />

As opposed to the straight leading edge of conventional hydrofoils, the humpback<br />

whale flipper has a number of sinusoid-like rounded bumps, called tubercles, which<br />

are arranged periodically along the leading edge. The presence of the tubercles<br />

modifies the water flow over the wing-like surface, creating regions of vortex generation<br />

between the tubercles. These vortices interact with the flow over the tubercle<br />

<strong>and</strong> accelerate that flow, helping to maintain a partially attached boundary layer.<br />

This hydrodynamic effect can delay stall to higher angles of attack, increases lift,<br />

<strong>and</strong> reduces drag compared to the post-stall condition of conventional wings. As<br />

the humpback whale functions in the marine environment in a Reynolds regime<br />

similar to some engineered marine systems, the use of tubercles has the potential<br />

to enhance the performance of wing-like structures. Specific applications of the tubercles<br />

for marine technology include sailboat masts, fans, propellers, turbines, <strong>and</strong><br />

control surfaces, such as rudders, dive planes, stabilizers, spoilers, <strong>and</strong> keels.<br />

Keywords: tubercles, delayed stall, Megaptera novaeangliae, leading edge,<br />

bio-inspired design<br />

the marine environment. The cetacean<br />

lineage dates back 55 million years.<br />

The intense selection pressures for a<br />

fast swimming, maneuverable marine<br />

predator have culminated in a highly<br />

streamlined body with advanced sensory<br />

capabilities that is propelled by a<br />

highly efficient propulsion mechanism.<br />

There are a number of examples<br />

where cetaceans have been the inspiration<br />

for the development or improvements<br />

of marine technology. The<br />

cetacean body shape was used by<br />

Cayley (circa 1800) as a solid of leastresistance<br />

for the development of airplane<br />

fuselage <strong>and</strong> boat hull designs<br />

(Gibbs-Smith, 1962). The famous<br />

but erroneous “Gray’s paradox” led<br />

to examination of special drag reduction<br />

mechanisms (Gray, 1936; Fish<br />

& Rohr, 1999; Fish, 2006), including<br />

the biomimetic development of compliant<br />

coatings for viscous dampening<br />

(Kramer, 1960; Riley et al., 1988;<br />

Carpenter & Pedley, 2003). The compactness<br />

<strong>and</strong> high resolution of the<br />

echolocation system of dolphins provides<br />

a benchmark for the improvement<br />

of SONAR systems (Au, 1993).<br />

The thrust performance of oscillating,<br />

wing-like systems, such as the flukes of<br />

dolphins, has been considered superior<br />

to screw propellers (Peterson, 1925;<br />

Liu & Bose, 1993; Triantafyllou &<br />

Triantafyllou, 1995). The flexibility<br />

of the oscillatory flukes can allow operation<br />

at high efficiency over an extended<br />

speed range without cavitation<br />

(Fish & Lauder, 2006; Iosilevskii &<br />

Weihs, 2007).<br />

This report focuses on a unique morphology<br />

of a highly derived aquatic<br />

198 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


mammal that has general biomimetic<br />

applications for marine systems. The<br />

flipper of the humpback whale <strong>and</strong><br />

its bumpy leading edge provide a<br />

novel approach to enhance the hydrodynamic<br />

performance of wing-like<br />

structures for operation in water. A<br />

principal attribute of using the whale<br />

as a model to construct a biomimetic<br />

system is the scale. As the whale is of<br />

a large size <strong>and</strong> swims at speeds that<br />

compliment the scale <strong>and</strong> operation<br />

of engineered marine systems, application<br />

for marine technologies can be<br />

readily undertaken.<br />

Humpback Whale<br />

as Inspiration<br />

The humpback whale (Megaptera<br />

novaeangliae) hasthelongestflipper<br />

of any cetacean (i.e., whale, dolphin,<br />

porpoise), with regard to both absolute<br />

<strong>and</strong> relative size (Figure 1; Fish<br />

& Battle, 1995). The flippers are involved<br />

with the underwater maneuvers<br />

performed by the species that are associated<br />

with their mode of feeding<br />

(Friedlaender et al., 2009; Hazen<br />

et al., 2009). Humpback whales are<br />

the only baleen whales (e.g., blue<br />

whale, fin whale, minke whale, right<br />

whale) that rely on tight, rapid turns<br />

to capture prey (Fish & Battle,<br />

1995). The humpback whales use<br />

their flippers as biological hydroplanes<br />

to achieve tight circles to corral <strong>and</strong><br />

engulf prey (Fish et al., 2011).<br />

FIGURE 1<br />

Flippers on humpback whale (left), showing<br />

scalloped pattern of tubercles (center), <strong>and</strong> flipper<br />

cross-section (right) (from Fish & Lauder,<br />

2006).<br />

The humpback whale flippers are<br />

unique because of the presence of<br />

large tubercles along the leading edge,<br />

which gives this surface a scalloped<br />

appearance (Figure 1). The distances<br />

between tubercles decrease distally,<br />

although these distances remain relatively<br />

constant at 7-9% of span over<br />

the midspan of the flipper (Fish &<br />

Battle, 1995).<br />

The planform <strong>and</strong> cross-sectional<br />

views of the humpback whale flipper<br />

are shown in Figure 1. Whereas typical<br />

wing-like structures have a straight<br />

leading edge without the presence of<br />

irregularities or perturbations, the<br />

humpback flipper defies convention<br />

with prominent rounded bumps that<br />

are regularly spaced along the leading<br />

edge of its high-aspect ratio flipper<br />

(Fish et al., 2008). Humpback<br />

whale flippers closely resemble the<br />

21% thick, low drag NACA 63 4 -021<br />

foil in cross section (Abbott & von<br />

Doenhoff, 1959; Fish & Battle,<br />

1995). Furthermore, the flippers have<br />

high mobility (Edel & Winn, 1978).<br />

The elongate flippers function as<br />

wings to generate the forces necessary<br />

for turning maneuvers (Fish et al.,<br />

2011). Turning is important in the<br />

capture of elusive prey. Humpback<br />

whales feed on shoals of small fish<br />

<strong>and</strong> krill. The preys are forced into a<br />

tight ball by the whales striking the<br />

water surface with their flukes or circling<br />

the prey from underneath while<br />

emitting bubbles. The bubbles rise to<br />

corral the prey as a bubble net. In either<br />

case, the whale maneuvers under<br />

the prey for engulfment by executing<br />

a rapid turn. Lift generated by the flippers<br />

is used to produce a centripetal<br />

force for the turn. The tubercles produce<br />

vortical flows over the surface of<br />

the flipper <strong>and</strong> control lift characteristics<br />

at high angles of attack <strong>and</strong> delay<br />

stall (Fish & Battle, 1995; Miklosovic<br />

et al., 2004; Fish & Lauder, 2006; Fish<br />

et al., 2011).<br />

Hydrodynamic Effect<br />

of Tubercles<br />

The prominence of the tubercles<br />

on the leading edge of the humpback<br />

whale flippers <strong>and</strong> the swimming pattern<br />

of the whale suggests that these<br />

novel structures have a distinct hydrodynamic<br />

effect (Figure 2). In addition,<br />

the pattern of barnacle distribution on<br />

the flippers (i.e., barnacles are confined<br />

to the peak of the tubercle <strong>and</strong> do<br />

not occur between tubercles; Fish &<br />

Battle, 1995) indicates that the flow<br />

over the flipper is affected by tubercles.<br />

This section reviews potential hydrodynamic<br />

advantages, flow control,<br />

<strong>and</strong> limitations of the tubercles on<br />

wing-like structures.<br />

Hydrodynamic Advantages<br />

The presence of leading-edge tubercles<br />

on a wing-like structure can have<br />

a positive influence on the hydrodynamic<br />

performance. Wind tunnel<br />

tests showed that wings with tubercles<br />

improved maximum lift by over 6%,<br />

increased the ultimate stall angle by<br />

40%, <strong>and</strong> decreased drag by as much<br />

as 32% (Figure 3; Miklosovic et al.,<br />

2004). The tubercles, when facing<br />

FIGURE 2<br />

Idealized humpback flipper with leading-edge<br />

tubercles.<br />

July/August 2011 Volume 45 Number 4 199


FIGURE 3<br />

Humpback whale flipper models <strong>and</strong> results of wind tunnel experiments. The models (left) with<br />

<strong>and</strong> without tubercles were machined from clear polycarbonate, based on a symmetrical NACA<br />

0020 foil section. Lift <strong>and</strong> drag data (right) for the flipper models were obtained from tests in a<br />

wind tunnel. The solid lines in a, b, <strong>and</strong> c show the average of the data for the flipper model without<br />

tubercles, <strong>and</strong> open triangles are for the model with tubercles. Lift coefficient C L (a), drag coefficient<br />

C D (b), <strong>and</strong> aerodynamic efficiency L/D (c) are plotted against angle of attack, α (from<br />

Miklosovic et al., 2004).<br />

intothefreestreamflow, alter the<br />

fluid flow over wing-like structures<br />

(Bushnell & Moore, 1991; Fish &<br />

Battle, 1995; Fish et al., 2011). Furthermore,<br />

the lift to drag (L/D) ratio,<br />

which represents the aerodynamic efficiency,<br />

displayed a greater peak L/D<br />

for a wing geometry with tubercles<br />

(Miklosovic et al., 2004, 2007;<br />

Hansen et al., 2009).<br />

The position <strong>and</strong> number of tubercles<br />

on the flipper suggest analogues<br />

with specialized leading edge control<br />

devices that improve the hydrodynamic<br />

performance of wings. The<br />

occurrence of “morphological complexities”<br />

on a lifting body could reduce<br />

or use pressure variation at the tip to decrease<br />

drag <strong>and</strong> improve lift to prevent<br />

tip stall (Bushnell & Moore, 1991). Alternatively,<br />

various biological wings<br />

utilize leading-edge control devices to<br />

maintain lift <strong>and</strong> avoid stall at high attack<br />

angles <strong>and</strong> low speeds.<br />

The function of the tubercles may<br />

be analogous to strakes used on aircraft.<br />

Strakes are large vortex generators<br />

that change the stall characteristics<br />

of a wing (Hoerner, 1965;<br />

Shevell, 1986; Bertin & Smith,<br />

1998). Stall is postponed because the<br />

vortices exchange momentum within<br />

the boundary layer to keep it partially<br />

attached over the wing surface. Lift is<br />

thus maintained at higher angles of attack<br />

with strakes compared to wings<br />

without strakes, although maximum<br />

lift is not increased by strakes (Shevell,<br />

1986). Another leading-edge device are<br />

slots or slats that delay stall or move<br />

the angle of attack of maximum lift<br />

to a lower value (Wegener, 1991;<br />

Bertin & Smith, 1998). The moveable<br />

slats create a space anterior of the fixed<br />

wing to allow higher-pressure fluid to<br />

rise from the underside of the wing.<br />

The movement of fluid from the<br />

high-pressure side to the low-pressure<br />

side of the wing improves mixing <strong>and</strong><br />

helps to maintain the boundary layer<br />

<strong>and</strong> delay stall (Wegener, 1991). However,<br />

the tubercles have distinct advantages<br />

over slats. Tubercles are<br />

passive structures that have no drag<br />

penalty when designed onto wings<br />

(Miklosovicetal.,2004),whereas<br />

slats are actively deployed <strong>and</strong> incur increased<br />

drag (Hoerner, 1965).<br />

Vortex Generation<br />

The mechanism for enhanced hydrodynamic<br />

performance due to the<br />

presence of tubercles appears due to<br />

the specific pattern of vortex generation<br />

over the surface of the flipper.<br />

Flow visualization experiments on<br />

wavy bluff bodies showed periodic<br />

variation in the wake width across the<br />

span (Owen et al., 2000). A wide wake<br />

with two simultaneous vortices occurred<br />

where the body protruded<br />

downstream <strong>and</strong> a narrow wake occurred<br />

where the body protruded upstream.<br />

The flowinthewakeofthe<br />

wavy body was different from the<br />

wake of a straight cylinder, which exhibited<br />

a typical von Karman Vortex<br />

Street of alternating vortex pairs. A<br />

bluff body with a spanwise sinusoidal<br />

form could reduce drag by at least<br />

30%, compared to equivalent straight<br />

bodies (Bearman & Owen, 1998).<br />

The vortices produced by a wing<br />

section with tubercles are shown in<br />

Figure 4. The tubercles generate<br />

separated, chordwise vortices in the<br />

troughs at high angles of attack.<br />

These vortices are formed as the flow<br />

strikes the leading edge of the trough.<br />

As the flow does not strike the leading<br />

edge normally, the flow is sheared into<br />

the trough’s center to generate the<br />

vortices. These vortices are convected<br />

along the chord. The spanwise arrangement<br />

of the vortices is in a pair<br />

on each side of the tubercle crest with<br />

opposite spins (Hansen et al., 2010).<br />

The flow directly over the tubercle interacts<br />

with the vortices located downstream<br />

<strong>and</strong> lateral to the tubercle crest.<br />

200 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 4<br />

Pressure contours <strong>and</strong> streamlines at α = 10° for NACA 63-021 with straight leading edge (left) <strong>and</strong><br />

with tubercles (right). An unsteady Reynolds-averaged Navier-Stokes (RANS) simulation was<br />

used. A separation line is shown on the wing section without tubercles. For the wing with tubercles,<br />

large vortices are formed posterior of the troughs along the leading edge <strong>and</strong> flow posterior of<br />

the tubercles is shown as straight streamlines without separation. Images courtesy of E. Paterson.<br />

The tangential velocities of the inward<br />

facing flows of the pair of vortices are<br />

directed toward the trailing edge of<br />

the wing section. The flow from the<br />

tubercle peak is accelerated posteriorly<br />

due to the interaction with the vortex<br />

pair. These effects prevent the local<br />

boundary layer downstream of the tubercles<br />

from separating <strong>and</strong> push the<br />

stall line further posterior toward the<br />

trailing edge. When integrated over<br />

the entire structure, the wing with tubercles<br />

will stall at a higher angle of<br />

attack than a wing without tubercles.<br />

Flow experiments conducted on a<br />

model wing section with leading-edge<br />

tubercles at low speeds showed flow<br />

separation from the troughs between<br />

adjacent tubercles but attached flow<br />

on the tubercles (Johari et al., 2007).<br />

Flow separation pattern <strong>and</strong> surface<br />

pressure was dramatically altered by<br />

the tubercles. For regions downstream<br />

of tubercle crest, separation was delayed<br />

almost to the trailing edge. Although<br />

this flow pattern did not<br />

result in improved lift generation,<br />

drag reduction or delay in the stall<br />

angle of attack, the post-stall characteristics<br />

were greatly smoothed (Johari<br />

et al., 2007; Saadat et al., 2010),<br />

which could provide important performance<br />

benefits for systems that routinely<br />

operate beyond the stall point.<br />

The vortices produced from the<br />

tubercles re-energize the boundary<br />

layer by carrying high-momentum<br />

flow close to the flipper’s surface<br />

(Figure 5; Wu et al., 1991; Pedro &<br />

FIGURE 5<br />

Kobayashi, 2008; Hansen et al.,<br />

2010). The flow dynamics are improved<br />

also by confining separation<br />

to the tip region. Tubercles delay stall<br />

by causing a greater portion of the flow<br />

to remain attached on a wing, with the<br />

attached flow localized behind the tubercle<br />

crests (Weber et al., 2011).<br />

Thesize<strong>and</strong>frequencyofthetubercles<br />

along the leading edge influences<br />

the performance of a wing.<br />

Small amplitude tubercles show the<br />

best performance with regard to lift<br />

<strong>and</strong> stall characteristics (Johari et al.,<br />

2007; Hansen et al., 2009, 2011).<br />

The wavelength <strong>and</strong> thus frequency<br />

of tubercles was found, however, to<br />

have little effect on performance<br />

(Johari et al., 2007).<br />

The tubercle effect is further enhanced<br />

when sweepback is added to<br />

the wings (Murray et al., 2005).<br />

Wings with sweep angles of 15° <strong>and</strong><br />

30°requiredhigheranglesofattack<br />

to achieve stall than nonswept wings<br />

<strong>and</strong> showed superior drag performance<br />

over most of the range of α compared<br />

to models without tubercles (Murray<br />

et al., 2005). Flow tests on delta<br />

Vorticity computed from detached eddy simulation (DES) for flippers with (a) <strong>and</strong> without (b) tubercles<br />

at an angle of attack of 15°. Vortices re-energize boundary layer to delay separation <strong>and</strong><br />

stall. Images courtesy of H. T. C. Pedro <strong>and</strong> M. H. Kobayashi.<br />

July/August 2011 Volume 45 Number 4 201


wings with a sweep of 50° showed that<br />

at high angles of attack large-scale<br />

three-dimensional separation occurred<br />

for the wing with a straight<br />

leading edge (Goruney & Rockwell,<br />

2009). However when tubercles are<br />

added, the flow is radically transformed.<br />

Tubercles with amplitude of<br />

4% of wing chord can completely<br />

eradicate the negative effect of the<br />

separation <strong>and</strong> foster re-attachment.<br />

Experiments performed on flapping<br />

wings with tubercles (Figure 6)<br />

showed an affect on the spanwise<br />

flow (Ozen & Rockwell, 2010). Typically<br />

a straight wing, whether flapping<br />

FIGURE 6<br />

or static, will develop a spanwise flow<br />

due to the pressure differential that develops<br />

between the upper <strong>and</strong> lower<br />

surfaces. This spanwise flow becomes<br />

manifest as a wing tip vortex, which increases<br />

the drag <strong>and</strong> reduces the efficiency<br />

of a wing. A flapping wing<br />

with tubercles does not produce a pronounced<br />

region of spanwise flow, but<br />

the wing tip vortex generation is unaffected<br />

(Ozen & Rockwell, 2010).<br />

Limitations of Tubercles<br />

Enhanced performance due to the<br />

presence of the tubercles is not universal.<br />

There exist limitations to the<br />

Comparison of flapping plate at Reynolds number of 1300 at angle of attack of 8° with <strong>and</strong> without<br />

tubercles from Ozen <strong>and</strong> Rockwell (2010). The pattern of the flow structure for the plate with tubercles<br />

produces a series of spanwise vortices that limit spanwise flow compared with the flapping<br />

plate without tubercles. Image courtesy of D. Rockwell.<br />

advantages of the tubercles. Tubercles<br />

improve performance when in concert<br />

with wing geometries that are characterized<br />

by the combination of finite<br />

span, swept wing, tapered planform,<br />

<strong>and</strong> thick foil.<br />

Foil sections with no wing tip that<br />

emulate infinite wings do not demonstrate<br />

reduced drag <strong>and</strong> increased lift, although<br />

stall is still delayed (Miklosovic<br />

et al., 2004; Johari et al., 2007; Van<br />

Nierop et al., 2008). Tip effects occur<br />

as a consequence of lift generation<br />

when a fully three-dimensional wing<br />

is canted at an angle of attack to an incident<br />

flow. Induced drag is produced<br />

in lift generation from kinetic energy<br />

imparted to the fluid from pressure differences<br />

between the two surfaces of<br />

the wing as there is leakage of fluid<br />

from high pressure to low pressure<br />

around the distal tip of a lifting surface<br />

resulting in spanwise flow <strong>and</strong> the formation<br />

of tip vortices (Vogel, 1981).<br />

The flow pattern set up by the tubercles<br />

helps to maintain a chordwise<br />

flow <strong>and</strong> reduce the induced drag due<br />

to tip vortices. This effect is only realized<br />

for finite wings.<br />

The tubercles require wings with<br />

thick sectional geometries to function.<br />

The section must have a prominent<br />

nose radius. In part, this is due to the<br />

necessity to contour the tubercles threedimensionally<br />

into the leading edge<br />

to avoid any flat surfaces. Leadingedge<br />

tubercles were tested on foils<br />

based on NACA 0021, 63 4 -021 <strong>and</strong><br />

65-021 sections (Miklosovic et al.,<br />

2004; Johari et al., 2007; Custodio<br />

et al., 2010; Hansen et al., 2009,<br />

2011). These designs approach the<br />

cross-sectional geometry of the humpback<br />

whale flipper (Fish & Battle,<br />

1995).<br />

It is necessary to have a relatively<br />

steady flow to maintain the pattern of<br />

thevortices<strong>and</strong>incurthehydrodynamic<br />

202 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


advantages (Stanway, 2008). Foils<br />

with tubercles that were oscillated in<br />

roll <strong>and</strong> pitch demonstrated that the<br />

tubercles did not improve hydrodynamic<br />

performance (Stanway, 2008).<br />

Flapping degraded performance by<br />

the redirection of energy to tuberclegenerated<br />

vortices from the vortices<br />

of the wake, which are necessary for<br />

thrust production during flapping.<br />

Furthermore, if the period of oscillations<br />

is too rapid, there may be<br />

insufficient time to allow the full development<br />

of the vortices over a wing. It is<br />

necessary to have a relatively steady<br />

flow to maintain the pattern of the<br />

vortices <strong>and</strong> incur the hydrodynamic<br />

advantages (Stanway, 2008).<br />

Tubercle Technologies<br />

Application of natural technologies<br />

into biomimetic-engineered systems<br />

has a number of problems that are inherent<br />

due to differences in biological<br />

<strong>and</strong> engineered systems (Fish, 2006).<br />

Engineered systems are relatively large<br />

in size, are composed of dry rigid<br />

materials, including metals <strong>and</strong> ceramics,<br />

use rotation motors, <strong>and</strong> are<br />

controlled by computational systems<br />

with limited sensory feedback. Biological<br />

structures associated with animal<br />

systems are relatively small in size, are<br />

composed of wet compliant materials,<br />

including composites of ceramics,<br />

polysaccharides <strong>and</strong> proteins, move<br />

by translational displacements generated<br />

by muscles, <strong>and</strong> are controlled<br />

by complex neural networks with multiple<br />

sensory inputs <strong>and</strong> fine scale<br />

motor outputs.<br />

The humpback whale tubercles<br />

provide an ideal solution for application<br />

to engineered marine systems.<br />

The size of the whale <strong>and</strong> its flippers<br />

operate near or at the same scale as<br />

some marine vehicles. The mature<br />

whales have a maximum length of<br />

17 m <strong>and</strong> weigh 40,000 kg (Clapham<br />

& Mead, 1999). Flipper length is approximately<br />

one-third the length of<br />

the animal <strong>and</strong> can be over 5 m. The<br />

whale cruises between 1.1 <strong>and</strong> 4.0 m/s<br />

<strong>and</strong> is able to burst to a speed of 7.5 m/s<br />

(Fish & Rohr, 1999). The flow experienced<br />

by <strong>and</strong> modified by the tubercles<br />

is within the same Reynolds regime<br />

that coincides with a large array of<br />

engineered applications. The Reynolds<br />

number of a flipper is 1.6 × 10 6 when<br />

the whale is lunge feeding (2.6 m/s).<br />

Thus, the flipper <strong>and</strong> tubercles are operating<br />

in a turbulent flow regime,<br />

which is the st<strong>and</strong>ard operating condition<br />

for most marine systems. Furthermore,<br />

the tubercle functions passively<br />

to modify flow <strong>and</strong> maintain favorable<br />

hydrodynamic conditions. Therefore,<br />

control systems can be simplified.<br />

Control Surfaces<br />

There are few other passive means<br />

of altering fluid flow around a winglike<br />

structure that can delay stall <strong>and</strong><br />

both increase lift <strong>and</strong> reduce drag at<br />

the same time. As a result, the application<br />

of leading-edge tubercles for passive<br />

flow control has potential in the<br />

design of marine technologies. The<br />

ubiquity of wing-like structures with<br />

marine applications for stability <strong>and</strong><br />

maneuverability presents an opportunity<br />

to enlist tubercles to improve performance.<br />

Included in such structures<br />

are fixed surfaces, such as keels, fins<br />

<strong>and</strong> skegs, <strong>and</strong> mobile control surfaces,<br />

such as rudders <strong>and</strong> dive planes.<br />

Delay of stall by tubercles on both<br />

fixed <strong>and</strong> mobile control surfaces<br />

provides a benefit in tight turning situations.<br />

To produce the required centripetal<br />

force to effect a turn, a lift force<br />

that is directed toward the center of the<br />

turn is generated by the control surface.<br />

The magnitude of the lift is<br />

directly dependent on the angle of attack<br />

with a higher angle providing a<br />

greater centripetal force. As stall can<br />

be delayed with tubercles, the control<br />

surface can operate at higher angles of<br />

attack producing a tighter turn radius<br />

with more control. If stall were to<br />

occur, the control surface would not<br />

be able to generate the centripetal<br />

force to maintain the turn. In effect,<br />

itwouldbelikedrivingacaralonga<br />

curved road <strong>and</strong> slipping on a patch<br />

of ice. The reduced friction <strong>and</strong> centripetal<br />

force between the ice <strong>and</strong> the<br />

tires would cause the car to drive off<br />

the road tangential to the curve, rather<br />

than following the original curved<br />

trajectory.<br />

A low-aspect ratio rudder with<br />

tubercles (Figure 7) <strong>and</strong> an unswept<br />

leading edge generated more lift at angles<br />

of attack above 22° compared to a<br />

smooth rudder at a Reynolds number<br />

of 200,000 (Weber et al., 2010). At<br />

higher Reynolds numbers, this effect<br />

diminishes <strong>and</strong> the tubercles accelerate<br />

the onset of cavitation. A humanpowered<br />

submarine, Umpty Squash,<br />

utilized tubercled dive planes <strong>and</strong> rudders<br />

(Figure 8). Students of the Sussex<br />

FIGURE 7<br />

Rudder with leading-edge tubercles.<br />

July/August 2011 Volume 45 Number 4 203


FIGURE 8<br />

Human-powered submarine (left) with rudder <strong>and</strong> dive planes with tubercles (right). Courtesy of<br />

Chris L<strong>and</strong> <strong>and</strong> the Sussex County Technical School.<br />

FIGURE 11<br />

Iceboat with leading-edge tubercles on the<br />

mast supporting a sail.<br />

County Technical High School,<br />

Sparta, NJ, constructed the submarine.<br />

In 2005, the submarine competed<br />

in the International Submarine<br />

RacesheldattheDavidTaylor<br />

Model Basin in Bethesda, MD. The<br />

submarine was capable of making a<br />

90° turn within 25 feet. The designers<br />

at Feadship De Voogt developed<br />

Breathe, a concept superyacht that<br />

incorporates biomimicry into the<br />

design (www.feadship.nl). The<br />

stabilizers <strong>and</strong> steering fins were<br />

based on the humpback whale flipper<br />

with tubercles (Figure 9).<br />

Commercially, the company Fluid<br />

Earth markets a surfboard skeg with<br />

leading-edge tubercles (Figure 10;<br />

Anders, 2009). The addition of tubercles<br />

would provide enhanced control<br />

during a cutback, when a surfer rapidly<br />

changes direction by 180° to maneuver<br />

the surfboard opposite to the direction<br />

of the wave’s braking motion.<br />

Application of tubercles to the mast<br />

of a sailboat (Figure 11) could be<br />

useful during close reach maneuvers.<br />

A close reach would have the sail set<br />

FIGURE 10<br />

Surfboard skeg with leading-edge tubercles.<br />

with a high angle of attack to the<br />

apparent wind. The lack of a thick<br />

cross-section by the sail itself may preclude<br />

any advantage in lift <strong>and</strong> stall.<br />

However, the presence of tubercles<br />

on the mast, representing a bluff<br />

body, could have advantages in terms<br />

of drag. Bluff bodies, like cylinders,<br />

FIGURE 9<br />

Conceptual yacht incorporating biomimetic<br />

structures. Stabilizers are shaped like the flippers<br />

of the humpback whale. Image courtesy<br />

of Feadship.<br />

204 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 12<br />

Two views of marine propeller designs with<br />

tubercles.<br />

can experience lowered drag when the<br />

leading edge has a sinusoidal design<br />

(Bearman & Owen, 1998).<br />

Propellers <strong>and</strong> Turbines<br />

The use of tubercles can effectively<br />

be employed in the generation of<br />

power by wing-like structures. Propellers<br />

operating in a marine system have<br />

the potential to be improved by the<br />

addition of tubercles (Figure 12).<br />

The effective angle of attack of a propeller<br />

blade can be increased by increasing<br />

the blade angle (Larrabee,<br />

1980). A higher angle of attack can<br />

produce more lift to derive greater<br />

thrust <strong>and</strong> increase the effective pitch<br />

of the propeller. Prevention of stall<br />

by the tubercle effect would reduce<br />

flow-induced vibrations. Furthermore,<br />

suppression of tonal noise is possible<br />

by the addition of tubercles to a propeller<br />

(Hansen et al., 2010). Tonal<br />

noise is most effectively reduced by<br />

large amplitude <strong>and</strong> smaller wavelength<br />

tubercles. As propellers produce<br />

a particular noise signature, reduction<br />

of noise would be advantageous for<br />

stealth in naval operations. Similarly,<br />

a reduction of noise pollution by commercial<br />

marine traffic could be beneficial<br />

to marine organisms, although the<br />

use of radiated noise by whales to avoid<br />

collisions with ships may be negatively<br />

impacted.<br />

Tubercle modified blades were also<br />

found to be effective in power generationofamarinetidalturbineatlow<br />

flow speeds (Murray et al., 2010). Tubercles<br />

were placed on the distal 40%<br />

of the three turbine blades. Compared<br />

to blades with smooth leading edges,<br />

blades with leading-edge tubercles<br />

demonstrated enhanced performance.<br />

The marine tidal turbine is analogous<br />

to wind turbines. A variable pitch<br />

wind turbine with retrofitted blades<br />

with tubercles demonstrated increased<br />

electrical generation at moderate wind<br />

speeds compared to unmodified blades<br />

(Howle, 2009; Wind Energy Institute<br />

of Canada, 2008).<br />

Conclusions<br />

A passive means of altering fluid<br />

flow around a wing-like structure that<br />

can delay stall <strong>and</strong> both increase lift<br />

<strong>and</strong> reduce drag simultaneously is<br />

highly novel. This performance by<br />

leading-edge tubercles, therefore, has<br />

potential application for passive flow<br />

control in the design of various marine<br />

technologies. The flow is modified by<br />

the formation of paired vortices in the<br />

troughs between tubercles. These vortices<br />

interact with the flow over the tubercles<br />

to keep the flow attached to the<br />

wing surface <strong>and</strong> delay stall. The tubercles<br />

perform best when designed into<br />

tapered wings with finite span, swept<br />

planform, <strong>and</strong> with a thick foil section<br />

that have limited oscillatory movement.<br />

Such applications for marine technology<br />

include fins, rudders, dive planes,<br />

water turbines <strong>and</strong> propellers. The fusion<br />

of these marine devices <strong>and</strong> tubercles<br />

can produce biomimetic designs<br />

that can exhibit superior performance<br />

to conventional engineered systems.<br />

Acknowledgments<br />

We thank the technical support<br />

staff of the United States Naval<br />

Academy. This work was supported<br />

by the National Science Foundation<br />

(IOS-0640185) to FEF <strong>and</strong> the National<br />

Defense Science <strong>and</strong> Engineering<br />

Graduate (NDSEG) Fellowship<br />

to PWW.<br />

Lead Author:<br />

Frank E. Fish<br />

Department of Biology<br />

West Chester University<br />

West Chester, PA 19383<br />

Email: ffish@wcupa.edu<br />

References<br />

Abbott, I.H., & von Doenhoff, A.E. 1959.<br />

Theory of Wing Sections. New York: Dover.<br />

Anders, M. 2009. Fins I like. Surfer’s Path,<br />

April/May 2009:116.<br />

Au, W.W.L. 1993. The Sonar of Dolphins.<br />

New York: Springer-Verlag.<br />

Bearman, P.W., & Owen, J.C. 1998.<br />

Reduction of bluff body drag <strong>and</strong> suppression<br />

of vortex shedding by the introduction of<br />

wavy separation lines. J Fluid Struct.<br />

12:123-30. doi: 10.1006/jfls.1997.0128.<br />

Bertin, J.J., & Smith, M.L. 1998. Aerodynamics<br />

for Engineers. Upper Saddle River,<br />

NJ: Prentice Hall.<br />

Bushnell, D.M., & Moore, K.J. 1991. Drag<br />

reduction in nature. Annu Rev Fluid Mech.<br />

23:65-79. doi: 10.1146/annurev.<br />

fl.23.010191.000433.<br />

Carpenter, P.W., & Pedley, T.J. 2003.<br />

IUTAM Symposium on Flow Past Highly<br />

Compliant Boundaries <strong>and</strong> in Collapsible<br />

Tubes. Dordrecht: Kluwer.<br />

Clapham, P.J., & Mead, J.G. 1999.<br />

Megaptera novaeangliae. Mamm Spec.<br />

604:1-9.<br />

Custodio, D., Henoch, C., & Johari, H.<br />

2010. The performance of finite-span<br />

hydrofoils with humpback whale-like leading<br />

edge protuberances. 63 rd Ann Mtg APS Div<br />

July/August 2011 Volume 45 Number 4 205


Fluid Dyn, 55 (16), Nov. 21, 2010, Long<br />

Beach, CA.<br />

Edel, R.K., & Winn, H.E. 1978. Observations<br />

on underwater locomotion <strong>and</strong> flipper<br />

movement of the humpback whale Megaptera<br />

novaeangliae. Mar Biol. 48:279-87.<br />

doi: 10.1007/BF00397155.<br />

Fish, F.E. 2006. The myth <strong>and</strong> reality of<br />

Gray’s paradox: Implication of dolphin drag<br />

reduction for technology. Bioinspir Biomim.<br />

1:R17-25. doi: 10.1088/1748-3182/1/2/R01.<br />

Fish, F.E., & Battle, J.M. 1995. Hydrodynamic<br />

design of the humpback whale flipper.<br />

J Morphol. 225:51-60. doi: 10.1002/<br />

jmor.1052250105.<br />

Fish, F.E., Howle, L.E., & Murray, M.M.<br />

2008. Hydrodynamic flow control in marine<br />

mammals. Integr Comp Biol. 211:1859-67.<br />

Fish, F.E., & Lauder, G.V. 2006. Passive<br />

<strong>and</strong> active flow control by swimming fishes<br />

<strong>and</strong> mammals. Annu Rev Fluid Mech.<br />

38:193-224. doi: 10.1146/annurev.<br />

fluid.38.050304.092201.<br />

Fish, F.E., & Rohr, J. 1999. Review of<br />

Dolphin Hydrodynamics <strong>and</strong> Swimming<br />

Performance. San Diego, CA: SPAWARS<br />

TR 1801.<br />

Fish, F.E., Weber, P.W., Murray, M.M., &<br />

Howle, L.E. 2011. The humpback whale’s<br />

flipper: Application of bio-inspired tubercle<br />

technology. Integr Comp Biol. 51:203-13.<br />

doi: 10.1093/icb/icr016.<br />

Friedlaender, A.S., Hazen, E.L., Nowacek,<br />

D.P., Halpin, P.N., Ware, C., Weinrich,<br />

M.T., … Wiley, D. 2009. Diel changes in<br />

humpback whale Megaptera novaeangliae<br />

feeding behavior in response to s<strong>and</strong> lance<br />

Ammodytes spp. behavior <strong>and</strong> distribution.<br />

Mar Ecol Prog Ser. 395:91-100. doi:<br />

10.3354/meps08003.<br />

Gibbs-Smith, C.H. 1962. Sir George<br />

Cayley’s Aeronautics 1796-1855. London:<br />

Her Majesty’s Stationery Office.<br />

Goruney, T., & Rockwell, D. 2009.<br />

Flow past a delta wing with a sinusoidal<br />

leading edge: Near-surface topology <strong>and</strong> flow<br />

structure. Exp Fluids. 47:321-31.<br />

doi: 10.1007/s00348-009-0666-x.<br />

Gray, J. 1936. Studies in animal locomotion<br />

VI. The propulsive powers of the dolphin.<br />

J Exp Biol. 13:192-9.<br />

Hansen, K.L., Kelso, R.M., & Dally, B.B.<br />

2009. The effect of leading edge tubercle<br />

geometry on the performance of different<br />

airfoils. In: 7th World Conf Exp Heat Transfer,<br />

Fluid Mechanics <strong>and</strong> Thermodynamics.<br />

28 June-03 July 2009. Krakow, Pol<strong>and</strong>.<br />

Hansen, K.L., Kelso, R.M., & Dally, B.B.<br />

2011. Performance variations of leading-edge<br />

tubercles for distinct airfoil profiles. AIAA<br />

J Aircraft. 49:185-94.<br />

Hansen, K.L., Kelso, R.M., & Doolan, C.J.<br />

2010. Reduction of flow induced tonal noise<br />

through leading edge tubercle modifications.<br />

In: 16th AIAA/CEAS Aeroacoustics Conf.<br />

7-9 June 2010. Stockholm, Sweden.<br />

Hazen, E.L., Friedlaender, A.S., Thompson,<br />

M.A., Ware, C.R., Weinrich, M.T., Halpin,<br />

P.N., & Wiley, D.N. 2009. Fine-scale<br />

prey aggregations <strong>and</strong> foraging ecology of<br />

humpback whales Megaptera novaeangliae.<br />

Mar Ecol Prog Ser. 395:75-89.<br />

doi: 10.3354/meps08108.<br />

Hoerner, S.F. 1965. Fluid-dynamic Drag.<br />

Published by author. Brick Town, NJ.<br />

Howle, L.E. 2009. WhalePower Wenvor<br />

blade. A report in the efficiency of a WhalePower<br />

Corp. 5 meter prototype wind turbine blade.<br />

BelleQuant Engineering, PLLC.<br />

Iosilevskii, G., & Weihs, D. 2007. Speed<br />

limits on swimming of fishes <strong>and</strong> cetaceans.<br />

J R Soc Interface. 5:329-38. doi: 10.1098/<br />

rsif.2007.1073.<br />

Johari, H., Henoch, C., Custodio, D., &<br />

Levshin, A. 2007. Effects of leading-edge<br />

protuberances on airfoil performance. AIAA<br />

J. 45:2634-42. doi: 10.2514/1.28497.<br />

Kramer, M.O. 1960. The dolphin’s secret.<br />

New Sci. 7:1118-20.<br />

Larrabee, E.E. 1980. The screw propeller.<br />

Sci Am. 243:134-48.<br />

Liu, P., & Bose, N. 1993. Propulsive performance<br />

of three naturally occurring oscillating<br />

propeller planforms. Ocean Eng. 20:57-75.<br />

doi: 10.1016/0029-8018(93)90046-K.<br />

Miklosovic, D.S., Murray, M.M., &<br />

Howle, L.E. 2007. Experimental evaluation of<br />

sinusoidal leading edges. J Aircraft. 44:1404-7.<br />

doi: 10.2514/1.30303.<br />

Miklosovic, D.S., Murray, M.M., Howle,<br />

L.E., & Fish, F.E. 2004. Leading-edge<br />

tubercles delay stall on humpback whale<br />

(Megaptera novaeangliae) flippers. Phys<br />

Fluids. 16:L39-42. doi: 10.1063/1.1688341.<br />

Murray, M.M., Fish, F.E., Howle, L.E., &<br />

Miklosovic, D.S. 2005. Stall delay by leading<br />

edge tubercles on humpback whale flipper<br />

at various sweep angles. In 14 th Int Symp<br />

Unmanned Untethered Submersible Tech.<br />

Durham New Hampshire: Autonomous<br />

Undersea Systems Insitute.<br />

Murray, M., Gruber, T., & Fredriksson, D.<br />

2010. Effect of leading edge tubercles on<br />

marine tidal turbine blades. 63 rd Ann Mtg<br />

APS Div Fluid Dyn, 55(16), Nov. 21, 2010,<br />

Long Beach, CA.<br />

Owen, J.C., Szewczyk, A.A., & Bearman,<br />

P.W. 2000. Suppression of Karman vortex<br />

shedding. Phys Fluids. 12:S9.<br />

Ozen, C.A., & Rockwell, D. 2010. Control<br />

of vortical structures on a flapping wing<br />

via a sinusoidal leading-edge. Phys Fluids.<br />

22:021701. doi: 10.1063/1.3304539.<br />

Pedro, H.T.C., & Kobayashi, M.H.<br />

2008. Numerical study of stall delay on<br />

humpback whale flippers. In: 46th AIAA<br />

Aerospace Mtg Exhibit. 7-10 Jan. 2008.<br />

Reno, Nevada.<br />

Peterson, C.J.G. 1925. The motion of<br />

whales during swimming. Nature. 116:327-9.<br />

doi: 10.1038/116327a0.<br />

Riley, J.J., Gad-el-Hak, M., & Metcalfe,<br />

R.W. 1988. Compliant coatings. Annu Rev<br />

Fluid Mech. 20:393-420. doi: 10.1146/<br />

annurev.fl.20.010188.002141.<br />

Saadat, M., Haj-Hariri, H., & Fish, F. 2010.<br />

Explanation of the effects of leading-edge<br />

206 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


tubercles on the aerodynamics of airfoils<br />

<strong>and</strong> finite wings. 63 rd Ann Mtg APS Div<br />

Fluid Dyn, 55(16), Nov 21, 2010, Long<br />

Beach, CA.<br />

Shevell, R.S. 1986. Aerodynamic anomalies:<br />

Can CFD prevent or correct them J Aircraft.<br />

23:641-9. doi: 10.2514/3.45356.<br />

Stanway, M.J. 2008. Hydrodynamic effects<br />

of leading-edge tubercles on control surfaces<br />

<strong>and</strong> in flapping foil propulsion. M.S. thesis,<br />

Massachusetts Institute of <strong>Technology</strong>.<br />

Triantafyllou, G.S., & Triantafyllou, M.S.<br />

1995. An efficient swimming machine.<br />

Sci Amer. 272:64-70. doi: 10.1038/<br />

scientificamerican0395-64.<br />

Van Nierop, E.A., Alben, S., & Brenner,<br />

M.P. 2008. How bumps on whale flippers<br />

delay stall: An aerodynamic model. Phys<br />

Rev Lett. 100:054502. doi: 10.1103/<br />

PhysRevLett.100.054502.<br />

Vogel, S. 1981. Life in Moving Fluids.<br />

Boston: Willard Grant Press.<br />

Weber, P.W., Howle, L.E., & Murray, M.M.<br />

2010. Lift, drag, <strong>and</strong> caviation onset on<br />

rudders with leading-edge tubercles. <strong>Marine</strong><br />

Technol Sname N. 47(1):26.<br />

Weber, P.W., Howle, L.E., Murray, M.M.,<br />

& Miklosovic, D.S. 2011. Computational<br />

evaluation of the performance of lifting<br />

surfaces with leading edge protuberances.<br />

J Aircraft. 48:591-600. doi: 10.2514/<br />

1.C031163.<br />

Wegener, P.P. 1991. What Makes Airplanes<br />

Fly New York: Springer-Verlag.<br />

Wind Energy Institute of Canada. 2008.<br />

WhalePower Tubercle Blade Power Performance<br />

Test Report. North Cape, PE:<br />

Wind Energy Institute of Canada.<br />

Wu, J.Z., Vakili, A.D., & Wu, J.M. 1991.<br />

Review of the physics of enhancing<br />

vortex lift by unsteady excitation.<br />

Prog Aerosp Sci. 28:73-131. doi: 10.1016/<br />

0376-0421(91)90001-K.<br />

July/August 2011 Volume 45 Number 4 207


PAPER<br />

Shark Skin Separation Control Mechanisms<br />

AUTHORS<br />

Amy Lang<br />

University of Alabama<br />

Philip Motta<br />

Maria Laura Habegger<br />

University of South Florida<br />

Robert Hueter<br />

Mote <strong>Marine</strong> Laboratory<br />

Farhana Afroz<br />

University of Alabama<br />

Introduction<br />

Natural selection, operating for<br />

hundreds of millions of years, has<br />

honed fast swimming marine organism<br />

forms to reduce energy expenditure<br />

through streamlining. Among<br />

the fastest swimming fishes, certain<br />

species of sharks have converged on<br />

a suite of adaptations to reduce drag.<br />

Consequently, fast swimming sharks<br />

have been studied for insight into potential<br />

drag-reducing mechanisms for<br />

well over three decades. Drag on a<br />

submerged, swimming body consists<br />

of three sources. These include (i) form<br />

drag due to a difference in pressure<br />

around the body, (ii) drag due to lift,<br />

<strong>and</strong> (iii) skin friction due to boundary<br />

layer formation (Bushnell & Moore,<br />

1991). At low Reynolds numbers<br />

(Re) skin friction predominates, while<br />

at higher Re pressure drag can dominate<br />

if not minimized. It is not surprising<br />

then to consider the fact that<br />

aquatic organisms have evolved to<br />

minimize drag (Fish, 1998), with<br />

the primary decrease coming from<br />

a streamlined body shape to reduce<br />

flow separation <strong>and</strong> thus form<br />

drag. Aquatic organisms that swim at<br />

ABSTRACT<br />

Drag reduction by marine organisms has undergone millions of years of natural<br />

selection, <strong>and</strong> from these organisms biomimetic studies can derive new technologies.<br />

The shortfin mako (Isurus oxyrinchus), considered to be one of the fastest <strong>and</strong><br />

most agile marine predators, is known to have highly flexible scales on certain locations<br />

of its body. This scale flexibility is theorized to provide a passive, flow-actuated<br />

mechanism for controlling flow separation <strong>and</strong> thereby decreasing drag. Recent biological<br />

observations have found that the shortfin mako has highly flexible scales,<br />

bristling to angles in excess of 50°, particularly on the sides of the body downstream<br />

of the gills. High “contragility,” which is explicitly definedhereastheabilityto<br />

change or move in a new or opposing direction while already in a turn, would<br />

occur if form drag were minimized. This would thus indicate the potential control<br />

of flow separation on body regions aft of the point of maximum girth or in regions of<br />

adverse pressure gradient. Thus results are consistent with the hypothesis that<br />

scale bristling controls flow separation. This scale flexibility appears to be a result<br />

of a reduction in the relative size of the base of the scales as well as a reorganization<br />

of the base shape as evidenced by histological examination of the<br />

skin <strong>and</strong> scales. Probable mechanisms leading to separation control are discussed.<br />

Keywords: shark skin, flow separation, drag reduction<br />

high Re (>10 3 )haveavarietyof<br />

shapes <strong>and</strong> structures to reduce drag,<br />

which we often attempt to duplicate<br />

(Vogel, 2003; Fish & Lauder, 2006).<br />

It has been deduced that a crescent tail<br />

design could decrease induced drag on<br />

the order of 8%, <strong>and</strong> not surprisingly<br />

this lunate tail design is found on<br />

many fast swimming marine animals<br />

(Fish, 1998; Donley et al., 2004).<br />

Several researchers (e.g., Anderson<br />

et al., 2001) have observed that swimming<br />

fish experience more friction<br />

drag than the same rigid body<br />

towed. This higher friction is attributed<br />

to motion of the body as it<br />

swims to produce thrust. The undulating<br />

body motion can result in measurably<br />

thinner boundary layers <strong>and</strong><br />

thus higher skin friction (Fish, 2006).<br />

The argument has been made that<br />

because of this higher power output<br />

requirement, to overcome drag <strong>and</strong><br />

maintain a certain speed, swimming<br />

drag reduction due to various morphological<br />

mechanisms is extremely<br />

probable (Schultz & Webb, 2002).<br />

Separation of the boundary layer<br />

from a body typically occurs in vicinities<br />

where the flow is decelerating<br />

along a curved body after the point of<br />

maximum thickness, resulting in an<br />

adverse pressure gradient. As a result<br />

separation typically occurs in areas<br />

posterior of the maximum body thickness.<br />

Incipient separation is characterized<br />

by regions of decreasing skin<br />

friction approaching zero, <strong>and</strong> consequent<br />

reversal of the flow at the surface<br />

(Doligalskietal.,1994).Swimming<br />

kinematics in thunniform fish such as<br />

tunas <strong>and</strong> mako sharks are characterized<br />

by cyclically repeating motions <strong>and</strong><br />

small linear <strong>and</strong> angular accelerations<br />

(Blake, 2004). Most fast swimming<br />

sharks, such as the shortfin mako Isurus<br />

208 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


oxyrinchus, are thunniform swimmers<br />

where oscillations are for the most<br />

part limited to the posterior end of<br />

the body. It has been reported that incipient<br />

separation (inflected boundary<br />

layer profile) is often observed during<br />

swimming movements, <strong>and</strong> this motion<br />

may be tuned by the fish to take<br />

advantage of the lower shear stress in<br />

a nearly separating boundary layer;<br />

yet separation must be avoided to reduce<br />

form drag <strong>and</strong> increase thrust<br />

produced by the caudal fin (Anderson<br />

et al., 2001).<br />

The skin of sharks is covered by<br />

minute scales, called dermal denticles<br />

or placoid scales, which originally<br />

evolved as a hard, protective covering<br />

to the animal (Raschi & Tabit,<br />

1992). The bases of these hard scales<br />

are embedded in the superficial collagenous<br />

layer of the skin (dermis) termed<br />

the stratum laxum, with the crowns<br />

of the scales exposed to the water. It is<br />

the unique geometry observed on fast<br />

swimming sharks for these tooth-like<br />

scales that is of interest from a drag reduction<br />

st<strong>and</strong>point. Beginning in the<br />

1970s researchers became interested<br />

in the small, streamwise ridges,<br />

or keels, located on the top of each<br />

crown (see scanning electron microscopy<br />

(SEM) pictures of shark scales<br />

shown in Figure 1). Now labeled as<br />

riblets, these ridges were found to result<br />

in a reduction in turbulent skin<br />

friction drag of up to 9.9% when<br />

sized correctly (Bechert et al., 1997).<br />

Even as early as the 1980s, riblets<br />

were utilized on boat hulls competing<br />

in the Olympics <strong>and</strong> America’s Cup<br />

but later were banned (Gad-el Hak,<br />

2000).<br />

However, the scales on some fast<br />

swimming sharks exhibit a different<br />

property that is flexibility or capability<br />

to bristle, <strong>and</strong> this is the focus of our<br />

current work. Previous work (Bechert<br />

FIGURE 1<br />

The average scale erection angles for 16 regions on the shortfin mako shark (Isurus oxyrinchus)<br />

together with scanning electron micrographs (top row) <strong>and</strong> histological sections (middle row) of<br />

the skin for three regions on the body. The flank region, including B2, has the most flexible scales,<br />

which are characterized by long backwardly projecting crowns <strong>and</strong> relatively short bases. The scale<br />

bases are embedded in the superficial part of the dermis. The surface of the scales has three riblets<br />

visible on the scanning electron micrographs. The scales in any region are oriented along the longitudinal<br />

axis of the shark, <strong>and</strong> anterior is to the left. The erection angles noted on the figure are the<br />

angles to which the scales in that region can be manually manipulated without damage to their<br />

attachment <strong>and</strong> which remain at that angle after release by the needle used to erect them.<br />

et al., 2000) investigating this aspect of<br />

the shark skin found no advantage<br />

from a skin friction reduction st<strong>and</strong>point,<br />

<strong>and</strong> only greatly increased friction<br />

drag if the scales were allowed to<br />

remain bristled. Thus, a new, passive<br />

flow-actuated separation control<br />

mechanism is proposed that is inspired<br />

by the scale flexibility found on the<br />

shortfin mako shark.<br />

This investigation chose to focus<br />

on the skin of the shortfin mako<br />

based on several factors. First, of fast<br />

swimming pelagic sharks, the shortfin<br />

mako is considered by most to be the<br />

fastest <strong>and</strong> most agile (Stevens, 2009).<br />

It is also one of the more derived<br />

species of shark <strong>and</strong> is recognized as<br />

making its appearance roughly 55 million<br />

years ago in the line of shark<br />

evolution dating back more than<br />

400 million years (Naylor et al.,<br />

1997). Next, it is one of two species<br />

of shark previously reported in literature,<br />

the other being the smooth hammerhead<br />

Sphyrna zygaena, ashaving<br />

July/August 2011 Volume 45 Number 4 209


flexible scales over large portions of its<br />

body (Bruse et al., 1993). Finally, the<br />

shortfin mako is readily obtainable off<br />

the Atlantic coast of the United States<br />

<strong>and</strong> is not currently listed as overfished<br />

or otherwise in a vulnerable state<br />

for conservation purposes (NMFS,<br />

2010).<br />

Materials <strong>and</strong> Methods<br />

To investigate scale structure <strong>and</strong><br />

erection, we acquired two subadult<br />

shortfin makos (female: total length,<br />

192 cm; fork length, 171.5 cm; male:<br />

total length, 158 cm; fork length,<br />

150 cm) from commercial <strong>and</strong> recreational<br />

fishers in the coastal waters off<br />

Montauk, New York. The frozen<br />

specimens were shipped to the University<br />

of South Florida in Tampa. There,<br />

following Reif (1985), scales at 16 regions<br />

along the body were marked in<br />

order to sample the scales under a<br />

variety of flow regimes (Figure 1).<br />

Three 1 cm 2 samples from each location<br />

were removed, two for histological<br />

analysis <strong>and</strong> one for SEM. Because<br />

swimming sharks have superambient<br />

subcutaneous pressure ranging as<br />

high as 100-200 kPa increasing skin<br />

stiffness (Wainwright et al., 1978;<br />

Martinez et al., 2002), scale erection<br />

angles were recorded with <strong>and</strong> without<br />

subcutaneous pressure, as follows.<br />

We measured scales erection angle<br />

under a dissecting microscope at a<br />

magnification of 135×. In seven of<br />

the 16 regions (B1, B2, B4, B5, A1,<br />

A2, A3; Figure 1), an aneroid sphygmomanometer<br />

was placed under the<br />

skin <strong>and</strong> underlying muscle <strong>and</strong> the<br />

pressure elevated <strong>and</strong> held at 15 psi<br />

(103 kPa), the maximum possible<br />

without damaging the underlying<br />

muscle tissue. Scale erection was not<br />

noted with the increase in subcutaneous<br />

pressure, leading us to suspect a<br />

passive mechanism. While the skin<br />

was pressurized, we used a fine acupuncture<br />

needle to gently manipulate<br />

five haphazardly selected scales to<br />

their maximum erected position without<br />

tearing them from the skin. Releasing<br />

the scales, we allowed them to<br />

settle at an erected angle, which we<br />

then measured by calculating the<br />

change in length of the scale crown<br />

when viewed from above. The inverse<br />

cosine of the apparent crown length divided<br />

by the resting or true crown<br />

length provided the angle of erection.<br />

Because the individual scales could actually<br />

be easily erected past this resting<br />

angle, we were in essence calculating a<br />

minimal erection angle. The pressure<br />

was then released <strong>and</strong> the angle similarly<br />

calculated on the flaccid skin.<br />

Our aprioritest determined that<br />

stretching the skin in this manner did<br />

not affect the non-pressurized erection<br />

angle.<br />

At the other regions (H2, B3, B6,<br />

P1, P2, P3, C1, C2, C3; Figure 1)<br />

where subcutaneous pressure was not<br />

possible, such as in the fins <strong>and</strong> over<br />

the body cavity, the erection angles<br />

were measured by gently erecting the<br />

scales (if possible) with the acupuncture<br />

needle <strong>and</strong> measuring their<br />

pre- <strong>and</strong> post-erection crown length.<br />

Finally, to get a more global picture<br />

of scale flexibility over the entire<br />

body, 35 equidistant sampling locations<br />

encompassing the entire dorsal,<br />

left lateral, <strong>and</strong> ventral surfaces of<br />

each shark were marked <strong>and</strong> scales in<br />

each area manually erected as before<br />

without subcutaneous pressure. The<br />

erected scales were simply recorded as<br />

greater or less than 50°, an angle that<br />

appeared to the approximate maximum.<br />

We also measured the crown<br />

<strong>and</strong> base length of scales in select<br />

areas, as well as the spacing of the riblets<br />

on the scales.<br />

To underst<strong>and</strong> the attachment of<br />

the scales to the skin, we prepared histological<br />

sections of the skin <strong>and</strong> scales,<br />

decalcified the scales, stained the samples<br />

to reveal the fibrous attachment,<br />

<strong>and</strong> examined the sections at 20× <strong>and</strong><br />

40× with a compound microscope. Finally,<br />

surface pictures of the scales, at<br />

all studied regions, were prepared by<br />

examining the skin at 100× <strong>and</strong> 200×<br />

under a SEM.<br />

Findings<br />

The placoid scales of sharks have<br />

a pulp cavity <strong>and</strong> a hard enameloid<br />

covering over dentine <strong>and</strong> are anchoredatthebaseofthescaletothe<br />

stratum laxum collagenous layer of<br />

the dermis (Figures 1 <strong>and</strong> 2). The exposed<br />

crowns overlap each other on<br />

the shark’s surface, <strong>and</strong> the majority<br />

of scales have small riblets or keels on<br />

their surface, oriented in the streamwise<br />

direction of the flow (Figures 1<br />

<strong>and</strong> 2). On the fast swimming shortfin<br />

mako, the flank scales (e.g., areas B2,<br />

B5, <strong>and</strong> A2) have a crown length of approximately<br />

0.18 mm, <strong>and</strong> each crown<br />

typically has three keels each having a<br />

height of 0.012 mm <strong>and</strong> a spacing of<br />

0.041 mm. This differs from other<br />

slower swimming sharks such as the<br />

blacktip shark (Carcharhinus limbatus),<br />

whereby preliminary data indicate the<br />

flank scales are typically 0.32 mm<br />

in length with each crown typically<br />

having five keels with a height of<br />

0.029 mm <strong>and</strong> a spacing of 0.065 mm.<br />

When considering shark species as a<br />

whole, length of the scales is typically<br />

fixed for specific regionsofthebody<br />

within a species but differs among regions<br />

<strong>and</strong> species. Similarly, the number<br />

of keels per scale is also consistent<br />

per body location for a species.<br />

Scale flexibility on the shortfin<br />

mako varies considerably across the<br />

210 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


FIGURE 2<br />

Scanning electron micrographs of the scales <strong>and</strong> histological sections of the embedded scales<br />

from three regions of the pectoral fin, P1, P2, <strong>and</strong> P3 of the shortfin mako (Isurus oxyrinchus).<br />

The scales on the leading edge of the fin cannot erect <strong>and</strong> lack riblets, whereas those on the trailing<br />

edge can erect to greater angles than either the scales on the leading edge or the midregion of the<br />

fin. The bases of the scales (B) are anchored in the dermis (D), <strong>and</strong> the thin epidermis (E) is visible<br />

between the scales. The pulp cavity (PC) is visible in some of the scales, <strong>and</strong> not all scales are<br />

sectioned through their center, resulting in some crown (CR) lengths appearing shorter than<br />

others. Ceratotrichia or fin rays (CE) are visible in region P3 because this part of the fin is very<br />

thin. Anterior is to the left.<br />

body <strong>and</strong> fins, with average erection<br />

angles varying from 0° on the leading<br />

edge of the pectoral fin to approximately<br />

50° on the widest part of the<br />

body just behind the gill region (Figure<br />

3). However, only certain portions<br />

ofthebodyhaveveryflexible scales.<br />

The most flexible scales are found on<br />

the flank of the body extending behind<br />

the gills to the tail; here scales<br />

are found to be easily erected with slight<br />

manipulation on dead specimens to angles<br />

of approximately 50° or greater<br />

(Figure 3). The lateral scales (B2, B5,<br />

A2) had significantly greater erection<br />

angles (mean angle = 44° ± 1°) than<br />

both the dorsal (mean angle = 26° ±<br />

1° SE) <strong>and</strong> ventral regions (mean<br />

angle = 25° ± 2° SE). Erection angles<br />

for the dorsal region did not differ<br />

from that of the ventral region<br />

(Kruskal-Wallis one-way ANOVA,<br />

Tukey’s pairwise test; H = 53.173,<br />

FIGURE 3<br />

Outline of a representative shortfinmako,Isurus<br />

oxyrinchus, showing the approximate region<br />

(lines) on the flank with the most flexible scales<br />

capable of erection to at least 50°. This region<br />

approximates the region of flow separation.<br />

df =2,P ≤ 0.001). Highly flexible<br />

scales are also found at the trailing<br />

edge of the pectoral fins. The scales<br />

on the trailing edge of the pectoral<br />

(P3) had the highest erection angles<br />

compared to the leading edge (P1)<br />

<strong>and</strong> the central region of the fin (P2),<br />

<strong>and</strong> all the regions were significantly<br />

different from each other (Kruskal-<br />

Wallis one-way ANOVA, Tukey’s<br />

pairwise test; H = 26.289, df =2,<br />

P ≤ 0.001). Conversely, for the caudal<br />

fin the mean angles were not significantly<br />

different among the three regions<br />

(ANOVA; F = 0.0614, df =2,<br />

P = 0.941).<br />

At least two factors appear to control<br />

scale flexibility on the body. The<br />

first is a reduction in the length of<br />

the base relative to the length of the<br />

crown over certain regions of the<br />

body such as the flank in the shortfin<br />

mako (Table 1). The flank scales<br />

have a greater ratio of crown length<br />

to base length compared to the dorsal<br />

scales (Kruskal-Wallis one-way<br />

ANOVA, Tukey’s pairwisetest;H =<br />

18.104, df =2,P < 0.001) due to a significantly<br />

shorter base on the flank<br />

scales (ANOVA, Holm-Sidak method;<br />

F = 45.967, df =2,P =0.001).The<br />

flank scales have a relatively wide base<br />

compared to its length, whereas the<br />

base of the dorsal scales is more uniform<br />

in shape (Table 1, BL/BW ratio;<br />

Figure 4). The ventral scales of the<br />

shortfin mako are smaller than the<br />

dorsal <strong>and</strong> flank scales as they are<br />

shorter in crown (ANOVA, Holm-<br />

Sidak method; F =45.967,df =2,<br />

P < 0.001) <strong>and</strong> base length (ANOVA,<br />

Holm-Sidak method; F = 42.916,<br />

df =2,P < 0.001) overall. Secondly, relative<br />

changes in the length of the leading<br />

<strong>and</strong> trailing edges of the scale base<br />

affect its anchoring in the dermis, similar<br />

to the root system of a tree. More<br />

firmly anchored scales have a more<br />

July/August 2011 Volume 45 Number 4 211


TABLE 1<br />

Means <strong>and</strong> st<strong>and</strong>ard errors for scale crown length (CL), base length (BL), base width (BW), ratio of CL/BL, <strong>and</strong> ratio BL/BW for three body areas of<br />

shortfin mako (Isurus oxyrinchus).<br />

Body Position CL (mm) BL (mm) BW (mm) CL/BL BL/BW<br />

B1, B4, A1 Dorsum 0.173 ± 0.004 0.145 ± 0.005 0.161 ± 0.005 1.201 ± 0.029 0.938<br />

B2, B5, A2 Flank 0.179 ± 0.004 0.104 ± 0.003 0.161 ± 0.005 1.745 ± 0.076 0.625<br />

B3, B6, A3 Ventrum 0.128 ± 0.004 0.091 ± 0.005 0.125 ± 0.003 1.472 ± 0.102 0.692<br />

evenly distributed base like a tree with<br />

a broad but evenly distributed shallow<br />

root system (Figure 4B). The more<br />

erectable scales on the flank have a narrow<br />

base on the trailing edge (Figure 4A).<br />

In this manner, the scales can pivot up,<br />

analogous to a tree with a narrow root<br />

system on one side that is blown over<br />

away from this side. The scales return<br />

to their resting position with the help<br />

of elastic fibers that anchor its base<br />

(stained black in the histology sections).<br />

A sideview picture of the skin with the<br />

scales in the foreground manually<br />

bristled is shown in Figure 5.<br />

Effects on Fluid<br />

Flow Patterns<br />

The findings of scale flexibility <strong>and</strong><br />

angle of erection on the shortfin mako<br />

FIGURE 4<br />

Representative scales of the shortfin mako<br />

(Isurus oxyrinchus) from(A)theflexible<br />

flankareaB5,<strong>and</strong>(B)thelessflexible area<br />

B4. The scales in (A) have a relatively short<br />

but wide base compared to those of (B) with<br />

a more uniformly shaped base.<br />

have led to a working hypothesis currently<br />

being investigated through<br />

hydrodynamic testing. Our discovery<br />

of highly flexible scales on the flank<br />

<strong>and</strong> trailing edges of the pectoral fin<br />

fits with the hypothesis that the scales<br />

act as a means of controlling flow separation.<br />

First, these results indicate that<br />

the erection of the scales is most likely<br />

initiated by a passive, flow-actuated<br />

mechanism, i.e., in a region of turbulent<br />

flow separation the flow consists of<br />

moments when the flow close to the<br />

wall is both in the main direction as<br />

well as reversed. Pressurizing the skin<br />

had no effect on scale erection. Flow<br />

reversal over a large region indicates<br />

FIGURE 5<br />

Coronal section through the shortfin mako<br />

skin showing the scales in the foreground<br />

that have been manually erected from location<br />

B2. Not all scales are erected to the same<br />

degree because of the individual manual erection.<br />

Flow would normally pass over the skin<br />

from left to right <strong>and</strong> reversed flow, as occurs<br />

during separation, is believed to cause bristling<br />

as shown.<br />

that global separation from the surface<br />

(body) occurs, resulting in increased<br />

pressure drag. In the case of the shark<br />

this would not only increase drag but<br />

also inhibit contragility or the ability<br />

to change direction quickly <strong>and</strong> to a<br />

large degree. As the shark swims <strong>and</strong><br />

turns, the body is bent laterally with<br />

regions of greatest curvature occurring<br />

on the flank. From the nose to<br />

the point of maximum girth (around<br />

the location of the gills), the flow will<br />

experience a favorable pressure gradient,<br />

<strong>and</strong> unfavorable pressure gradient<br />

regions will be located downstream of<br />

this point. Thus, the findings that the<br />

most flexible scales are found on the<br />

flank of the body <strong>and</strong> downstream of<br />

the gills corroborate the hypothesis<br />

that these scales are working to control<br />

flow separation. Likewise, the delay of<br />

flow separation over the pectoral fin<br />

can lead to increased performance in<br />

their ability to act as lifting surfaces<br />

during high-speed swimming maneuvers.<br />

The generation of high lift<br />

by the pectoral fins is important for<br />

quick upward maneuvers while attacking<br />

prey, as has been observed in video<br />

evidence of a shortfin mako in pursuit<br />

of a towed baitfish. This same video<br />

evidence shows the shark’s ability to<br />

turn in one direction <strong>and</strong> then change<br />

direction before the body completes<br />

the initial turn. This type of turning<br />

behavior, also defined as contragility,<br />

requires not only large muscular effort<br />

but also low form drag (Frank Fish,<br />

212 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


personal communication, February<br />

18, 2008).<br />

Flow visualization images (Figure<br />

6) of particles illuminated with a<br />

laser sheet show characteristic flow scenarios<br />

found in a turbulent boundary<br />

layer undergoing separation. In this<br />

case, separation is induced on a flat<br />

surface via the presence of a rotating<br />

cylinder located above the wall on<br />

which a turbulent boundary layer is<br />

formed. The free stream flow moves<br />

FIGURE 6<br />

Flow visualization generated by particle streaking<br />

in a plane parallel to the free stream flow<br />

(16.5 cm/s) illuminated by a laser sheet. A rotating<br />

(40 RPM) cylinder (5.1 cm diameter) is<br />

located above the image with its center, 6.4 cm<br />

above the wall, to induce boundary layer separation.<br />

Boundary layer thickness prior to the<br />

test area was approximately 1 cm <strong>and</strong> an area<br />

of approximately 3 cm × 1.5 cm is imaged.<br />

Main flow moves left to right. Distinguishing<br />

moments in time for this unsteady flow are<br />

shown when (a) the flow is attached, (b) the<br />

separation process is initiated with reversed<br />

flow close to wall, <strong>and</strong> (c) separated flow characterizedbyvortexburstingawayfromthe<br />

wall is observed, with a large region of reversed<br />

flow near the wall.<br />

at 17 cm/s, giving a local Reynolds<br />

number, based on distance from the<br />

leading edge of 0.3 m, in the boundary<br />

layer of 5 × 10 4 . In comparison, a<br />

shark’s boundary layer will have a Re of<br />

∼10 6 -10 7 when swimming at 10 m/s.<br />

However, the characteristics of the turbulent<br />

flow will be similar; the flow on<br />

a shark will be faster <strong>and</strong> the boundary<br />

layer thinner resulting in shorter time<br />

<strong>and</strong> length scales than the water tunnel<br />

experiments. Under these conditions,<br />

there are moments when the flow is<br />

for the most part attached (Figure 6a),<br />

developing into a large region of separation<br />

with reversed flow near the wall<br />

(Figure 6b), <strong>and</strong> finally times when<br />

large vortex bursting occurs as the separated<br />

region becomes unstable resulting<br />

in a shedding of vortices (Figure 6c).<br />

A time trace of the velocity measured at<br />

a location adjacent to the wall where<br />

the flow is reversed about 50% of the<br />

timeisshowninFigure7.Here,the<br />

cyclic nature of a separating turbulent<br />

boundary is clearly evident. Regions in<br />

FIGURE 7<br />

theplotwherethevelocityisnegativeindicate<br />

moments of reversed flow,<br />

which would actuate scale bristling<br />

on the shark. It is transitions in the<br />

time trace of the velocity where the<br />

flow moves (Figure 7) from positive<br />

(u > 0) to negative (u < 0) (or from<br />

point a to point b labeled in Figure 7)<br />

when scale actuation would be initiated<br />

<strong>and</strong> potentially disrupt the evolving<br />

flow that leads to the formation<br />

of a separation bubble as shown in<br />

Figures 6b <strong>and</strong> 6c.<br />

Cassel et al. (1996) provide a description<br />

of the process leading to<br />

flow separation. Because of the suction<br />

pressure upstream (adverse pressure<br />

gradient) the region closest to<br />

the wall, where the flow has the lowest<br />

momentum, is where flow reversal<br />

is first initiated. This patch of fluid<br />

moves upstream <strong>and</strong> thickens, ultimately<br />

leading to large scale flow separation<br />

from the surface. It is our<br />

hypothesis that on the shark’s body<br />

in the region close to the wall where<br />

Velocity (u) measured as a function of time at a point close to the wall corresponding to approximately<br />

two-thirds the distance downstream in Figure 6. Measurements were made using timeresolved<br />

digital particle image velocimetry (TR-DPIV), which sampled the flow at 1 kHz. At this<br />

location for about 50% of the time, flow reversal is occurring, as can be seen by the regions where<br />

u < 0. This is a point in the vicinity where flow reversal is initiated in the turbulent boundary layer<br />

<strong>and</strong> develops into a large-scale region of reversed flow due to the presence of the adverse pressure<br />

gradient induced by the rotating cylinder above. Points labeled (a), (b), <strong>and</strong> (c) are<br />

moments in the flow typified correspondingly to those shown in Figure 6.<br />

July/August 2011 Volume 45 Number 4 213


flow reversal begins to occur, the scales<br />

are actuated by the flow to erect, thereby<br />

disrupting the unsteady flow separation<br />

process. Future experiments<br />

will investigate this hypothesis through<br />

hydrodynamic testing using models<br />

<strong>and</strong> real specimens of shark skin.<br />

This aspect of shark skin resulting in<br />

asurfacewithapreferredflow direction<br />

is likely key to its ability to control<br />

flow separation.<br />

Previousexperimentsoverabristled<br />

shark skin model confirmed the<br />

presence of embedded vortices forming<br />

between replicas of the scales<br />

(Lang et al., 2008). Thus, if flow is<br />

induced to form between the scales<br />

when bristled, there are two additional<br />

mechanisms that may aid to control<br />

the flow. The formation of embedded<br />

vortices, similar as occurs with dimples<br />

on a golf ball, would allow the flow to<br />

pass over the skin with a resulting partial<br />

slip condition, thereby leading to<br />

higher momentum adjacent to the surface.<br />

Secondly, with a turbulent flow<br />

forming in the boundary layer above<br />

the cavities, there may be additional<br />

momentum exchange whereby high<br />

momentum fluid typically located<br />

away from the surface is induced at a<br />

greater rate to move towards the surface<br />

<strong>and</strong> into the cavities. This latter<br />

mechanism, resulting in turbulence<br />

augmentation (Gad el-Hak, 2000), is<br />

another potential means to increase<br />

the momentum overall in the flow adjacent<br />

to the wall. These three mechanisms<br />

may be working in conjunction<br />

to inhibit flow separation over the<br />

surface of the shark.<br />

This method has advantages, above<br />

<strong>and</strong> beyond other methods currently<br />

in use to control flow separation, in<br />

that it is passive with no energy input<br />

required. Also it causes no additional<br />

drag penalty when not in use in that<br />

scale bristling would be controlled by<br />

on-dem<strong>and</strong> erection of the scales induced<br />

by regions of flow reversal as<br />

occurs under conditions of incipient<br />

flow separation. This obviates the use<br />

of bristled scales to act as vortex generators<br />

as a means of separation control,<br />

as previously theorized by Bechert<br />

et al. (2000). Vortex generators, which<br />

consist of small, typically V-shaped<br />

protrusions, require careful placement<br />

<strong>and</strong> protrusion into the boundary layer<br />

flow upstream of the point of flow separation<br />

<strong>and</strong> work by mixing higher<br />

momentum flow down towards the<br />

surface (Lin, 2002). Vortex generators<br />

also result in a drag penalty due to their<br />

protrusion into the flow (Gad-el-Hak,<br />

2000). Our findings suggest that the<br />

scales are bristled passively <strong>and</strong> are activated<br />

in a region of flow reversal that<br />

occurs downstream of the point of separation<br />

<strong>and</strong> is thus a different methodology<br />

from that of vortex generators<br />

currently in use today. Finally, this<br />

new passive, flow-actuated mechanism<br />

may in fact go to the initial root cause<br />

of the separation, that of flow reversal<br />

adjacent to the wall, <strong>and</strong> disrupt it<br />

prior to growth into a fully separated<br />

region. Our ultimate aim is to gain a<br />

fundamental underst<strong>and</strong>ing of how<br />

the flow-actuated bristling of shark<br />

skin scales can control flow separation<br />

so that bio-inspired surfaces can be<br />

engineered for greater flow control in<br />

marine <strong>and</strong> other applications.<br />

Summary<br />

Fast swimming shortfin mako<br />

sharks Isurus oxyrinchus have highscale<br />

flexibility on the flank <strong>and</strong><br />

trailing edges of the pectoral fin, with<br />

bristling angles up to a range of approximately<br />

50° on the flank. These regions<br />

correspond to those on the body<br />

where flow separation control is likely<br />

to be most beneficial. In the case of the<br />

flank region, high curvature of the<br />

body will result from the shark’s lateral<br />

swimming motion, <strong>and</strong> this is also<br />

in the vicinity of the point of maximum<br />

girth; both conditions indicate<br />

the presence of an adverse pressure<br />

gradient. In the case of the pectoral<br />

fin, it is hypothesized that the flexible<br />

scales can lead to the control of flow<br />

separation which indicates high drag<br />

<strong>and</strong> loss of lift. Control of the flow in<br />

both regions will lead to increased swimming<br />

speeds with high contragility for<br />

the shark. Increased scale flexibility in<br />

this shark appears to be due to a reduction<br />

in the relative size of the scale base<br />

compared to the crown <strong>and</strong> changes<br />

in the shape of the base where it is anchored<br />

into the dermis. Future work<br />

will lead to the manufacturing of bioinspired<br />

surfaces based on shark skin<br />

microgeometry whereby a passive,<br />

flow-actuated surface patterning can<br />

be used for applications where flow<br />

separation control is required.<br />

Acknowledgments<br />

Funding for this work received<br />

through collaborative NSF grants<br />

(0932352, 0744670, <strong>and</strong> 0931787)<br />

to A. Lang, P. Motta, <strong>and</strong> R. Hueter<br />

to support both the engineering <strong>and</strong><br />

biological work is gratefully acknowledged.<br />

We also thank Jessica Davis<br />

for assisting in the shark measurements<br />

<strong>and</strong> C<strong>and</strong>y Mir<strong>and</strong>a for preparing the<br />

histological samples. Finally, we wish<br />

to express our gratitude to Paul <strong>and</strong><br />

Jane Majeski <strong>and</strong> crew, Captain Mark<br />

Sampson, Captain Al VanWormer,<br />

Philip Pegley, Lisa Natanson, Jack<br />

Morris, <strong>and</strong> Mote <strong>Marine</strong> Laboratory<br />

for assistance in obtaining shark<br />

specimens.<br />

214 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Lead Author:<br />

Amy Lang<br />

Department of Aerospace<br />

Engineering <strong>and</strong> Mechanics<br />

University of Alabama<br />

Box 870280<br />

Tuscaloosa, AL 35487<br />

Email: alang@eng.ua.edu<br />

References<br />

Anderson, E., McGillis, W., & Grosenbaugh,<br />

M. 2001. The boundary layer of swimming<br />

fish. J Exp Biol. 204:81-102.<br />

Bechert, D.W., Bruse, M., Hage, W., &<br />

Meyer, R. 2000. Fluid mechanics of biological<br />

surfaces <strong>and</strong> their technological application.<br />

Naturwissenschaften. 80:157-71.<br />

Bechert, D.W., Bruse, M., Hage, W.,<br />

Van der Hoeven, J., & Hoppe, G. 1997.<br />

Experiments on drag-reducing surfaces <strong>and</strong><br />

their optimization with an adjustable<br />

geometry. J Fluid Mech. 338:59-87.<br />

Blake, R. 2004. Fish functional design <strong>and</strong><br />

swimming performance. J Fish Biol. 65:1193-222.<br />

Bruse, M., Bechert, D., van der Hoeven, J.,<br />

Hage, W., & Hoppe, G. 1993. Experiments<br />

with conventional <strong>and</strong> with novel adjustable<br />

drag-reducing surfaces. In: Proc. of the Int.<br />

Cong. On Near-Wall Turbulent Flows.<br />

719-738. Tempe, AZ: Elsevier Science<br />

Publishers.<br />

Bushnell, D., & Moore, K. 1991. Drag<br />

reduction in nature. Annu Rev Fluid Mech.<br />

23:65-79.<br />

Cassel, K., Smith, F., & Walker, J. 1996. The<br />

onset of instability in unsteady boundary-layer<br />

separation. J Fluid Mech. 315:223-56.<br />

Doligalski, T., Smith, C., & Walker, J. 1994.<br />

Vortex interactions with walls. Annu Rev<br />

Fluid Mech. 26:573-616.<br />

Donley, J.M., Sepulveda, C.A., Konstantinidis,<br />

P., Gemballa, S., & Shadwick, R.E. 2004.<br />

Convergent evolution in mechanical design of<br />

lamnid sharks <strong>and</strong> tunas. Nature. 429:61-65.<br />

Fish, F.E. 1998. Imaginative solutions by<br />

marine organisms for drag reduction. In:<br />

Proceedings of the International Symposium<br />

on Seawater Drag Reduction, ed. Meng, J.C.S.,<br />

443-50. Newport, Rhode Isl<strong>and</strong>.<br />

Fish, F. 2006. The myth <strong>and</strong> reality of<br />

Gray’s paradox: Implication of dolphin drag<br />

reduction for technology. Bioinspir Biomim.<br />

1:17-25.<br />

Fish, F., & Lauder, G. 2006. Passive <strong>and</strong><br />

active flow control by swimming fishes <strong>and</strong><br />

mammals. Annu Rev Fluid Mech. 38:193-224.<br />

Gad-el-Hak, M. 2000. Flow Control: Passive,<br />

Active <strong>and</strong> Reactive Flow Management.<br />

Cambridge, UK: Cambridge University<br />

Press. 421 pp.<br />

Lang, A., Hidalgo, P., Motta, P., & Westcott,<br />

M. 2008. Bristled shark skin: A microgeometry<br />

for boundary layer control<br />

Bioinspir Biomim. 3:046005.<br />

Lin, J. 2002. Review of research on lowprofile<br />

vortex generators to control boundarylayer<br />

separation. Prog Aerosp Sci. 38:389-420.<br />

Martinez, G., Drucker, E., & Summers, A.<br />

2002. Under pressure to swim fast. Integr<br />

Comp Biol. 42(6):1273-4.<br />

Naylor, G., Martin, A., Mattison, E., & Brown,<br />

W. 1997. The inter-relationships of lamniform<br />

sharks: Testing phylogenetic hypotheses with<br />

sequence data. In: Molecular Systematics of<br />

Fishes, eds. Kocher, T.D., & Stepien, C.,<br />

199-217. New York: Academic Press.<br />

NMFS. 2010. Final Amendment 3 to the<br />

Consolidated Atlantic Highly Migratory<br />

Species Fishery Management Plan. Silver<br />

Spring, MD: National Oceanic <strong>and</strong> Atmospheric<br />

Administration, National <strong>Marine</strong><br />

Fisheries Service, Office of Sustainable Fisheries,<br />

Highly Migratory Species Management<br />

Division. 632 pp. Public Document.<br />

Raschi, W., & Tabit, C. 1992. Functional<br />

aspects of placoid scales: A review <strong>and</strong> update.<br />

Aust J Mar Fresh Res. 43:123-147.<br />

Reif, W. 1985. Squamation <strong>and</strong> ecology of<br />

sharks, no. 78, Courier Forschungs-Institut<br />

Senckenberg, Frankfurt am Main.<br />

Schultz, W., & Webb, P. 2002. Power<br />

requirements of swimming: Do new methods<br />

resolve old questions Integr Comp Biol.<br />

42:1018-25.<br />

Stevens, J. 2009. The biology <strong>and</strong> ecology of<br />

the shortfin mako shark, Isurus oxyrinchus.<br />

In: Sharks of the Open Ocean: Biology,<br />

Fisheries <strong>and</strong> Conservation, eds. Camhi,<br />

M.D., Pikitch, E.K., Babcock, E.A., 87-94.<br />

Oxford, UK: Blackwell Publishing Ltd.<br />

Vogel, S. 2003. Comparative Biomechanics:<br />

Life’s Physical World. Princeton: Princeton<br />

University Press.<br />

Wainwright, S., Vosburgh, F., & Hebrank, J.<br />

1978. Shark skin: A function in locomotion.<br />

Science. 202:747-9.<br />

July/August 2011 Volume 45 Number 4 215


PAPER<br />

Can Biomimicry <strong>and</strong> Bioinspiration Provide<br />

Solutions for Fouling Control<br />

AUTHORS<br />

Emily Ralston<br />

Geoffrey Swain<br />

Florida Institute of <strong>Technology</strong><br />

Introduction<br />

Antifouling is changing. The ban<br />

on tributyltin (TBT), arguably the<br />

most effective <strong>and</strong> yet environmentally<br />

damaging antifouling precipitated an<br />

increase in research to develop new<br />

technology (Swain, 1999; Omae,<br />

2003). The immediate response was<br />

to return to copper as the antifouling<br />

of choice. However, high levels of copper<br />

in many ports <strong>and</strong> harbors has led<br />

to concern <strong>and</strong> even the banning of<br />

copper based antifouling (i.e., San<br />

Diego, Washington State; Carson<br />

et al., 2009; Nehring, 2001; Qian<br />

et al., 2010; Thomas et al., 2001).<br />

The use of toxic coatings may also be<br />

linked to the transport of biocide tolerant<br />

non-indigenous species (Dafforn<br />

et al., 2008). Some users have switched<br />

to biocide-free silicone antifouling<br />

coatings for improved hydrodynamic<br />

<strong>and</strong> environmental performance. However,<br />

these coatings may foul, which<br />

necessitates cleaning, <strong>and</strong> there may<br />

be an increased risk of transporting<br />

non-native species. They are mechanically<br />

weak <strong>and</strong> can be damaged more<br />

easily than traditional coatings<br />

(Chambers et al., 2006; Nehring, 2001;<br />

Swain, 2010; Swain et al., 2007; Yebra<br />

et al., 2004). An ideal coating will be<br />

hydraulically smooth <strong>and</strong> control fouling<br />

for the lifetime of the vessel, be<br />

ABSTRACT<br />

Biomimicry, modeling biological systems to find engineering methods, <strong>and</strong> bioinspiration,<br />

improving upon or repurposing the biological model, may provide direction<br />

for the development of new antifouling solutions. Despite being subject to constant<br />

pressure from foulers, many organisms maintain a clean surface. The challenge lies<br />

in selecting the most effective <strong>and</strong> reproducible antifouling mechanisms from<br />

nature <strong>and</strong> mimicking or modifying them to provide a realistic engineered solution.<br />

Keywords: natural antifouling, marine coatings, biomimicry, bioinspiration, antifouling,<br />

foul release<br />

environmentally compliant, control<br />

invasive species, be easily applied, repaired<br />

<strong>and</strong> maintained, be compatible<br />

with materials <strong>and</strong> methods of hull<br />

construction <strong>and</strong> decommissioning,<br />

<strong>and</strong> be cost effective (Swain, 2010).<br />

In the decades since a ban on TBT<br />

was proposed, research has been directed<br />

towards new solutions to the<br />

fouling problem. Some of the discoveries<br />

have been in the form of new<br />

booster biocides to improve the performance<br />

of copper based antifouling<br />

against biofilms (slimes). These include<br />

Irgarol 1051, Seanine 211, diuron,<br />

pyrithiones, <strong>and</strong> many others.<br />

Unfortunately, some of these boosters<br />

may parallel TBT in terms of environmental<br />

impact. For example, Irgarol<br />

1051 is a photosynthesis inhibitor<br />

that reduces plants’ ability to create<br />

energy causing growth to slow, reproduction<br />

to stop <strong>and</strong> eventually death of<br />

the plant. It has a long half-life <strong>and</strong><br />

does not partition into sediment so it<br />

is found primarily in the water column.<br />

It has been found to occur in estuaries,<br />

ports <strong>and</strong> harbors at elevated<br />

levels <strong>and</strong> is highly toxic to non-target<br />

marine plants (Hall et al., 1994;<br />

Thomas & Brooks, 2010; Thomas<br />

et al., 2001; Yebra et al., 2004). Recently,<br />

it has been detected in water<br />

samples collected from the Caribbean,<br />

Bermuda, Florida <strong>and</strong> Australia in the<br />

vicinity of seagrass beds <strong>and</strong> coral reefs<br />

where the effects could be devastating<br />

(Carbery et al., 2006; Owen et al.,<br />

2002; Scarlett et al., 1999). Researchers<br />

must be careful to avoid any chemistry<br />

that may impact the environment<br />

<strong>and</strong> non-target organisms. One source<br />

of new technology is to underst<strong>and</strong><br />

how organisms prevent fouling <strong>and</strong><br />

then mimic or draw inspiration from<br />

these biological models to create an<br />

engineering solution.<br />

Organisms that are unfouled or<br />

only lightly fouled provide insights<br />

into the mechanisms that have evolved<br />

to prevent surface colonization or epibiosis<br />

(Wahl, 1989). These “clean” organisms<br />

are the focus of research into<br />

biomimetic <strong>and</strong> bioinspired solutions<br />

to fouling on human structures. Biomimicry<br />

refers to the study of the<br />

structure <strong>and</strong> function of biological<br />

systems as models for the design of<br />

engineering solutions (dictionary.com)<br />

while bioinspiration exp<strong>and</strong>s on biomimicry<br />

by not only copying or imitating<br />

nature but also improving them or<br />

216 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


epurposing the biological model for<br />

an idealized engineering solution.<br />

This review will discuss the methods<br />

by which nature controls fouling<br />

<strong>and</strong> characterize natural antifouling<br />

<strong>and</strong> engineered solutions in terms of<br />

chemical, physical, mechanical, behavioral<br />

<strong>and</strong> combined mechanisms.<br />

Natural Antifouling<br />

Organisms have evolved several<br />

different strategies to prevent fouling.<br />

These natural antifouling methods include<br />

chemical, physical, mechanical,<br />

behavioral <strong>and</strong> a combination of<br />

more than one of the others (Bers &<br />

Wahl, 2004; Pawlik, 1992; Wahl,<br />

1989).<br />

Chemical<br />

Chemical antifouling has a long<br />

history of research <strong>and</strong> has been the<br />

subject of many reviews (Armstrong<br />

et al., 2000; Clare, 1996; Fusetani,<br />

2004; Omae, 2006; Pawlik, 1992;<br />

Qian et al., 2010; Raveendran &<br />

Mol, 2009; Rittschoff, 2000; <strong>and</strong><br />

others). To date, thous<strong>and</strong>s of active<br />

natural products have been identified<br />

(Pawlik, 1992). The activity of natural<br />

chemistries includes low pH, deterrents,<br />

anesthetics, attachment <strong>and</strong><br />

metamorphosis inhibitors, or toxic<br />

chemicals. The chemicals may be surface<br />

bound or water soluble (Omae,<br />

2006; Rittschof, 2000; Wahl, 1989).<br />

Despite issues with the ecological relevance<br />

of some of the chemicals, active<br />

natural products have been isolated<br />

from an algae, sponges, soft corals<br />

<strong>and</strong> a limpet that are either available<br />

at the surface or released into the<br />

water column when the organism<br />

is disturbed (de Nys & Steinberg,<br />

2002; Fusetani, 2004; Hay, 1996;<br />

Hellio et al., 2002). The mucus of dolphins,<br />

echinoderms, fish <strong>and</strong> corals<br />

contain antifouling chemicals that<br />

have been found to dissolve glue, prevent<br />

attachment or act as antimicrobial<br />

toxins (Baum et al., 2002; Bavington<br />

et al., 2004; Ebran et al., 2000; Ritchie,<br />

2006; Shephard, 1994; Videler et al.,<br />

1999). Many tunicates have acidic<br />

body pH <strong>and</strong> low epibiosis, especially<br />

in areas where density of acidic<br />

vacuoles is high (Hirose et al., 2001;<br />

Stoeker, 1980). Additionally, when<br />

looking at whole animal extracts,<br />

some tunicates contain chemicals that<br />

are cytotoxic, antimicrobial <strong>and</strong> antiviral<br />

(Davis & Wright, 1989). The<br />

eggs of many organisms, including<br />

fish <strong>and</strong> coral, are well protected with<br />

antimicrobial chemistries (Marquis<br />

et al., 2005, Ramasamy & Marugan,<br />

2007). Terrestrial plants have also<br />

yielded interesting chemistries such<br />

as tannins, pyrethroids <strong>and</strong> capsaicin<br />

(Feng et al., 2009; Perez et al., 2007;<br />

Xu et al., 2005).<br />

More recently, attention has turned<br />

to microorganisms. Antifouling metabolites<br />

that had been attributed in the<br />

past to algae, sponges, corals, etc., have<br />

been found on closer study, to be produced<br />

by surface associated bacteria<br />

<strong>and</strong> cyanobacteria (Armstrong et al.,<br />

2000; Clare, 1996; Krug, 2006).<br />

These surface associated microorganisms<br />

are distinct from the communities<br />

in the water column, often found in<br />

higher densities than the water column<br />

community <strong>and</strong> are often highly<br />

pigmented (Dobretsov et al., 2005;<br />

Faimalietal.,2004;Holmstrom<br />

et al., 1992). Deterrent biofilms preventfoulingbytoxicordeterrent<br />

chemistries (Dobretsov et al., 2006;<br />

Pawlik, 1992).<br />

Physical<br />

The two primary physical means<br />

identified to prevent fouling are surface<br />

energy <strong>and</strong> surface texture. A surface<br />

energy range of 20-30 dynes/cm<br />

(Baier, 1972; Dexter, 1979) has been<br />

shown to minimize adhesion <strong>and</strong><br />

favor the removal of epibionts. Such<br />

surface energy values have been measured<br />

on the surface of killer whales<br />

(Baier & Meyer, 1986), gorgonians<br />

(Vrolijk et al., 1990) <strong>and</strong> healthy<br />

teeth (Baier & Meyer, 1986; Glantz<br />

et al., 1991). Surface energy also effects<br />

settlement of some organisms, albeit<br />

in a species specific manner (Anderson<br />

et al., 2003; Meyer et al., 1988; Molino<br />

& Weatherbee, 2008; Rittschof &<br />

Costlow, 1989). For example, it has<br />

been found that barnacles <strong>and</strong> bryozoans<br />

prefer to settle on different surface<br />

energies (Dahlstrom et al., 2004;<br />

Rittschof & Costlow, 1989). Diatoms<br />

<strong>and</strong> the green algae Ulva have different<br />

adhesion strengths on surfaces with<br />

different wettability. Diatoms are more<br />

easily removed from hydrophilic surfaces,<br />

whereas Ulva releases more easily<br />

from hydrophobic surfaces (Finlay<br />

et al., 2002; Kirshnan et al., 2006).<br />

Low adhesion surfaces in nature are associated<br />

with waxes, oils, surfactants,<br />

mucuses or fluorinated or methylated<br />

compounds (Baum et al., 2003; Krug,<br />

2006; Shephard, 1994; Wahl, 1989).<br />

The antifouling properties of surface<br />

topography have received extensive<br />

attention <strong>and</strong> review (Scardino<br />

& de Nys, 2011; Scardino et al., 2008).<br />

The effectiveness of topography as<br />

antifouling appears to be the relationship<br />

of scale between the texture <strong>and</strong><br />

the settling organism. An ultra-smooth<br />

surface offers no refuge from predation<br />

or hydrodynamic stresses <strong>and</strong> is therefore<br />

unattractive (Kohler et al., 1999;<br />

Walters & Wethey, 1996). Surfaces<br />

with textures that are smaller than<br />

the settler reduce settlement <strong>and</strong>/or<br />

attachment strength, the “attachment<br />

point theory” (Scardino et al., 2008).<br />

Textures that are the same size or<br />

July/August 2011 Volume 45 Number 4 217


slightly larger than the propagules offer<br />

the greatest number of attachment<br />

points, the strongest attachment <strong>and</strong><br />

the best protection to settling organisms<br />

(Callow et al., 2002; Scardino<br />

et al., 2006). Organisms that have<br />

only one attachment point (barnacles,<br />

arborescent bryozoans, etc.) are more<br />

specific in searching for high quality<br />

pits than colonial organisms with multiple<br />

attachment points, likely because<br />

the colonial organisms outgrow the<br />

“refuge” of the pit quickly <strong>and</strong> can<br />

survive partial mortality (Walters &<br />

Wethey, 1996). The use of hairs in<br />

mussel spat (Dixon et al., 1995), spicules<br />

in gorgonian coral (Scardino &<br />

de Nys, 2011; Vrolijk et al., 1990)<br />

<strong>and</strong> spines in some colonial organisms<br />

(Dyrynda, 1986; Wahl, 1989) has<br />

been shown to prevent fouling <strong>and</strong><br />

overgrowth of fouling organisms. It<br />

has been suggested that many other organisms<br />

including crabs, brittle stars,<br />

molluscs, marine mammals <strong>and</strong> sharks<br />

(adult skin <strong>and</strong> dogfish egg cases), use<br />

textured surfaces with one or more<br />

scales of complexity to prevent micro<strong>and</strong><br />

macro-fouling (Baum et al.,<br />

2002; Bers & Wahl, 2004; Scardino<br />

& de Nys, 2004; Scardino & de Nys,<br />

2011).<br />

Mechanical<br />

Grooming is a common antifouling<br />

mechanism found in nature.<br />

Grooming involves specialized structuresthateitherpickorsweepan<br />

animals surface clean (Wahl, 1989).<br />

Decapod crustaceans have highly<br />

evolved brushes used to remove epibionts<br />

<strong>and</strong> parasites from specific<br />

parts of their bodies like the gills <strong>and</strong><br />

carapace (Acosta & Poirrier, 1992;<br />

Batang & Suzuki, 2003; Bauer, 1981).<br />

Historically, echinoderms <strong>and</strong> bryozoans<br />

were thought to use specialized<br />

structures to clean their surfaces<br />

(Campbell & Rainbow, 1977; Dyrynda,<br />

1986), but recently this has been called<br />

into question as the pedicellaria of the<br />

crown of thorns starfish (Acanthaster<br />

planci) were found to be too unresponsive<br />

<strong>and</strong> widely placed to be effective in<br />

keeping their surfaces clean (Guenther<br />

et al., 2007). Many organisms use<br />

ciliary cleaning in conjunction with<br />

mucus to keep surfaces clean (Wahl<br />

et al., 1998). Symbiotic or mutualistic<br />

relationships such as fish visiting<br />

“cleaning stations” (Poulin & Grutter,<br />

1996), mutualistic grazing of snails<br />

within populations (Wahl & Sonnichsen,<br />

1992; Wahl et al., 1998) <strong>and</strong> branchiobdellid<br />

annelids that feed on epibionts<br />

in the gill chamber of crayfish<br />

(Brown et al., 2002) are examples of<br />

beneficial relationships that may prevent<br />

fouling.<br />

The other mechanical method of<br />

antifouling is surface renewal via shedding<br />

or molting of outer layers. Crustaceans,<br />

stone fish <strong>and</strong> algae all molt,<br />

either the entire surface simultaneously<br />

or in patches, which removes<br />

all attached fouling (Bakus et al.,<br />

1986; Keats et al., 1997; Wahl, 1989).<br />

Additionally, many organisms use<br />

mucus as membrane to separate themselves<br />

from their environment. The<br />

mucus sloughs off removing foulers,<br />

makes adhesion difficult <strong>and</strong> fouls<br />

sensory <strong>and</strong> attachment apparatus of<br />

epibionts (Brown & Bythell, 2005;<br />

Davies & Hawkins, 1998; Denny,<br />

1989; Dyrynda, 1986; Shephard,<br />

1994; Wahl, 1989; Wahl et al., 1998).<br />

Behavioral<br />

Behavioral antifouling is the direct<br />

or indirect active avoidance of fouling<br />

organisms (Becker & Wahl, 1996).<br />

Burrowing into sediment, moving<br />

into the air, between fresh <strong>and</strong> salt<br />

water or into areas with very different<br />

oxygen contents <strong>and</strong> nocturnal activity<br />

or hiding in crevices are all mechanisms<br />

that remove less tolerant epibiotic<br />

organisms (Becker & Wahl, 1996;<br />

Wahl, 1989; Wahl et al., 1998). Organisms<br />

with a similar range of tolerances<br />

to their hosts will be unaffected<br />

by these behavioral methods (Brock<br />

et al., 1999).<br />

Combination<br />

Most organisms that are well studied<br />

use a combination of methods to<br />

prevent surface fouling. This was highlighted<br />

in reviews by Ralston <strong>and</strong><br />

Swain (2009) <strong>and</strong> Scardino <strong>and</strong><br />

de Nys (2011). Crustaceans groom<br />

<strong>and</strong> shed their shells <strong>and</strong> use behavioral<br />

mechanisms like burrowing <strong>and</strong> moving<br />

among habitats to ensure clean surfaces<br />

(Becker & Wahl, 1996; Wahl<br />

et al., 1998). Echinoderms groom,<br />

slough, excrete anti-adhesive mucus,<br />

have chemical antifoulants <strong>and</strong> may<br />

even use a strong negatively charged<br />

cuticle to prevent surface colonization<br />

(Bakus et al., 1986; Bavington et al.,<br />

2004; Bryan et al., 1996; McKenzie<br />

& Grigolava, 1996). Corals use antibacterial<br />

mucus, select specific microbial<br />

colonists which in turn protect<br />

them from other microbes, slough<br />

mucus <strong>and</strong> surface layers <strong>and</strong> secrete<br />

secondary metabolites to keep their<br />

surfaces clean (Brown & Bythell,<br />

2005; Ritchie, 2006; Targett et al.,<br />

1983). Dolphins <strong>and</strong> whales are hypothesized<br />

to use microtopography<br />

in conjunction with an enzymatically<br />

active zymogel to prevent attachment<br />

of macrofoulers (Baum et al., 2003;<br />

Meyer & Seegers, 2004) <strong>and</strong> surface<br />

sloughing, skin compliance <strong>and</strong> a critical<br />

surface tension in the preferred<br />

range for minimal adhesion combined<br />

with breaching may remove fouling at<br />

an early stage (Baum et al., 2003; Fish<br />

& Rohr, 1999; Scardino & de Nys,<br />

2011). Algae have provided many<br />

218 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


new chemical metabolites that prevent<br />

fouling, shed their outer layers <strong>and</strong><br />

may remove settled epibionts by flexing<br />

beyond what their epibionts can<br />

withst<strong>and</strong> (Nylund & Pavia, 2005;<br />

Scardino & de Nys, 2011; Walters<br />

et al., 2003; Wikstrom & Pavia, 2004).<br />

Biomimetic <strong>and</strong> Bioinspired<br />

Engineering Solutions<br />

Chemical<br />

Due to our vast experience with incorporating<br />

chemicals into coatings, it<br />

is not surprising that natural products<br />

are the most investigated biological<br />

antifouling. SeaNine 211 is a booster<br />

biocide added to copper coatings to<br />

boost efficacy against fouling plants.<br />

It degrades quickly in water <strong>and</strong> sediment,<br />

binds strongly to sediment, has<br />

low environmental toxicity <strong>and</strong> has an<br />

excellent performance record from lab<br />

<strong>and</strong> field tests <strong>and</strong> ship trials (Thomas<br />

& Brooks, 2010; Yebra et al., 2004).<br />

SeaNine 211 is based on the natural<br />

product isothiozolone originally isolated<br />

in the 1980s from the soft coral<br />

Eunicea (Raveendran & Mol, 2009).<br />

Econea is a halogenated pyrrol that is<br />

the active ingredient in copper-free<br />

antifouling paints from manufacturers<br />

such as Petit, Interlux, Sea Hawk <strong>and</strong><br />

others. Halogenated pyrrols are common<br />

secondary metabolites in bacteria<br />

<strong>and</strong> sponges (or possibly surface bacteria<br />

associated with sponges) <strong>and</strong> are<br />

potent settlement <strong>and</strong> metamorphosis<br />

inhibitors for barnacles <strong>and</strong> other<br />

animal foulers (Dahms et al., 2006;<br />

Omae, 2006). Because of its specificity<br />

against animal fouling, Econea is often<br />

combined with a booster like SeaNine<br />

to prevent plant fouling as well.<br />

Perhaps the most studied natural<br />

chemistry is the halogenated furanone<br />

originally isolated from the red algae<br />

Delisea pulchra. The chemical is present<br />

on the surface of the plant in concentrations<br />

that prevent fouling <strong>and</strong><br />

the coverage of epibionts corresponds<br />

to concentrations of the furanone<br />

(de Nys & Steinberg, 2002). It has<br />

not yet been successfully incorporated<br />

into a long-lasting ship hull coating<br />

(Chambers et al., 2006); however,<br />

some have reported that it is available<br />

products called “Netsafe” <strong>and</strong> “Pearlsafe”<br />

marketed in Australia for use in<br />

commercial aquaculture (Raveendran<br />

& Mol, 2009). We were unable to<br />

find any record of these products for<br />

sale at this time so they may no longer<br />

be available. Many other natural<br />

chemicals from marine macroorganisms<br />

show promise for non-toxic or<br />

low-toxicity antifouling paints <strong>and</strong><br />

are being investigated (see reviews by<br />

Armstrong et al., 2000; Fusetani,<br />

2004; Omae, 2006; Qian et al.,<br />

2010; Raveendran & Mol, 2009).<br />

Terrestrial plants have also yielded<br />

promising chemistries for antifouling.<br />

These include products like tannins,<br />

pyrethroids <strong>and</strong> capsaicin (Feng et al.,<br />

2009; Perez et al., 2007; Thomas &<br />

Brooks, 2010; Xu et al., 2005). Pyrethroids,<br />

synthetic analogs of pyrethrin<br />

from chrysanthemum flowers, are<br />

of particular interest because they are<br />

already approved for use as environmentally<br />

safe insecticides. These insecticides<br />

have low toxicity to mammals,<br />

do not persist, do not bioaccumulate<br />

<strong>and</strong> are available in industrial quantities<br />

(Feng et al., 2009). Tannin is present<br />

in terrestrial plants, mangroves <strong>and</strong><br />

in some marine algae, primarily as an<br />

anti-herbivory chemical. However,<br />

some have found it to have antifouling<br />

properties as well (Brock et al., 2007;<br />

Lau & Qian, 1997; Perez et al., 2007;<br />

Wikstrom & Pavia, 2004). The longterm<br />

efficacy of tannin isolated from<br />

the quebracho tree was improved by<br />

precipitating it with aluminum forming<br />

a salt which increased the life<br />

span of the coating to 1 month in the<br />

field (Perez et al., 2007).<br />

Another avenue of research is isolating<br />

chemicals from microorganisms<br />

<strong>and</strong> using the microorganisms themselves.<br />

There are many benefits to<br />

this strategy including culturability,<br />

abundance <strong>and</strong> ability to trick or stress<br />

the organisms into producing large<br />

quantities of the necessary chemical<br />

(Dobretsov et al., 2006; Holmstrom<br />

& Kjelleberg, 1994). Holmstrom et al.<br />

(2000) were able to keep bacteria alive<br />

in a coating for 14 days in the laboratory.<br />

Microencapsulation is another<br />

strategy being investigated, not just<br />

for microorganisms but for all natural<br />

products, as a way to increase length<br />

of efficacy. Coatings with microencapsulated<br />

living bacteria were able<br />

to prevent fouling up to 7 weeks in<br />

field trials (Chambers et al., 2006; Yee<br />

et al., 2007).<br />

The use of enzymes <strong>and</strong> hormones<br />

that are commercially available is another<br />

strategy for chemical antifouling.<br />

Many patents have been awarded <strong>and</strong><br />

there is an enzymatic coating available<br />

on the Danish yacht market, although<br />

little scientific evidence of effectiveness<br />

exists (Olsen et al., 2007). Enzymes<br />

may act directly by dissolving<br />

glues, lysing cells or decomposing<br />

exoskeletons of barnacles (Abarzua &<br />

Jakubowski, 1995; Evans & Clarkson,<br />

1993; Olsen et al., 2007). They may<br />

also act indirectly by increasing the<br />

effectiveness of an antifoulant or by<br />

acting on the coating to improve release<br />

or polishing rates (Olsen et al.,<br />

2007). Hormones, such as noradrenaline,<br />

may also be used as a non-toxic<br />

deterrent in antifouling coatings<br />

(Gohad et al., 2010). However, both<br />

enzymes <strong>and</strong> hormones have several<br />

drawbacks to their widespread use<br />

July/August 2011 Volume 45 Number 4 219


including expense, instability, potentially<br />

specific response <strong>and</strong> need for<br />

environmental approval (Gohad et al.,<br />

2010; Olsen et al., 2007; Rittschof,<br />

2000).<br />

There are several challenges in getting<br />

new chemicals approved for use in<br />

antifouling paints. The environmental<br />

problems associated with TBT have<br />

increased awareness of the potential<br />

risks associated with introducing new<br />

chemistries <strong>and</strong> attention must be<br />

given to proving environmental compliance.<br />

This greatly increases the<br />

time <strong>and</strong> the cost to go from the identification<br />

<strong>and</strong> isolation of a chemical<br />

to commercialization, which can cost<br />

millions of dollars <strong>and</strong> take over<br />

10 years to get approval. Furthermore,<br />

many natural products are structurally<br />

complex <strong>and</strong> only available in small<br />

amounts in the organism so there are<br />

issues with obtaining or synthesizing<br />

the active chemicals (Fusetani, 2004;<br />

Rittschof, 2000; Rittschof, 2001).<br />

Natural products tend to have a short<br />

life span as they cannot be toxic to the<br />

organism.Thisleadstoissueswhen<br />

incorporated into a coating as coatings<br />

need to maintain efficacy for<br />

3-12 years of service life (de Nys &<br />

Steinberg, 2002; Fusetani, 2004;<br />

Ingle, 2007; Marechal & Hellio,<br />

2009; Rittschof, 2000; Rittschof,<br />

2001). Conversely, the short life span<br />

is beneficial for environmental compliance<br />

as one of the characteristics of an<br />

ideal chemical antifoulant is short half<br />

life (Clare, 1996). Other factors for an<br />

ideal chemical antifoulant include<br />

non-toxicity, activity against a wide<br />

variety of fouling organisms, easy incorporation<br />

into a controlled release<br />

coating <strong>and</strong> should come from a culturable<br />

organism or have an active<br />

chemistry that can be industrially synthesized<br />

(Clare, 1996; Hellio et al.,<br />

2002; Marechal & Hellio, 2009;<br />

Raveendran & Mol, 2009; Rittschof,<br />

2000; Yebra et al., 2004).<br />

Physical<br />

Slippery coatings are not new<br />

technology. Commercially available<br />

fouling release coatings have been<br />

available since the mid-1970s <strong>and</strong> are<br />

proving to be an increasingly successful<br />

method for fouling control. These<br />

coatings combine polydimethylsiloxane<br />

silicone with low surface energy,<br />

oils <strong>and</strong> compliance to reduce the<br />

adhesion strength of organisms to a<br />

surface. Fouling is removed by hydrodynamic<br />

shear forces or with gentle<br />

cleaning (Anderson et al., 2003).<br />

There are several lanolin based waxes<br />

on the market for use on ship hulls<br />

<strong>and</strong> propellers. While the wax may<br />

provide short-term antifouling, the<br />

main purpose is to lessen attachment<br />

strength <strong>and</strong> make cleaning easier.<br />

These products are often used over a<br />

tough epoxy, however the duration of<br />

effect is unknown <strong>and</strong> we were unable<br />

to find any published data in a peer reviewed<br />

journal to back the claims of<br />

manufacturers. Results obtained from<br />

these coatings may vary depending<br />

on the fouling communities. Effects<br />

of surface energy <strong>and</strong> wettability on<br />

surface colonization are species specific<br />

with some responding favorably to hydrophobic<br />

surfaces <strong>and</strong> some to hydrophilic<br />

or intermediate surfaces (Callow<br />

et al., 2002, Dahlstrom et al., 2004).<br />

Additionally primary colonizers often<br />

change the surface energy of surfaces<br />

which will change the effect on subsequent<br />

settlers (Scardino & de Nys,<br />

2011).<br />

Mimicking the surface texture of<br />

marine organisms has been investigated<br />

as an environmentally friendly<br />

antifoulant. “Sealcoat” is a commercially<br />

available antifouling coating<br />

that is flocked with fibers mimicking<br />

the fur of a seal. According to the company’s<br />

website, fouling is prevented for<br />

up to 5 years. However, no scientific<br />

data exists to back this claim. Other<br />

studies looking at flocked or furred<br />

coatings found mixed responses with<br />

green <strong>and</strong> brown algae, encrusting<br />

bryozoans <strong>and</strong> barnacles deterred, red<br />

algae <strong>and</strong> hydroids unaffected <strong>and</strong><br />

solitary tunicates <strong>and</strong> tube worms<br />

increased by these coatings (Phillippi<br />

et al., 2001). It must also be remembered<br />

that seals do not only depend<br />

on their fur to keep them fouling free<br />

but also groom <strong>and</strong> spend large amounts<br />

of time out of the water.<br />

Mimics of topographies from other<br />

organisms such as crustose coralline<br />

algae, molluscs, crabs, brittle stars,<br />

soft corals <strong>and</strong> dogfish egg cases, have<br />

been investigated for antifouling activity<br />

<strong>and</strong> have shown short-term efficacy<br />

in laboratory assays. Additionally,<br />

topographies from pilot whale <strong>and</strong><br />

shark skins have been characterized<br />

<strong>and</strong> had an antifouling activity attributed<br />

to the microstructures. These active<br />

topographies range in scale from 1<br />

to 300 μm with multiple length scales<br />

occurring on natural surfaces (Baum<br />

et al., 2002; Bers & Wahl, 2004;<br />

Scardino & de Nys, 2004; Scardino<br />

& de Nys, 2011). The “Sharklet” is<br />

an example of a biomimetic texture<br />

used as an engineered surface to prevent<br />

fouling. It has performed well in<br />

laboratory assays against Ulva spores<br />

<strong>and</strong> Balanus amphitrite cyprids (Carman<br />

et al., 2006; Schumacher et al., 2007).<br />

In order to improve the effect of this<br />

<strong>and</strong> other topographies, a mathematical<br />

model was created called<br />

the “Engineered Roughness Index”;<br />

this index can also be used to predict<br />

settlement of marine organisms on<br />

the engineered topographies (Long<br />

et al., 2010). Engineered surfaces<br />

with hierarchically wrinkled surfaces<br />

220 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


have shown promising results in field<br />

trials, especially against barnacles<br />

(Efimenko et al., 2009; Scardino & de<br />

Nys, 2011).<br />

Sound has been suggested as an<br />

antifouling method. However, there<br />

are no published field test data that scientifically<br />

prove that it can provide<br />

long-term antifouling. This is not a<br />

truly biomimetic method as it is not reported<br />

as a natural antifouling mechanism.<br />

Sound is used by competent<br />

larvae of fish <strong>and</strong> invertebrates like<br />

crabs to navigate to appropriate settlement<br />

sites (Radford et al., 2010;<br />

Simpson et al., 2008; Stanley et al.,<br />

2010). Specific habitats have different<br />

auditory signatures <strong>and</strong> larvae can use<br />

these to differentiate <strong>and</strong> pilot to their<br />

adult habitats (Radford et al., 2010).<br />

Both high- <strong>and</strong> low-frequency sound<br />

waves have been shown to be effective<br />

at inhibiting settlement of barnacles<br />

<strong>and</strong> mussels (Branscomb & Rittschof,<br />

1984; Donskoy & Ludyanskiy, 1995;<br />

Guo et al., 2011). Additionally, ultrasound<br />

waves have been used to destroy<br />

barnacle larvae via cavitation for ballast<br />

water treatment (Seth et al., 2010).<br />

The use of low-frequency sound is limited<br />

because it is audible to humans<br />

<strong>and</strong> other organisms <strong>and</strong> therefore<br />

noise pollution is an issue. Several<br />

companies worldwide (Ultrasonic Antifouling,<br />

ASM, Sonihull <strong>and</strong> others)<br />

offer ultrasonic units that can be installed<br />

that are purported to prevent<br />

fouling or conversely to kill settling<br />

fouling organisms thereby making<br />

them easy to remove. However, this<br />

method is variable in effect, with settlement<br />

rates ranging from 1% up to<br />

55% for low <strong>and</strong> high frequency, respectively<br />

(Branscomb & Rittschof,<br />

1984; Guo et al., 2011). Guo <strong>and</strong> colleagues<br />

(2011) reported a settlement<br />

rate for barnacle cyprids in the laboratory<br />

of about 20% for their best ultrasonic<br />

treatment compared to a rate of<br />

around 70% for the control so the<br />

method is not perfect. Additionally,<br />

Sonihull reports changes in fish behavior<br />

when their ultrasonic units are<br />

in use so there are noise pollution concerns<br />

with high-frequency sound as<br />

well.<br />

Physical methods of antifouling are<br />

often inferred but seldom proved due<br />

to challenges associated with testing<br />

living materials. Surface topography<br />

effects are scale dependent (Scardino<br />

et al., 2006; Schumacher et al., 2007)<br />

<strong>and</strong> effectiveness in fouling prevention<br />

may vary geographically (Bers et al.,<br />

2010). Finding a universal physical<br />

antifoulant may be difficult <strong>and</strong> the<br />

results are often short lived, lasting a<br />

monthorlessinfield testing (Holm<br />

et al., 1997).<br />

Mechanical<br />

Mechanical cleaning is performed<br />

on ships, aquaculture nets, instruments<br />

<strong>and</strong> other marine structures<br />

when they become fouled, either because<br />

an antifouling coating was not<br />

used or if that coating becomes fouled.<br />

The U.S. Navy cleans their vessels<br />

when a set level of fouling is reached<br />

as set out in the Naval Ships’ Technical<br />

Manual (Cologer, 1984; NSTM,<br />

2006). Cleaning is reactive <strong>and</strong> has<br />

been shown to speed the rate of recolonization<br />

<strong>and</strong> increase the risk of<br />

transport of nonindigenous species<br />

(Floerl et al., 2005). Additionally,<br />

commercially available brush cleaning<br />

devices (i.e., SCAMP, Mini-Pamper,<br />

etc.) are harsh <strong>and</strong> may damage the<br />

antifouling coating. A new direction<br />

for mechanical antifouling is to mimic<br />

natural grooming. This is the idea behind<br />

the HullBUG (Hull Bioinspired<br />

Underwater Grooming), an autonomous<br />

robot that will proactively pass<br />

over a hull while a ship is in port. Its<br />

mode of action is a gentle wiping or<br />

brushing of the surface on a frequent<br />

schedule sufficient to remove fouling<br />

at its earliest stages before it can become<br />

established (Borchardt, 2010;<br />

Tribou & Swain, 2010). Results<br />

from field testing of fouling release<br />

<strong>and</strong> copper-coated panels subjected<br />

to grooming are so promising that further<br />

experiments <strong>and</strong> scale up on this<br />

method are being investigated.<br />

Ecospeed is a commercially available<br />

hull coating system. It is a tough<br />

glass flake reinforced vinyl ester<br />

coating. When combined with hull<br />

cleaning, this non-toxic coating is<br />

purported to maintain a fouling free<br />

surface with no repainting for up to<br />

25 years. Additionally, the coating<br />

smooths during cleaning, decreasing<br />

drag. Again, no scientifically published<br />

data exists to back the claims made<br />

by the manufacturer.<br />

Surface renewal has been attempted<br />

as an antifoulant for ship<br />

hull coatings. Polymers were developed<br />

that hydrolyze in seawater leaving<br />

a clean surface as they dissolve<br />

(C<strong>and</strong>ries et al., 2000). To date, however,<br />

these coatings have only been<br />

successful when combined with biocides<br />

as the rate of dissolution <strong>and</strong><br />

thickness of the coating required<br />

would be too great without the help<br />

of toxic chemicals.<br />

Mechanical antifouling has proven<br />

to be an effective but imperfect method<br />

of keeping submerged surfaces clean.<br />

Cleaning requires the deployment of<br />

equipment <strong>and</strong> usually divers, which<br />

increases both the expense <strong>and</strong> human<br />

risk factor of this antifouling<br />

mechanism. When applied to toxic<br />

coatings, cleaning may increase the release<br />

of biocides, at least in the short<br />

term (Schiff et al., 2004). Cleaning<br />

may cause damage to coatings which<br />

increases the rate of re-colonization<br />

July/August 2011 Volume 45 Number 4 221


<strong>and</strong> may increase the risk of transport<br />

of invasive species (Floerl et al., 2005;<br />

Piola & Johnston, 2008). Mechanical<br />

antifouling works better when combined<br />

with another antifouling method<br />

such as a biocidal coating or a fouling<br />

release surface. Grooming, however,<br />

is proactive <strong>and</strong> more closely matches<br />

many of the behavioral activities<br />

found in nature. Many organisms<br />

benefit from self or mutual grooming<br />

to maintain their surfaces free of<br />

fouling.<br />

Behavioral<br />

Behavioral methods include removing<br />

a vessel from the water when<br />

not in use or moving between fresh<br />

<strong>and</strong> salt water. The former is commonly<br />

practiced by recreational boat<br />

owners; however, removing a large<br />

vessel from the water is impractical, especially<br />

if that ship is frequently used.<br />

Moving vessels between fresh <strong>and</strong><br />

salt water, by traversing through the<br />

Panama Canal, for instance, is performed<br />

occasionally <strong>and</strong> has been<br />

credited with preventing the unobstructed<br />

movement of Caribbean <strong>and</strong><br />

Pacific species between the two bodies<br />

of water. Brock <strong>and</strong> colleagues (1999)<br />

found that moving a ship into fresh<br />

waterfor9dayswassufficient to remove<br />

90% of fouling from the hull.<br />

However, tolerant fouling organisms<br />

will not be affected by this antifouling<br />

method as shown by the survival <strong>and</strong><br />

subsequent introduction of the mussel<br />

Mytilus galloprovincialis to Oahu,<br />

Hawaii, from Washington. The mussel<br />

was one of the 10% of fouling organisms<br />

remaining on the USS Missouri<br />

after its Pacific transit <strong>and</strong> was seen<br />

spawning shortly after arrival in Pearl<br />

Harbor <strong>and</strong> later found colonizing<br />

the ballast tanks of a submarine (Apte<br />

et al., 2000).<br />

Combined<br />

It is unlikely that any one antifouling<br />

mechanism will be sufficient to<br />

prevent all fouling in all situations<br />

that may be encountered by submerged<br />

structures. Indeed, every<br />

organism that is well studied with<br />

regards to natural antifouling uses a<br />

combination of strategies to maintain<br />

a clean surface. The most effective<br />

coatings in use today also use more<br />

than one antifouling mechanism; antifouling<br />

coatings use a biocide combined<br />

with self-polishing or ablative<br />

mechanism to keep an active layer at<br />

the surface. Fouling release coatings<br />

combine low surface energy, oils <strong>and</strong><br />

compliance to maximize self-cleaning.<br />

To date, most researchers investigating<br />

a biomimetic solution to antifouling<br />

have focused on only one method.<br />

However, that is beginning to change<br />

with the recent publication of reviews<br />

focusing on combined antifouling<br />

mechanisms (Ralston & Swain, 2009;<br />

Scardino & de Nys, 2011).<br />

Bioinspired Approaches<br />

The examples highlighted above<br />

represent biomimetic solutions for<br />

biofouling control. Very little research<br />

exists that takes lessons from nature<br />

<strong>and</strong> adapts or alters them for a true<br />

bioinspired solution. It has only been<br />

recently that novel uses have been proposed<br />

from biological models. For example,<br />

the dopamine based adhesive<br />

system in mussels, a common fouling<br />

organism, has been investigated as a way<br />

to obtain better adhesion of non-stick<br />

coatings to a substrate. The dopamine<br />

adhesive allows testing on polyethylene<br />

glycol (PEG) <strong>and</strong> other slippery<br />

polymers where before it was not possible<br />

because the polymers would not<br />

sticktoanything.Thosecoatings<br />

using the bioinspired PEG-DOPA system<br />

outperformed traditional silicone<br />

fouling release coatings in laboratory<br />

assays, comparing both the settlement<br />

<strong>and</strong> adhesion of a common fouling<br />

diatom <strong>and</strong> alga (Statz et al., 2006).<br />

Conclusions<br />

<strong>Marine</strong> organisms can achieve<br />

long-term protection from fouling<br />

using short-lived renewable mechanisms.<br />

This is attributed to using a<br />

combination of chemical, physical,<br />

mechanical <strong>and</strong> behavioral mechanisms.<br />

Much research has been published<br />

investigating the specific ways<br />

that organisms maintain a clean surface<br />

but frequently focus on only one<br />

mechanism without considering the efficacy<br />

of a holistic combined method.<br />

The challenge, for us, is to identify <strong>and</strong><br />

select the best natural systems to solve<br />

the problem of biofouling. Through<br />

improved knowledge of natural systems,wewillbebetterabletoboth<br />

mimic <strong>and</strong> innovate using biological<br />

models to find engineering solutions.<br />

Results so far have been promising<br />

but better interactions between biologists,<br />

ecologists, engineers, chemists<br />

<strong>and</strong> materials scientists are needed<br />

<strong>and</strong> publishing of results is vitally important.<br />

Despite some issues, biomimetics<br />

<strong>and</strong> bioinspiration hold great<br />

promise for new antifouling solutions.<br />

For example, the HullBUG grooming<br />

method now being developed by the<br />

Office of Naval Research demonstrates<br />

how a proactive grooming method will<br />

enhance the long-term effectiveness<br />

of the presently available commercial<br />

antifouling or fouling release surfaces.<br />

Acknowledgments<br />

The authors would like to thank<br />

the Office of Naval Research (grants<br />

N000140210217, N000140810034<br />

<strong>and</strong> N000140910843), who has<br />

222 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


funded much of this work <strong>and</strong> continues<br />

to support research directed<br />

towards discovering improved antifouling<br />

technology <strong>and</strong> the participants<br />

in the ONR Coatings Research<br />

Group.<br />

Authors:<br />

Emily Ralston <strong>and</strong> Geoffrey Swain<br />

Florida Institute of <strong>Technology</strong><br />

150 W. University Blvd.,<br />

Melbourne, FL 32901<br />

Emails: eralston@fit.edu;<br />

swain@fit.edu<br />

References<br />

Abarzua, S., & Jakubowski, S. 1995. Biotechnological<br />

investigation for the prevention<br />

of biofouling. I. Biological <strong>and</strong> biochemical<br />

principles for the prevention of biofouling.<br />

Mar Ecol Prog Ser. 123:301-12. doi: 10.3354/<br />

meps123301.<br />

Acosta, C., & Poirrier, M. 1992. Grooming<br />

behavior <strong>and</strong> associated structures of the<br />

mysid Mysidopsis bahia. J Crustacean Biol.<br />

12:383-91. doi: 10.2307/1549032.<br />

Anderson, C., Atlar, M., Callow, M., C<strong>and</strong>ries,<br />

M., Milne, A., & Townsin, R. 2003. The<br />

development of foul-release coatings for<br />

seagoing vessels. J Mar Des Operations.<br />

B4:11-23.<br />

Apte, S., Holl<strong>and</strong>, B., Godwin, L., & Gardner,<br />

J. 2000. Jumping ship: A stepping stone<br />

event mediated transfer of a non-indigenous<br />

species via a potentially unsuitable environment.<br />

Biol Invasions. 2:75-79. doi: 10.1023/<br />

A:1010024818644.<br />

Armstrong, E., Boyd, K., & Burgess, J. 2000.<br />

Prevention of marine biofouling using natural<br />

compounds from marine organisms. Biotech<br />

Ann Rev. 6:221-41. doi: 10.1016/S1387-<br />

2656(00)06024-5.<br />

Baier, R. 1972. Influence of the initial surface<br />

condition of materials on bioadhesion.<br />

In: Proc. 3rd Int. Congr. Mar. Corr. Foul.<br />

Maryl<strong>and</strong>: Northwestern University Press.<br />

Baier, R., & Meyer, A. 1986. Biosurface<br />

chemistry for fun <strong>and</strong> profit. Chemtech.<br />

16:178-85.<br />

Bakus, G., Targett, N., & Schulte, B. 1986.<br />

Chemical ecology of marine organisms.<br />

J Chem Ecol. 12:951-87. doi: 10.1007/<br />

BF01638991.<br />

Batang, Z., & Suzuki, H. 2003. Gill-cleaning<br />

mechanisms of the burrowing thalassinidean<br />

shrimps Nihonotrypaea japonica <strong>and</strong> Upogebia<br />

major (Crustacea: Decapoda). J Zool. 261:<br />

69-77. doi: 10.1017/S0952836903003959.<br />

Bauer, R. 1981. Grooming behavior <strong>and</strong><br />

morphology in the decapod crustacea.<br />

J Crustacean Biol. 1:153-73. doi: 10.2307/<br />

1548154.<br />

Baum, C., Meyer, W., Stelzer, R., &<br />

Fleischer, L-G. 2002. Average nanorough<br />

skin surface of the pilot whale (Globicephala<br />

melas, Delphinidae): Considerations on the<br />

self-cleaning abilities based on nanoroughness.<br />

Mar Biol. 140:653-7. doi: 10.1007/s00227-<br />

001-0710-8.<br />

Baum, C., Simon, F., Meyer, W., Fleischer,<br />

L-G., Siebers, D., Kacza, J., & Seeger, J. 2003.<br />

Surface properties of the skin of the pilot whale<br />

Globicephala melas. Biofouling. 19:181-6.<br />

doi: 10.1080/0892701031000061769.<br />

Bavington, C., Lever, R., Mulloy, B.,<br />

Grundy, M., Page, C., Richardson, N., &<br />

McKenzie, J. 2004. Anti-adhesive glycoproteins<br />

in echinoderm mucus secretions.<br />

Comp Biochem Physiol B. 139:607-17.<br />

doi: 10.1016/j.cbpc.2004.07.008.<br />

Becker, K., & Wahl, M. 1996. Behaviour<br />

patterns as natural antifouling mechanisms<br />

of tropical marine crabs. J Exp Mar Biol<br />

Ecol. 203:245-58. doi: 10.1016/0022-<br />

0981(96)02575-0.<br />

Bers, A., Diaz, E., Da Gama, B., Vieira-Silva,<br />

F., Dobretsov, S., Valdivia, N., … Wahl, M.<br />

2010. Relevance of mytilid shell microtopographies<br />

for fouling defence—A global<br />

comparison. Biofouling. 26:367-77.<br />

doi: 10.1080/08927011003605888.<br />

Bers, A., & Wahl, M. 2004. The influence of<br />

natural surface microtopographies on fouling.<br />

Biofouling. 20:43-51. doi: 10.1080/<br />

08927010410001655533.<br />

Borchardt, J. 2010. Grooming the fleet.<br />

Mech Eng. 132:33-35.<br />

Branscomb, E., & Rittschof, D. 1984. An<br />

investigation of low frequency sound waves<br />

as a means of inhibiting barnacle settlement.<br />

J Exp Mar Biol Ecol. 79:149-54. doi: 10.1016/<br />

0022-0981(84)90215-6.<br />

Brock, E., Nylund, G., & Pavia, H. 2007.<br />

Chemical inhibition of barnacle larval settlement<br />

by the brown alga Fucus vesiculosus.<br />

Mar Ecol Prog Ser. 337:165-74. doi: 10.3354/<br />

meps337165.<br />

Brock, R., Bailey-Brock, J., & Goody, J.<br />

1999. A case study of the efficacy of freshwater<br />

immersion in controlling introduction of<br />

alien marine fouling communities: The USS<br />

Missouri. Pac Sci. 53:223-31.<br />

Brown, B., & Bythell, J. 2005. Perspectives on<br />

mucus secretion in reef corals. Mar Ecol Prog<br />

Ser. 296:291-309. doi: 10.3354/meps296291.<br />

Brown, B., Creed, R., & Dobson, W. 2002.<br />

Branchiobdellid annelids <strong>and</strong> their crayfish<br />

hosts: Are they engaged in a cleaning symbiosis<br />

Oecologia. 132:250-5. doi: 10.1007/s00442-<br />

002-0961-1.<br />

Bryan, P., Rittschof, D., & McClintock, J.<br />

1996. Bioactivity of echinoderm ethanolic<br />

body-wall extracts: An assessment of marine<br />

bacterial attachment <strong>and</strong> macroinvertebrate<br />

larval settlement. J Exp Mar Biol Ecol. 196:<br />

79-96. doi: 10.1016/0022-0981(95)00124-7.<br />

Callow, M., Jennings, A., Brennan, A.,<br />

Seegert, C., Gibson, A., Wilson, L., …<br />

Callow, J. 2002. Microtopographic cues for<br />

settlement of zoospores of the green fouling<br />

alga Enteromorpha. Biofouling. 18:237-45.<br />

doi: 10.1080/08927010290014908.<br />

Campbell, A., & Rainbow, P. 1977. The<br />

role of pedicellariae in preventing barnacle<br />

settlement on the sea-urchin test. Mar Behav<br />

Physiol. 4:253-60. doi: 10.1080/<br />

10236247709386957.<br />

C<strong>and</strong>ries, M., Atlar, M., & Anderson, C.<br />

2000. Considering the use of alternative<br />

July/August 2011 Volume 45 Number 4 223


antifoulings: The advantages of foul-release<br />

systems. In: Conference Proceedings, ENSUS<br />

2000. pp. 88-95. Newcastle, UK: University<br />

of Newcastle-upon-Tyne.<br />

Carbery, K., Owen, R., Frickers, T., Otero,<br />

E., & Readman, J. 2006. Contamination of<br />

Caribbean coastal waters by the antifouling<br />

herbicide Irgarol 1051. Mar Pollut Bull.<br />

52:635-44. doi: 10.1016/j.marpolbul.2005.<br />

10.013.<br />

Carman, M., Estes, T., Feinberg, A.,<br />

Schumacher, J., Wilkerson, W., Wilson, L., …<br />

Brennan, A. 2006. Engineered antifouling<br />

microtopographies—correlating wettability<br />

with cell attachment. Biofouling. 22:11-21.<br />

doi: 10.1080/08927010500484854.<br />

Carson, R., Damon, M., Johnson, L., &<br />

Gonzalez, J. 2009. Conceptual issues in<br />

designing a policy to phase out metal-based<br />

antifouling paints on recreational boats in<br />

San Diego Bay. J Environ Manag. 90:2460-8.<br />

doi: 10.1016/j.jenvman.2008.12.016.<br />

Chambers, L., Stokes, K., Walsh, F., &<br />

Wood, R. 2006. Modern approaches to<br />

marine antifouling coatings. Surf Coat Tech.<br />

201:3642-52. doi: 10.1016/j.surfcoat.2006.<br />

08.129.<br />

Clare, A. 1996. <strong>Marine</strong> natural product<br />

antifoulants: Status <strong>and</strong> potential. Biofouling.<br />

9:211-29. doi: 10.1080/08927019609378304.<br />

Cologer, C. 1984. Six year interaction of<br />

underwater cleaning with copper based antifouling<br />

paints on Navy surface ships. Nav Eng J.<br />

96:200-8. doi: 10.1111/j.1559-3584.1984.<br />

tb01829.x.<br />

Dafforn, K., Glasby, T., & Johnston, E.<br />

2008. Differential effects of tributyltin <strong>and</strong><br />

copper antifoulants on recruitment of nonindigenous<br />

species. Biofouling. 24:23-33.<br />

doi: 10.1080/08927010701730329.<br />

Dahlstrom, M., Jonsson, H., Jonsson, P.,<br />

& Elwing, H. 2004. Surface wettability as a<br />

determinant in the settlement of the barnacle<br />

Balanus improvisus (Darwin). J Exp Mar Biol<br />

Ecol. 305: 223-32. doi: 10.1016/j.jembe.<br />

2003.12.013.<br />

Dahms, H., Ying, X., & Pfeiffer, C. 2006.<br />

Antifouling potential of cyanobacteria:<br />

A mini-review. Biofouling. 22:317-27.<br />

doi: 10.1080/08927010600967261.<br />

Davis, A., & Wright, A. 1989. Interspecific<br />

differences in fouling of two congeneric ascidians<br />

(Eudistoma olivaceum <strong>and</strong> E. capsulatum): Is<br />

surface acidity an effective defense Mar Biol.<br />

102:491-7. doi: 10.1007/BF00438350.<br />

Davies, M., & Hawkins, S. 1998. Mucus<br />

from marine molluscs. In: Advances in <strong>Marine</strong><br />

Biology, eds. Blaxter, J., Southward, A., &<br />

Tyler, P., 1-71. San Diego: Academic Press.<br />

Denny, M. 1989. Invertebrate mucous secretions:<br />

Functional alternatives to vertebrate<br />

paradigms. Symp Soc Exp Biol. 43:337-66.<br />

de Nys, R., & Steinberg, P. 2002. Linking<br />

marine biology <strong>and</strong> biotechnology. Curr Opin<br />

Biotech. 13:244-8. doi: 10.1016/S0958-<br />

1669(02)00311-7.<br />

Dexter, S. 1979. Influence of substratum<br />

critical surface tension on bacterial adhesion—<br />

in situ studies. J Colloid Interf Sci. 70:<br />

346-54. doi: 10.1016/0021-9797(79)90038-9.<br />

Dixon, D., Sole-Cava, A., Pascoe, P., &<br />

Holl<strong>and</strong>, P. 1995. Periostracal adventitious<br />

hairs on spat of the mussel Mytilus edulis.<br />

J Mar Biol Assoc UK. 75:363-72.<br />

doi: 10.1017/S0025315400018233.<br />

Dobretsov, S., Dahms, H., & Qian, P. 2006.<br />

Inhibition of biofouling by marine microorganisms<br />

<strong>and</strong> their metabolites. Biofouling.<br />

22:43-54. doi: 10.1080/08927010500504784.<br />

Dobretsov, S., Dahms, H-U., Tsoi, M., &<br />

Qian, P-Y. 2005. Chemical control of epibiosis<br />

by Hong Kong sponges: The effect of sponge<br />

extracts on micro- <strong>and</strong> macrofouling communities.<br />

Mar Ecol Prog Ser. 297:119-29.<br />

doi: 10.3354/meps297119.<br />

Donskoy, D., & Ludyanskiy, M. 1995. Low<br />

frequency sound as a control measure for zebra<br />

mussel fouling. In: Proceedings of the 5th<br />

International Zebra Mussel <strong>and</strong> Other Aquatic<br />

Nuisance Organisms Conference. pp. 103-112.<br />

Pembroke, Ontario: Professional Edge.<br />

Dyrynda, P. 1986. Defensive strategies of<br />

modular organisms. Philos T Roy Soc Lon B.<br />

313:227-43. doi: 10.1098/rstb.1986.0035.<br />

Ebran, N., Julien, S., Orange, N., Auperin,<br />

B., & Molle, G. 2000. Isolation <strong>and</strong> characterization<br />

of novel glycoproteins from fish<br />

epidermal mucus: Correlation between their<br />

pore-forming properties <strong>and</strong> their antibacterial<br />

activities. Biochim Biophys Acta. 1467:271-80.<br />

doi: 10.1016/S0005-2736(00)00225-X.<br />

Efimenko, K., Finlay, J., Callow, M., Callow,<br />

J., & Genzer, J. 2009. Development <strong>and</strong><br />

testing of hierarchically wrinkled coatings<br />

for marine antifouling. ACS Appl Mat Int.<br />

1:1031-40. doi: 10.1021/am9000562.<br />

Evans, L., & Clarkson, N. 1993. Antifouling<br />

strategies in the marine environment. J Appl<br />

Bacteriol Symp Suppl. 74:119S-24S.<br />

Faimali, M., Garaventa, F., Terlizzi, A.,<br />

Chiantore, M., & Cattaneo-Vietti, R. 2004.<br />

The interplay of substrate nature <strong>and</strong> biofilm<br />

formation in regulating Balanus amphitrite<br />

Darwin, 1854 larval settlement. J Exp Mar<br />

Biol Ecol. 306:37-50. doi: 10.1016/j.jembe.<br />

2003.12.019.<br />

Feng, D., Ke, C., Li, S., Lu, C., & Guo, F.<br />

2009. Pyrethroids as promising marine antifoulants:<br />

Laboratory <strong>and</strong> field studies. Mar<br />

Biotechnol. 11:153-60. doi: 10.1007/s10126-<br />

008-9130-9.<br />

Finlay, J., Callow, M., Ista, L., Lopez, G.,<br />

& Callow, J. 2002. The influence of surface<br />

wettability on the adhesion strength <strong>and</strong><br />

settled spores of the green algae Enteromorpha<br />

<strong>and</strong> the diatom Amphora. Integr Comp Biol.<br />

42:1116-22. doi: 10.1093/icb/42.6.1116.<br />

Fish, F., & Rohr, J. 1999. Review of<br />

dolphin hydrodynamics <strong>and</strong> swimming<br />

performance. Technical Report no. 1801. SSC.<br />

136 pp.<br />

Floerl, O., Inglis, G., & Marsh, H. 2005.<br />

Selectivity in vector management <strong>and</strong> investigation<br />

of the effectiveness of measures used to<br />

prevent transport of non-indigenous species.<br />

Biol Invasions. 7:459-75. doi: 10.1007/<br />

s10530-004-4863-5.<br />

224 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Fusetani, N. 2004. Biofouling <strong>and</strong> antifouling.<br />

Nat Prod Res. 21:94-104. doi: 10.1039/<br />

b302231p.<br />

Glantz, P., Baier, R., Attstrom, R., Meyer, A.,<br />

& Gucinski, H. 1991. Comparative clinical<br />

wettability of teeth <strong>and</strong> intraoral mucosa.<br />

J Adhes Sci Technol. 5:401-8. doi: 10.1163/<br />

156856191X00459.<br />

Gohad, N., Shah, N., Metters, A., & Mount,<br />

A. 2010. Noradrenaline deters marine invertebrate<br />

biofouling when covalently bound<br />

in polymeric coatings. J Exp Mar Biol Ecol.<br />

394:63-73. doi: 10.1016/j.jembe.2010.07.014.<br />

Guenther, J., Heimann, K., & de Nys, R.<br />

2007. Pedicellariae of the crown-of-thorns sea<br />

star Acanthaster planci are not an effective<br />

defence against fouling. Mar Ecol Prog Ser.<br />

340:101-108. doi: 10.3354/meps340101.<br />

Guo, S., Lee, H., Chaw, K., Miklas, J., Teo,<br />

S., Dickinson, G., … Khoo, B. 2011. Effect<br />

of ultrasound on cyprids <strong>and</strong> juvenile barnacles.<br />

Biofouling. 27:185-92. doi: 10.1080/<br />

08927014.2010.551535.<br />

Hall, J., Giddings, J., Solomon, K., & Balcomb,<br />

R. 1994. An ecological risk assessment for the<br />

use of Irgarol 1051 as an algaecide for antifouling<br />

paints. Crit Rev Toxicol. 29:367-437.<br />

Hay, M. 1996. <strong>Marine</strong> chemical ecology:<br />

What’s known <strong>and</strong> what’s next J Exp Mar<br />

Biol Ecol. 200:103-34. doi: 10.1016/S0022-<br />

0981(96)02659-7.<br />

Hellio, C., Berge, J., Beaupoil, C., Le Gal, Y.,<br />

& Bourgougnon, N. 2002. Screening of<br />

marine algal extracts for anti-settlement<br />

activities against microalgae <strong>and</strong> macroalgae.<br />

Biofouling. 18:205-15. doi: 10.1080/<br />

08927010290010137.<br />

Hirose, E., Yamashiro, H., & Mori, Y. 2001.<br />

Properties of tunic acid in the ascidian Phallusia<br />

nigra (Ascidiidae, Phlebobranchia). Zool Sci.<br />

18:309-14. doi: 10.2108/zsj.18.309.<br />

Holm, E., Cannon, G., Roberts, D., Schmidt,<br />

A., Sutherl<strong>and</strong>, J., & Rittschof, D. 1997.<br />

The influence of initial surface chemistry on<br />

development of the fouling community at<br />

Beaufort, North Carolina. J Exp Mar Biol<br />

Ecol. 215:189-203. doi: 10.1016/S0022-<br />

0981(97)00040-3.<br />

Holmstrom, C., & Kjelleberg, S. 1994. The<br />

effect of external biological factors on settlement<br />

of marine invertebrate <strong>and</strong> new antifouling<br />

technology. Biofouling. 8:147-60.<br />

doi: 10.1080/08927019409378269.<br />

Holmstrom, C., Rittschof, D., & Kjelleberg,<br />

S. 1992. Inhibition of settlement by larvae<br />

of Balanus amphitrite <strong>and</strong> Ciona intestinalis<br />

by a surface-colonizing marine bacterium.<br />

Appl Environ Microb. 58:2111-5.<br />

Holmstrom, C., Steinberg, P., Christov, V.,<br />

Christie, G., & Kjelleberg, S. 2000. Bacteria<br />

immobilised in gels: Improved methodologies<br />

for antifouling <strong>and</strong> biocontrol applications.<br />

Biofouling. 15:109-17. doi: 10.1080/<br />

08927010009386302.<br />

Ingle, M. 2007. Naval Sea Systems Comm<strong>and</strong><br />

Antifouling Program. Presented at the National<br />

Paint & Coatings Association International<br />

<strong>Marine</strong> <strong>and</strong> Offshore Coatings Expo. Virginia<br />

Beach, VA.<br />

Keats, D., Knight, M., & Pueschel, C. 1997.<br />

Antifouling effects of epithallial shedding in<br />

three crustose coralline algae (Rhodophyta,<br />

Coralinales) on a coral reef. J Exp Mar Biol<br />

Ecol. 213:281-93. doi: 10.1016/S0022-<br />

0981(96)02771-2.<br />

Kohler, J., Hansen, P., & Wahl, M. 1999.<br />

Colonization patterns at the substratum-water<br />

interface: How does surface microtopography<br />

influence recruitment patterns of sessile<br />

organisms Biofouling. 14:237-48. doi: 10.1080/<br />

08927019909378415.<br />

Krishnan, S., Wang, N., Ober, C., Finlay, J.,<br />

Callow, M., Callow, J., … Fischer, D. 2006.<br />

Comparison of the fouling release properties<br />

of hydrophobic fluorinated <strong>and</strong> hydrophilic<br />

PEGylated block copolymer surfaces: Attachment<br />

strength of the diatom Navicula <strong>and</strong><br />

the green algae Ulva. Biomacromolecules.<br />

7:1449-62. doi: 10.1021/bm0509826.<br />

Krug, P. 2006. Defense of benthic invertebrates<br />

against surface colonization by larvae: A<br />

chemical arms race. In: Progress in Molecular<br />

<strong>and</strong> Subcellular Biology Subseries <strong>Marine</strong><br />

Molecular Biotechnology, eds. Fusetani, N.,<br />

& Clare, A., 1-53. Berlin Heidelberg: Springer-<br />

Verlag.<br />

Lau, S., & Qian, P. 1997. Phlorotannins<br />

<strong>and</strong> related compounds as larval settlement<br />

inhibitors of the tube-building polychaete<br />

Hydroides elegans. Mar Ecol Prog Ser.<br />

159:219-27. doi: 10.3354/meps159219.<br />

Long, C., Schumacher, J., Robinson, P., III,<br />

Finlay, J., Callow, M., Callow, J., & Brennan,<br />

A. 2010. A model that predicts the attachment<br />

behavior of Ulva linza zoospores on surface<br />

topography. Biofouling. 26:411-9. doi: 10.1080/<br />

08927011003628849.<br />

Marechal, J-P., & Hellio, C. 2009. Challenges<br />

for the development of new non-toxic antifouling<br />

solutions. Int J Mol Sci. 10:4623-37.<br />

doi: 10.3390/ijms10114623.<br />

Marquis, C., Baird, A., de Nys, R., Holmstrom,<br />

C., & Koziumi, N. 2005. An evaluation<br />

of the antimicrobial properties of the eggs<br />

of 11 species of scleractinian corals. Coral<br />

Reefs. 24:248-53. doi: 10.1007/s00338-005-<br />

0473-7.<br />

McKenzie, J., & Grigolava, I. 1996. The<br />

echinoderm surface <strong>and</strong> its role in preventing<br />

microfouling. Biofouling. 10:261-72.<br />

doi: 10.1080/08927019609386285.<br />

Meyer, A., Baier, R., & King, R. 1988. Initial<br />

fouling of nontoxic coatings in fresh, brackish,<br />

<strong>and</strong> sea water. Can J Chem Eng. 66:55-62.<br />

doi: 10.1002/cjce.5450660108.<br />

Meyer, W., & Seegers, U. 2004. A preliminary<br />

approach to epidermal antimicrobial<br />

defense in the Delphinidae. Mar Biol.<br />

144:841-4. doi: 10.1007/s00227-003-<br />

1256-8.<br />

Molino, P., & Weatherbee, R. 2008. The<br />

biology of biofouling diatoms <strong>and</strong> their role<br />

in the development of microbial slimes.<br />

Biofouling. 24:365-79. doi: 10.1080/<br />

08927010802254583.<br />

Naval Ships’ Technical Manual. 2006.<br />

Chapter 081—Waterborne underwater hull<br />

cleaning of Navy ships. Publication # S9086-<br />

CQ-STM-010/CH-081 Revision 5. Washington,<br />

DC: Naval Sea Systems Comm<strong>and</strong>.<br />

July/August 2011 Volume 45 Number 4 225


Nehring, S. 2001. After the TBT era: Alternative<br />

anti-fouling paints <strong>and</strong> their ecological<br />

risks. Senck Marit. 31:341-51.<br />

Nylund, G., & Pavia, H. 2005. Chemical<br />

versus mechanical inhibition of fouling in the<br />

red alga Dilsea carnosa. Mar Ecol Prog Ser.<br />

299:111-21. doi: 10.3354/meps299111.<br />

Olsen, S., Pedersen, L., Laursen, M., Kiil, S.,<br />

& Dam-Johansen, K. 2007. Enzyme-based<br />

antifouling coatings: A review. Biofouling.<br />

23:369-83. doi: 10.1080/08927010701566384.<br />

Omae, I. 2003. Organotin antifouling paints<br />

<strong>and</strong> their alternatives. Appl Organometal<br />

Chem. 17:81-105. doi: 10.1002/aoc.396.<br />

Omae, I. 2006. General aspects of natural<br />

products antifoulants in the environment.<br />

H<strong>and</strong>book Env Chem. 5:227-62. doi: 10.1007/<br />

698_5_057.<br />

Owen, R., Knap, A., Toaspern, M., & Carbery,<br />

K. 2002. Inhibition of coral photosynthesis<br />

by the antifouling herbicide Irgarol 1051. Mar<br />

Pollut Bull. 44:623-32. doi: 10.1016/S0025-<br />

326X(01)00303-4.<br />

Pawlik, J. 1992. Chemical ecology of the<br />

settlement of benthic marine invertebrates.<br />

Oceanogr Mar Biol Annu Rev. 30:273-335.<br />

Perez, M., Garcia, M., Blustein, G., & Stupak,<br />

M. 2007. Tannin <strong>and</strong> tannate from the<br />

quebracho tree: An eco-friendly alternative<br />

for controlling marine biofouling. Biofouling.<br />

23:151-9. doi: 10.1080/08927010701189484.<br />

Phillippi, A., O’Connor, N., Lewis, A., &<br />

Kim, Y. 2001. Surface flocking as a possible<br />

anti-biofoulant. Aquaculture. 195:225-38.<br />

doi: 10.1016/S0044-8486(00)00556-1.<br />

Piola, R., & Johnston, E. 2008. The potential<br />

for translocation of marine species via small<br />

scale disruptions to antifouling surfaces.<br />

Biofouling. 24:145-55. doi: 10.1080/<br />

08927010801930480.<br />

Poulin, R., & Grutter, A. 1996. Cleaning<br />

symbioses: Proximate <strong>and</strong> adaptive explanations.<br />

Bioscience. 46:512-7. doi: 10.2307/1312929.<br />

Qian, P., Xu, Y., & Fusetani, N. 2010.<br />

Natural products as antifouling compounds:<br />

Recent progress <strong>and</strong> future perspectives.<br />

Biofouling. 26:223-34. doi: 10.1080/<br />

08927010903470815.<br />

Radford, C., Stanley, J., Tindle, C.,<br />

Montgomery, J., & Jeffs, A. 2010. Localised<br />

coastal habitats have distinct underwater<br />

sound signatures. Mar Ecol Prog Ser.<br />

401:21-9. doi: 10.3354/meps08451.<br />

Ralston, E., & Swain, G. 2009. Bioinspiration—the<br />

solution for biofouling control<br />

Bioinspir Biomim. 4:015007. doi: 10.1088/<br />

1748-3182/4/1/015007.<br />

Ramasamy, M., & Murugan, A. 2007. Fouling<br />

deterrent chemical defense in three muricid<br />

gastropod egg masses from the Southeast<br />

coast of India. Biofouling. 23:259-65.<br />

doi: 10.1080/08927010701278857.<br />

Raveendran, T., & Mol, V. 2009. Natural<br />

products antifoulants. Curr Sci. 97:508-20.<br />

Ritchie, K. 2006. Regulation of microbial<br />

populations by coral surface mucus <strong>and</strong><br />

mucus-associated bacteria. Mar Ecol Prog Ser.<br />

322:1-14. doi: 10.3354/meps322001.<br />

Rittschof, D. 2000. Natural product antifoulants:<br />

One perspective on the challenges<br />

related to coating development. Biofouling.<br />

15:119-27. doi: 10.1080/08927010009386303.<br />

Rittschof, D. 2001. Natural product antifoulants<br />

<strong>and</strong> coatings development. In:<br />

<strong>Marine</strong> Chemical Ecology, eds. McClintock,<br />

J., & Baker, B., 543-66. Boca Raton, Florida:<br />

CRC Press.<br />

Rittschof, D., & Costlow, J., Jr. 1989.<br />

Bryozoan <strong>and</strong> barnacle settlement in relation<br />

to initial surface wettability: A comparison<br />

of laboratory <strong>and</strong> field studies. Topics Mar<br />

Biol. 53:411-6.<br />

Scardino, A., & de Nys, R. 2004. Fouling<br />

deterrence on the bivalve shell Mytilus galloprovincialis:<br />

A physical phenomenon. Biofouling.<br />

20:249-57. doi: 10.1080/08927010400016608.<br />

Scardino, A., & de Nys, R. 2011. Minireview:<br />

Biomimetic models <strong>and</strong> bioinspired<br />

surfaces for fouling control. Biofouling. 27:<br />

73-86. doi: 10.1080/08927014.2010.536837.<br />

Scardino, A., Guenther, J., & de Nys, R.<br />

2008. Attachment point theory revisited: The<br />

fouling response to a microtextured matrix.<br />

Biofouling. 24:45-53. doi: 10.1080/<br />

08927010701784391.<br />

Scardino, A., Harvey, E., & de Nys, R. 2006.<br />

Testing attachment point theory: Diatom<br />

attachment on microtextured polyamide biomimetics.<br />

Biofouling. 22:55-60. doi: 10.1080/<br />

08927010500506094.<br />

Scarlett, A., Donkin, P., Fileman, T., &<br />

Morris, R. 1999. Occurrence of the antifouling<br />

herbicide, Irgarol 1051, within coastalwater<br />

seagrasses from Queensl<strong>and</strong>, Australia.<br />

Mar Pollut Bull. 38:687-91. doi: 10.1016/<br />

S0025-326X(99)00003-X.<br />

Schiff, K., Diehl, D., & Valkirs, A. 2004.<br />

Copper emissions from antifouling paint<br />

on recreational vessels. Mar Pollut Bull.<br />

48:371-7. doi: 10.1016/j.marpolbul.2003.<br />

08.016.<br />

Schumacher, J., Aldred, N., Callow, M.,<br />

Finlay, J., Callow, J., Clare, A., & Brennan, A.<br />

2007. Species-specific engineered antifouling<br />

topographies: Correlations between the settlement<br />

of algal zoospores <strong>and</strong> barnacle cyprids.<br />

Biofouling. 23:307-17. doi: 10.1080/<br />

08927010701393276.<br />

Seth, N., Chakravarty, P., Kh<strong>and</strong>eparker, L.,<br />

Anil, A., & P<strong>and</strong>it, A. 2010. Quantification<br />

of the energy required for the<br />

destruction of Balanus Amphitrite larva<br />

by ultrasonic treatment. J Mar Biol Assoc<br />

UK. 90:1475-82. doi: 10.1017/<br />

S0025315409991548.<br />

Shephard, K. 1994. Functions for fish<br />

mucus. Rev Fish Biol Fisher. 4:401-29.<br />

doi: 10.1007/BF00042888.<br />

Simpson, S., Meekan, M., Jeffs, A.,<br />

Montgomery, J., & McCauley, R. 2008.<br />

Settlement-stage coral reef fish prefer the<br />

higher-frequency invertebrate-generated<br />

audible component of reef noise. Anim<br />

Behav. 75:1861-8. doi: 10.1016/j.anbehav.<br />

2007.11.004.<br />

Stanley, J., Radford, C., & Jeffs, A. 2010.<br />

Induction of settlement in crab megalopae by<br />

ambient underwater reef sound. Behav Ecol.<br />

21:113-20. doi: 10.1093/beheco/arp159.<br />

226 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


Statz, A., Finlay, J., Dalsin, J., Callow, M.,<br />

Callow, J., & Messersmith, P. 2006. Algal<br />

antifouling <strong>and</strong> fouling release properties of<br />

metal surfaces coated with a polymer inspired<br />

by marine mussels. Biofouling. 22:391-9.<br />

doi: 10.1080/08927010601004890.<br />

Stoeker, D. 1980. Relationships between<br />

chemical defense <strong>and</strong> ecology in benthic<br />

ascidians. Mar Ecol Prog Ser. 3:257-65.<br />

doi: 10.3354/meps003257.<br />

Swain, G. 1999. Redefining antifouling<br />

coatings. J Prot Coat Linings. 16:26-35.<br />

Swain, G. 2010. The importance of ship<br />

hull coatings <strong>and</strong> maintenance as drivers for<br />

environmental sustainability. In: Ship Design<br />

<strong>and</strong> Operation for Environmental Sustainability.<br />

pp 8. London: Royal Institute of<br />

Naval Architects.<br />

Swain, G., Kovach, B., Touzot, A., Casse, F.,<br />

& Kavanagh, C. 2007. Measuring the<br />

performance of today’s antifouling coatings.<br />

J Ship Prot. 23:164-170.<br />

Targett, N., Bishop, S., O’Connell, O., &<br />

Yoder, J. 1983. Antifouling agents against the<br />

benthic marine diatom, Navicula salinicola<br />

Homarine from the gorgonians Leptogorgia<br />

virgulata <strong>and</strong> L. setacea <strong>and</strong> analogs. J Chem<br />

Ecol. 9:817-29. doi: 10.1007/BF00987807.<br />

Thomas, K., & Brooks, S. 2010. The environmental<br />

fate <strong>and</strong> effects of antifouling paint<br />

biocides. Biofouling. 26:73-88. doi: 10.1080/<br />

08927010903216564.<br />

Thomas, K., Fileman, T., Readman, J., &<br />

Waldock, M. 2001. Antifouling paint booster<br />

biocides in the UK coastal environment <strong>and</strong><br />

potential risks of biological effects. Mar<br />

Pollut Bull. 42:677-88. doi: 10.1016/S0025-<br />

326X(00)00216-2.<br />

Tribou, M., & Swain, G. 2010. The use of<br />

proactive in-water grooming to improve the<br />

performance of ship hull antifouling coatings.<br />

Biofouling. 26:47-56. doi: 10.1080/<br />

08927010903290973.<br />

Videler, H., Geertjes, G., & Videler, J. 1999.<br />

Biochemical characteristics <strong>and</strong> antibiotic<br />

properties of the mucus envelope of the<br />

queen parrotfish. J Fish Biol. 54:1124-7.<br />

doi: 10.1111/j.1095-8649.1999.tb00864.x.<br />

Vrolijk, N., Targett, N., Baier, R., & Meyer,<br />

A. 1990. Surface characterisation of two<br />

gorgonian coral species: Implications for a<br />

natural antifouling defence. Biofouling.<br />

2:39-54. doi: 10.1080/08927019009378128.<br />

Wahl, M. 1989. <strong>Marine</strong> epibiosis. I. Fouling<br />

<strong>and</strong> antifouling: Some basic aspects. Mar<br />

Ecol Prog Ser. 58:175-89. doi: 10.3354/<br />

meps058175.<br />

Wahl, M., Kroger, K., & Lenz, M. 1998.<br />

Non-toxic protection against epibiosis.<br />

Biofouling. 12:205-26. doi: 10.1080/<br />

08927019809378355.<br />

Wahl, M., & Sonnichsen, H. 1992. <strong>Marine</strong><br />

epibiosis. IV. The periwinkle Littorina littorea<br />

lacks typical antifouling defences—why are<br />

some populations so little fouled Mar Ecol<br />

Prog Ser. 88:225-35. doi: 10.3354/meps088225.<br />

Walters, L., Smith, C., & Hadfield, M. 2003.<br />

Recruitment of sessile marine invertebrates<br />

on Hawaiian macrophytes: Do pre-settlement<br />

or post-settlement processes keep plants free<br />

from fouling Bull Mar Sci. 72:813-39.<br />

Walters, L., & Wethey, D. 1996. Settlement<br />

<strong>and</strong> early post-settlement survival of sessile<br />

marine invertebrates on topographically<br />

complex surfaces: The importance of refuge<br />

dimensions <strong>and</strong> adult morphology. Mar<br />

Ecol Prog Ser. 137:161-71. doi: 10.3354/<br />

meps137161.<br />

Wikstrom, S., & Pavia, H. 2004. Chemical<br />

settlement inhibition versus post-settlement<br />

mortality as an explanation for differential<br />

fouling of two congeneric seaweeds. Oecologia.<br />

138:223-30. doi: 10.1007/s00442-003-1427-9.<br />

Xu, Q., Barrios, C., Cutright, T., & Newby,<br />

B. 2005. Assessment of antifouling effectiveness<br />

of two natural product antifoulants<br />

by attachment study with freshwater bacteria.<br />

Environ Sci Pollut Res. 12:278-84.<br />

doi: 10.1065/espr2005.04.244.<br />

Yebra, D., Kiil, S., & Dam-Johansen, K. 2004.<br />

Antifouling technology—past, present <strong>and</strong><br />

future steps towards efficient <strong>and</strong> environmentally<br />

friendly antifouling coatings. Prog Org<br />

Coat. 50: 75-104. doi: 10.1016/j.porgcoat.<br />

2003.06.001.<br />

Yee, L., Holmstrom, C., Fuary, E., Lewin, N.,<br />

Kjelleberg, S., & Steinberg, P. 2007. Inhibition<br />

of fouling by marine bacteria immobilised<br />

in K-carrageenan beads. Biofouling. 23:<br />

287-94. doi: 10.1080/08927010701366280.<br />

July/August 2011 Volume 45 Number 4 227


BOOK REVIEW<br />

Sex, Drugs, <strong>and</strong> Sea Slime: The Oceans’<br />

Oddest Creatures <strong>and</strong> Why They Matter<br />

By Ellen Prager<br />

University of Chicago Press, April 15, 2011 (International Publication: May 15, 2011)<br />

184 pp., $26.00 (Hardcover)<br />

Reviewed by Jason Goldberg<br />

U.S. Fish <strong>and</strong> Wildlife Service<br />

When it comes to reviewing this book,<br />

please let me be c<strong>and</strong>id about a personal<br />

bias: as a volunteer at the Smithsonian’s<br />

National Museum of Natural History,<br />

I’ve been known to give tours focused<br />

on the Sant Ocean Hall’s oddities, such<br />

as the two-horned narwhal skull, giant<br />

squid, <strong>and</strong> the seadevil. There’s a method<br />

to such madness: if you can capture a lay<br />

audience’s attention with the exciting <strong>and</strong><br />

unusual, they may be more receptive to<br />

discussing more serious messages about<br />

the importance of our oceans. With Sex,<br />

Drugs, <strong>and</strong> Sea Slime, Dr. Ellen Prager has<br />

unquestionably written one of the more<br />

bizarre <strong>and</strong> fascinating books in ocean literature.<br />

Combine Tina Fey with Carl<br />

Sagan or perhaps Mel Brooks <strong>and</strong> Rachel<br />

Carson <strong>and</strong> you have this book. Prager’s<br />

writing is reminiscent of the works of<br />

Mary Roach, Anthony Aveni, <strong>and</strong><br />

Michael Shermer, all of whom also have<br />

a talent for writing about science’s odder<br />

side. To the best of my knowledge,<br />

however, none has written as eloquently<br />

about, as Prager calls it, the lobster’s<br />

“Super Soaker Pee Blaster.” As she writes,<br />

the purpose of her book is to be “a brief<br />

<strong>and</strong> entertaining look at some of the<br />

oceans’ most fascinating creatures, their<br />

unusual tactics for survival, <strong>and</strong> their invaluable<br />

links to humankind. The end<br />

goal is to showcase the importance of<br />

the great diversity of life in the sea, why<br />

it is at risk, <strong>and</strong> why we should all care.”<br />

With this book, she has succeeded exceptionally<br />

well in achieving her goals.<br />

Everyone in the <strong>Marine</strong> <strong>Technology</strong><br />

<strong>Society</strong> likely has some favorite story of a<br />

bizarre creature or other factoid about<br />

the oceans they learned while in school<br />

or on the job. Prager has thoroughly researched<br />

<strong>and</strong> captured the best of these<br />

tales. More importantly, what she has<br />

really done is write entertainingly about<br />

why the ocean is relevant regardless of<br />

where you might live. She has an ability<br />

to wax poetic about the ocean’s strangest<br />

denizens, whether it is the unassuming<br />

dinoflagellate or the majestic humpback<br />

whale. The range of material she covers<br />

in such a slender volume is really quite astonishing.<br />

I often found myself wondering<br />

as I read the book whether she would cover<br />

this species or that, <strong>and</strong> inevitably she did.<br />

Plankton, hagfishes, corals, eels, parrotfish,<br />

conesnails, cephalopods, kelp, <strong>and</strong> more—<br />

they’re all in here, along with all the<br />

(copious) mucus <strong>and</strong> other excretions<br />

they produce. As the book’s title indicates,<br />

there’s also plenty of sex, as well as a few sex<br />

changes. For those of you who might be<br />

wondering, yes, she included the pearlfish,<br />

a species that provides clear proof that<br />

evolution has a sense of humor.<br />

Some chapters focus on specific species,<br />

such as those on plankton or the<br />

denizens of a coral reef, while others target<br />

specialized functions, such as species that<br />

might compete in the “X-Games” or live<br />

in extreme environments. The descriptions<br />

of the species <strong>and</strong> their unusual habits are<br />

always entertaining <strong>and</strong> sometimes laughout-loud<br />

funny. She then turns serious at<br />

the end of each chapter when she covers<br />

why these species matter. I was very<br />

pleased to see that Prager doesn’t just<br />

cover the usual reasons, such as food,<br />

drugs, <strong>and</strong> recreation, but that she highlights<br />

things many people don’t realize,<br />

such as the value of corals in mitigating<br />

storm damages. She concludes the book<br />

onamoreseriousbutoptimisticnote,<br />

highlighting the dangers that still lurk in<br />

ocean conservation <strong>and</strong> suggesting actions<br />

we can all take so we can continue to enjoy<br />

the ocean for generations to come.<br />

Overall, Prager’s work makes for fascinating<br />

reading for anyone interested in<br />

marine science. While it does sometimes<br />

get a little technical, it is certainly appropriate<br />

for lay audiences. It’s also an invaluable<br />

resource for anyone who talks<br />

about the oceans <strong>and</strong> needs some good references<br />

to spice up their talk, although the<br />

inclusion of an index would have been<br />

beneficial for such purposes. If you want<br />

to get another sense of her book, you can<br />

download a free National Public Radio<br />

podcast from http://www.npr.org/<br />

2011/04/07/135043954/under-the-seasex-is-slimy-business.<br />

Thesamecombination<br />

of titillating humor <strong>and</strong> practical<br />

discussion of the ocean’s valueisdemonstrated<br />

in Prager’s interview.<br />

Our livelihoods depend on the ocean,<br />

<strong>and</strong> the talent that has helped develop <strong>and</strong><br />

effectively use technology is extraordinary.<br />

It is therefore incumbent upon each of us<br />

to be able to talk in some way with the<br />

public <strong>and</strong> decision-makers about how<br />

228 <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Journal


the well-being of everyone living on Terra<br />

Firma relies on blue, brown, or white<br />

water. Books such as Prager’s offer an outline<br />

that many of us can use to foster a discussion<br />

about our own respective fields.<br />

Yes, the book is about weird science. The<br />

giggle factor is unmistakable. Even so, each<br />

time it gets weird, it casts light on the wonders<br />

of the ocean <strong>and</strong> makes that weirdness<br />

important <strong>and</strong> meaningful. Perhaps for<br />

that reason, more than any other, the<br />

book is highly recommended.<br />

July/August 2011 Volume 45 Number 4 229


UPCOMING MTS JOURNAL ISSUES<br />

September/October 2011<br />

General Issue<br />

November/December 2011<br />

Legacy Underwater Munitions: Assessment,<br />

Evaluation of Impacts, <strong>and</strong> Potential<br />

Response Technologies<br />

Guest Editors: Geoffrey Carton <strong>and</strong> Terrance Long<br />

CALL FOR PAPERS<br />

The MTS Journal<br />

includes technical<br />

papers, notes <strong>and</strong><br />

commentaries<br />

of general interest in<br />

most issues.<br />

Contact Managing Editor<br />

Amy Morgante<br />

(morganteeditorial@verizon.net)<br />

if you have material you<br />

would like considered.<br />

Specifications for<br />

submitting a manuscript<br />

are on the MTS website<br />

under the Publications menu.<br />

www.mtsociety.org<br />

E-mail:<br />

morganteeditorial@verizon.net<br />

Or visit our homepage at<br />

Check the <strong>Society</strong> website for future Journal topics.<br />

www.mtsociety.org<br />

www.mtsociety.org


Open Positions<br />

The MTS Career Center is an<br />

online resource to help employers<br />

<strong>and</strong> job seekers make career<br />

connections in the marine<br />

technology industry. Career<br />

professionals bring fresh ideas<br />

<strong>and</strong> new energy, <strong>and</strong> can keep<br />

your company current with modern<br />

trends <strong>and</strong> tactics.<br />

MTS Career Center<br />

Visit www.mtsociety.org/careers<br />

The MTS Career Center Provides:<br />

• Unmatched, Targeted Exposure for Job Postings<br />

MTS represents the largest audience of qualied marine technology<br />

industry professionals.<br />

• Easy Online Job & Applicant Management<br />

You can enter job descriptions, check the status of postings, renew or<br />

discontinue postings <strong>and</strong> even make payments online. Plus, you can<br />

easily manage your applicants through automatic email noti cations<br />

<strong>and</strong> applicant history access.<br />

• Resume Database Access<br />

With a paid job posting, you can search the resume database <strong>and</strong><br />

use an automatic notication system to receive email notications<br />

when new resumes match your criteria.<br />

• Employment Br<strong>and</strong>ing Opportunities<br />

Along with each job posting, you can include information about your<br />

company <strong>and</strong> a link to your website.<br />

International <strong>Marine</strong> Forensics Symposium<br />

April 3 – 5, 2012 • National Harbor, MD<br />

The Symposium is sponsored by the<br />

<strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong>,<br />

American <strong>Society</strong> of Naval Engineers,<br />

<strong>Society</strong> of Naval Architects & <strong>Marine</strong> Engineers.<br />

This symposium will bring together marine professionals<br />

<strong>and</strong> historians to exchange information on historic marine<br />

losses, on marine forensic investigation processes <strong>and</strong><br />

tools, <strong>and</strong> on case studies where causes of failures <strong>and</strong><br />

losses have been determined or are under continued<br />

study. The Symposium will report on the latest research<br />

<strong>and</strong> underst<strong>and</strong>ing of the Titanic, Lusitania, Edmund<br />

Fitzgerald, the Monitor <strong>and</strong> Passaic, HMS Prince of<br />

Wales, Bismarck, HMS Hood, <strong>and</strong> the Andrea Doria.<br />

Keynote speaker James<br />

Cameron (conrmed).<br />

His undersea documentaries<br />

include Expedition Bismarck,<br />

Ghosts of the Abyss,<br />

Volcanoes of the Deep Sea,<br />

Aliens of the Deep <strong>and</strong><br />

Last Mysteries of the Titanic.<br />

The International <strong>Marine</strong> Forensics Symposium will be held at the Gaylord Hotel, National Harbor, MD.<br />

Exhibition space is available at $650 for a 10 x 10-foot space. Freeman Exhibit Services will h<strong>and</strong>le the<br />

show decorating. A number of Sponsorship Opportunities are also available to maximize your impact<br />

with Symposium attendees. Contact Mary Beth Loutinsky at mbloutinsky@gmail.com for details.


Oceans of Opportunity:<br />

International Cooperation <strong>and</strong> Partnerships Across the Pacic<br />

September 19–22, 2011<br />

Kona, Hawaii<br />

Register Today for Outst<strong>and</strong>ing Topics in Oceanology,<br />

the Latest <strong>Technology</strong> in the Exhibition Hall, <strong>and</strong><br />

Excellent Networking Opportunities!<br />

go to:<br />

http://www.oceans11mtsieeekona.org/main.cfm/CID/16/Registration/<br />

Of cial Notice of <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Annual Meeting<br />

2011 Ofcer Elections <strong>and</strong> Notice of Annual Membership Meeting<br />

The terms of the individuals holding the following<br />

<strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> (“MTS”) ofces<br />

end on December 31, 2011: Vice President of<br />

Publications, Vice President of Industry <strong>and</strong><br />

<strong>Technology</strong>, Vice President of Education <strong>and</strong><br />

Research, <strong>and</strong> Vice President of Government <strong>and</strong><br />

Public Affairs. Elections for these ofces will<br />

be conducted at the MTS Annual Membership<br />

Meeting <strong>and</strong> Awards Luncheon held in conjunction<br />

with the OCEANS’11 MTS/IEEE Kona<br />

Conference, September 21, 2011, at the Hilton<br />

Waikoloa Village, Kona, Hawaii. The meeting will<br />

begin at noon PST.<br />

Article IV, Section 1.6 of the MTS Bylaws denes<br />

the nominations <strong>and</strong> election procedures for MTS<br />

ofcer elections. On June 20, 2011, the MTS Nominating<br />

Committee presented MTS’s President-<br />

Elect with at least two (2) nominees for each of the<br />

ofces to be elected at this year’s Annual Meeting.<br />

After reviewing these nominations, the President-<br />

Elect directed MTS staff to prepare the ballot for<br />

this year’s election <strong>and</strong> Member Proxy, both of<br />

which contain the names of the nominees for each<br />

ofce to be elected at this year’s election. The<br />

MTS Bylaws establish August 1, 2011 as the record<br />

date for determining the MTS members entitled to<br />

notice of the Annual Meeting <strong>and</strong> eligible to vote in<br />

this year’s election. If you are an active member of<br />

MTS as of August 1, 2011 (membership dues paid),<br />

you are a member in good st<strong>and</strong>ing <strong>and</strong> therefore<br />

entitled to vote in this year’s ofcer elections.<br />

Pursuant to MTS’s Bylaws, a member entitled to<br />

vote may do so in person at the Annual Meeting<br />

or by appointing a proxy to vote on the member’s<br />

behalf. If you wish to appoint a proxy to vote on<br />

your behalf at the MTS Annual Membership<br />

Meeting, please complete the online Member<br />

Proxy (http://www.votenet.com).<br />

This Proxy form must be completed no later than<br />

5:00 p.m. (Eastern St<strong>and</strong>ard Time) on September<br />

11, 2011 (“Proxy Deadline”). Proxies that are<br />

incomplete or received after the Proxy Deadline<br />

will not be accepted or counted. A hardcopy proxy<br />

form may be requested by calling the MTS ofce at<br />

410.884.5330.


<strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> Member Organizations<br />

CORPORATE MEMBERS<br />

ABCO Subsea<br />

Houston, Texas<br />

AMETEK Sea Connect Products, Inc.<br />

Westerly, Rhode Isl<strong>and</strong><br />

C & C Technologies, Inc.<br />

Lafayette, Louisiana<br />

C-MAR Group<br />

Houston, Texas<br />

Compass Publications, Inc.<br />

Arlington, Virginia<br />

Converteam<br />

Houston, Texas<br />

Delcor USA<br />

Houston, Texas<br />

DOF Subsea USA<br />

Houston, Texas<br />

Dynacon, Inc.<br />

Bryan, Texas<br />

E.H. Wachs Company<br />

Houston, Texas<br />

Fluor Offshore Solutions<br />

Sugar L<strong>and</strong>, Texas<br />

Fugro Chance, Inc.<br />

Lafayette, Louisiana<br />

Fugro Geoservices, Inc.<br />

Houston, Texas<br />

Fugro-McClell<strong>and</strong> <strong>Marine</strong> Geosciences<br />

Houston, Texas<br />

Fugro Pelagos, Inc.<br />

San Diego, California<br />

Geospace Offshore Cables<br />

Houston, Texas<br />

GS-Hydro US<br />

Houston, Texas<br />

HARRIS CapRock Communications<br />

Melbourne, Florida<br />

Hydroid, LLC<br />

Pocasset, Massachusetts<br />

Innerspace Corporation<br />

Covina, California<br />

INTECSEA<br />

Houston, Texas<br />

InterMoor, Inc.<br />

Houston, Texas<br />

iRobot Corporation<br />

Durham, North Carolina<br />

J P Kenny, Inc.<br />

Houston, Texas<br />

Kongsberg Maritime, Inc.<br />

Houston, Texas<br />

L-3 Communications<br />

Houston, Texas<br />

L-3 MariPro<br />

Goleta, California<br />

Lockheed Martin Sippican<br />

Marion, Massachusetts<br />

<strong>Marine</strong> Cybernetics AS<br />

Trondheim, Norway<br />

Mitsui Engineering <strong>and</strong> Shipbuilding Co. Ltd.<br />

Tokyo, Japan<br />

Mohr Engineering & Testing<br />

Houston, Texas<br />

Oceaneering Advanced Technologies<br />

Hanover, Maryl<strong>and</strong><br />

Oceaneering International, Inc.<br />

Houston, Texas<br />

Odyssey <strong>Marine</strong> Exploration<br />

Tampa, Florida<br />

Oil States Industries, Inc.<br />

Arlington, Texas<br />

Perry Slingsby Systems, Inc.<br />

Houston, Texas<br />

Phoenix International Holdings, Inc.<br />

Largo, Maryl<strong>and</strong><br />

QinetiQ North America – <strong>Technology</strong> Solutions<br />

Group<br />

Slidel, Louisiana<br />

Raytheon Technical Service Company, LLC<br />

Saipem America, Inc.<br />

Houston, Texas<br />

Schilling Robotics, LLC<br />

Davis, California<br />

SEA CON Brantner <strong>and</strong> Associates, Inc.<br />

El Cajon, California<br />

SonTek/YSI, Inc.<br />

San Diego, California<br />

South Bay Cable Corp.<br />

Idyllwild, California<br />

Subconn, Inc.<br />

Burwell, Nebraska<br />

Subsea 7 (US), LLC<br />

Houston, Texas<br />

Technip<br />

Houston, Texas<br />

Teledyne Geophysical Instruments<br />

Houston, Texas<br />

Teledyne Oil & Gas<br />

Daytona Beach, Florida<br />

Teledyne RD Instruments, Inc.<br />

Poway, California<br />

Tyco Electronics Subsea Communications<br />

Morristown, New Jersey<br />

UniversalPegasus International, Inc.<br />

Houston, Texas<br />

BUSINESS MEMBERS<br />

A<strong>and</strong>eraa Data Instruments, Inc.<br />

Attleboro, Massachusetts<br />

Ashtead <strong>Technology</strong>, Inc.<br />

Houston, Texas<br />

Bastion Technologies, Inc.<br />

Houston, Texas<br />

Bennex Subsea, Houston<br />

Houston, Texas<br />

BioSonics, Inc.<br />

Seattle, Washington<br />

BIRNS, Inc.<br />

Oxnard, California<br />

C.A. Richards <strong>and</strong> Associates, Inc.<br />

Houston, Texas<br />

Cochrane Technologies, Inc.<br />

Lafayette, Louisiana<br />

Compass Personnel Services, Inc.<br />

Katy, Texas<br />

Contros Systems & Solutions GmbH<br />

Kiel, Germany<br />

DeepSea Power <strong>and</strong> Light<br />

San Diego, California<br />

Deepwater Rental <strong>and</strong> Sypply<br />

New Iberia, Louisiana<br />

DOER <strong>Marine</strong><br />

Alameda, California<br />

DPS Offshore Inc.<br />

Houston, Texas<br />

DTC International<br />

Houston, Texas<br />

Energy Sales, Inc.<br />

Redmond, Washington<br />

Environ-Tech Diving<br />

Stanwood, Washington<br />

Falmat, Inc.<br />

San Marcos, California<br />

Fugro Atlantic<br />

Norfolk, Virginia<br />

Fugro GeoSurveys, Inc.<br />

St. John’s, Newfoundl<strong>and</strong> <strong>and</strong> Labrador,<br />

Canada<br />

Global Industries Offshore, LLC<br />

Houston, Texas<br />

Horizon <strong>Marine</strong>, Inc.<br />

Marion, Massachusetts<br />

ICAN<br />

Mt. Pearl, Newfoundl<strong>and</strong> <strong>and</strong> Labrador, Canada<br />

Intrepid Global, Inc.<br />

Houston, Texas<br />

IPOZ Systems, LLC<br />

Katy, Texas<br />

IVS 3D<br />

Portsmouth, New Hampshire<br />

iXBlue, Inc.<br />

Cambridge, Massachusetts<br />

KDU Worldwide Technical Services<br />

Sarjah, United Arab Emirates<br />

KnightHawk Engineering<br />

Houston, Texas<br />

Liquid Robotics, Inc.<br />

Palo Alto, California<br />

Makai Ocean Engineering, Inc.<br />

Kailua, Hawaii<br />

Matthews-Daniel Company<br />

Houston, Texas<br />

Oceanic Imaging Consultants, Inc.<br />

Honolulu, Hawaii<br />

The <strong>Marine</strong> <strong>Technology</strong> <strong>Society</strong> gratefully acknowledges the critical support of the Corporate, Business, <strong>and</strong> Institutional members listed.<br />

Member organizations have aided the <strong>Society</strong> substantially in attaining its objectives since its inception in 1963.<br />

OceanWorks International<br />

Houston, Texas<br />

Poseidon Offshore Mining<br />

Oslo, Norway<br />

Quest Offshore Resources<br />

Sugar L<strong>and</strong>, Texas<br />

Remote Ocean Systems, Inc.<br />

San Diego, California<br />

RRC Robotica Submarina<br />

Macaé, Brazil<br />

SeaBotix<br />

San Diego, California<br />

SeaL<strong>and</strong>Aire Technologies, Inc.<br />

Jackson, Mississippi<br />

SeaView Systems, Inc.<br />

Dexter, Michigan<br />

Sonardyne, Inc.<br />

Houston, Texas<br />

SIMCorp <strong>Marine</strong> Environmental, Inc.<br />

St. Stephens, Canada<br />

Sound Ocean Systems, Inc.<br />

Redmond, Washington<br />

Stress Subsea, Inc.<br />

Houston, Texas<br />

Subsea Riser Products, Inc.<br />

Houston, Texas<br />

SURF Subsea, Inc.<br />

Magnolia, Texas<br />

Team Trident<br />

Cypress, Texas<br />

<strong>Technology</strong> Systems Corporation<br />

Palm City, Florida<br />

Teledyne Impulse<br />

San Diego, California<br />

Tension Member <strong>Technology</strong><br />

Huntington Beach, California<br />

VideoRay, LLC<br />

Phoenixville, Pennsylvania<br />

WET Labs, Inc.<br />

Philomath, Oregon<br />

Xodus Group<br />

Houston, Texas<br />

INSTITUTIONAL MEMBERS<br />

Associacio Institut Ictineu Centre, Catala De<br />

Recerca Submarina<br />

Barcelona, Spain<br />

Canadian Coast Guard<br />

St. John’s, Newfoundl<strong>and</strong> <strong>and</strong> Labrador, Canada<br />

City of St. John’s<br />

Newfoundl<strong>and</strong> <strong>and</strong> Labrador, Canada<br />

CLS America, Inc.<br />

Largo, Maryl<strong>and</strong><br />

Consortium for Ocean Leadership<br />

Washington, DC<br />

Department of Innovation, Trade <strong>and</strong> Rural<br />

Development<br />

St. John’s, Newfoundl<strong>and</strong> <strong>and</strong> Labrador, Canada<br />

Fundação Homem do Mar<br />

Rio de Janeiro, Brazil<br />

Harbor Branch Oceanographic Institute<br />

Fort Pierce, Florida<br />

International Seabed Authority<br />

Kingston, Jamaica<br />

<strong>Marine</strong> Applied Research & Exploration<br />

Richmond, California<br />

<strong>Marine</strong> Institute<br />

Newfoundl<strong>and</strong> <strong>and</strong> Labrador, Canada<br />

Monterey Bay Aquarium Research Institute<br />

Moss L<strong>and</strong>ing, California<br />

National Research Council Institute for Ocean<br />

<strong>Technology</strong><br />

St. John’s, Newfoundl<strong>and</strong> <strong>and</strong> Labrador,<br />

Canada<br />

Naval Facilities Engineering Service Center<br />

Port Hueneme, California<br />

NOAA/PMEL<br />

Seattle, Washington<br />

Noblis<br />

Falls Church, Virginia<br />

OceanGate, LLC<br />

Everett, Washington<br />

Oregon State University College of Oceanic <strong>and</strong><br />

Atmospheric Sciences<br />

Corvalis, Oregon<br />

<strong>Society</strong> of Ieodo Research<br />

Jeju-City, South Korea


5565 Sterrett Place, Suite 108<br />

Columbia, Maryl<strong>and</strong> 21044<br />

Postage for periodicals<br />

is paid at Columbia, MD,<br />

<strong>and</strong> additional mailing offices.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!