r w: . .quu , .35... ~ ma. r. r b. £ \ v3. 4‘“ 4:.»
Hf.-(s)Q.-(s). (4.32)
V(s) i=1
Hybrid Tail
From (4.1) and (4.2), the tail width S, at the end has a significant impact on the speed U.
One could increase S, by simply using a wider IPMC beam. Due to the IPMC actuation
mechanism, however, a too wide beam (i.e., plate) will produce curling instead of bending
motion and is thus not desirable. Therefore, it has been chosen to increase the edge width
by attaching a passive plastic piece, as illustrated in Fig. 4.3. While such a hybrid tail
is expected to increase the thrust, one has to also consider that the extra hydrodynamic
force on the passive fin adds to the load of IPMC and may reduce the bending amplitude.
Therefore, it is necessary to model these interactions carefully.
Figure 4.3: Illustration of an IPMC beam with a passive fin. The lower schematic shows
the definitions of dimensions.
The hydrodynamic force acting on the passive fin can be written as [76]
ft... (2.5) = {0.st (z)2 r2(w)w(z,s), £0 < z < 1., (4.33)
80
where 13(0)) is the hydrodynamic function of the passive fin. Note that the hydrodynamic
force acting on the active IPMC beam has been incorporated in (4.10), so only the hydro-
dynamic force on the passive fin needs to be considered here. Since the passive fin used is
very light, its inertial mass is negligible compared to the propelled virtual fluid mass and
is thus ignored in the analysis here. Considering that the passive fin is rigid compared to
IPMC, its width b(z) and deflection w(z,s) can be expressed as
b —b
b (2) = ‘ ° (2 —Lo) + be, (434)
L1 —L0
aw(Lo,S)
<92
W(2,8) = W(L0,S) + (z-Lo), (435)
where b0, b1, L, L0, L1 are as defined in Fig. 4.3. Then one can calculate the moment
introduced by the passive fin: for L0 g 2 g L1,
M... (2,.) = [12.1 (as) (r—zwr
L0
L1 L1
= [1.105) (T—Lo)dT+(Lo —2) [16.1 (r.s)d1.
L0 L0
(4.36)
If we define
L1 L1
Mtail (S) = /f1a11(T,S) (T—Lo)d’€, 1”111110) = fftail(TaS)dT, (437)
L0 L0
then (4.36) can be written as
Mfin (2,5) = Mraii (S) +Ftail (S) (L0 - Z)- (438)
Fig. 4.4 shows the forces and moments acting on the hybrid tail.
81
.Y---
Figure 4.4: Forces and moments acting on the hybrid tail.
One can get the generalized force as:
Lo
fzi (S) = 1117, (de(ZaS)¢i(Z)dZ+ 5—(-f,fl. (AA)
Solving for 15(p,s), one obtains
_ p . __ 1 A(S)V(S)
1.(p,s)—p———2_B(S)z.(o,s) p2_3(s) 2 , (A5)
which can be rewritten through partial fraction expansion as:
15(p,S) :
2 p—W+m
0.5 0.5
130,5 ————- —-——-— , A.6
+ ‘ ’b—WUWW) ‘ ’
192
A(s>V Q. a. ((2.1)
Let’s start with the case when k > 0. Because 1(0) 2 —k < 0, 1(1)) is continuous in
(—-IZ,+°°) and
a
b
there exist pl 6 (—Z’O)’ p2 E (O,+oo) such that x(p1) = O and x(p2) = 0. Since
= P
ap+b’
X'(P) (CZ)
withp> —g,wecmgetx’(p)>0whenp>0andx’(p)<0when0>p> ~53. Sox(p)
is monotonically increasing in (0, +00), and monotonically decreasing in ( — — , 0). Then p,
a
and p2 should be unique. When k = O, and 75(0) = 0, then p1 2 p2 = 0. El E1
199
C.2 Derivation of Eq. (5.17)
D
Define Dépz —p1. Eq. (5.12) can then be written as V = b° From Fig. 5.4,
Pz-Pl
= a, C3
772 - Th ( )
where n, and 172 are the n-coordinates corresponding to p1 and p2, respectively, in Fig. 5.4.
Eq. (C.3) implies
b a D a = a. (C4)
;2' (“1(131’2+ ‘) ‘1“(‘5P' + 1))
With p2 = D + p, and (C4), one can solve p, in terms of D,
7'5— - 5,1) > 0
P1 = eb _ 1 - (C-5)
O,D=O
When D -—> O, with 1’ Hospital’ 3 rule,
a
_ 1 b
So p1 is still a continuous function of D. Since
3111(3 +1)=fl— (C6)
from (C5), one can get
b D 1
71m 3 a —b +1 =— GD 3 —k. (07)
a b -D a -D a
eb — 1 eb — 1
200
Since D = bV, one can get Eq. (5.17) from Eq. (C.7). Note that with 1’ Hospital’ 8 rule,
one gets
So k is a continuous function of V.
201
BIBLIOGRAPHY
[1] K. Abdelnour, E. Mancia, S. D. Peterson, and M. Porfiri. Hydrodynamics of under-
water propulsors based on ionic polymer metal composites: A numerical study. Smart
Materials and Structures, 181085006: 1—1 1, 2009.
[2] B. J. Akle, M. D. Bennett, and D. J. Leo. High-strain ionomeric-ionic liquid elec-
troactive actuators. Sensors and Actuators A, 126:173—1 81, 2006.
[3] Barbar J. Akle. Characterization and Modeling of the Ionomer—Conductor Interface
in Ionic Polymer Transducers. PhD thesis, Virginia Polytechnic Institute and State
University, 2005.
[4] G. Alici, G. Spinks, N. N. Huynh, L. Sarmadi, and R. Minato. Establishment of a
biomimetic device based on tri-layer polymer actuatorslpropulsion fins. Bioinspira-
tion & Biomimetics, 2:818—830, 2007.
[5] P. V. Alvarado and K. Youcef-Toumi. Design of machines with compliant bodies for
biomimetic locomotion in liquid environments. Journal of Dynamic Systems, Mea-
surement, and Control, 128:3—13, 2006.
[6] Mart Anton, Andres Punning, Alvo Aabloo, Madis Listak, and Maarja Kruusmaa. To-
wards a biomimetic BAP robot. In Proc. of Towards the Autonomous Mobile Robots,
pages 1—7, 2004.
[7] G. A. Baker and P. Graves-Morris. Padé Approximants. Cambridge University Press,
New York, 1996.
[8] Y. Bar-Cohen. Electroactive polymers as artificial muscles - capabilities, potentials
and challenges. In Y. Osada, editor, Handbook on Biomimetics, pages 1-13 (Chapter
8). NTS Inc., 2000.
[9] Y. Bar-Cohen. Electroactive Polymers as Artificial Muscles - Capabilities, Potentials
and Challenges. Handbook on Biomimetics. NTS Inc., 2000.
[10] Y. Bar-Cohen. Electric flex. IEEE Spectrum, 41(6):29—33, 2004.
202
[11] M. D. Bennett and D. Leo. Ionic liquids as stable solvents for ionic polymer trans-
ducers. Sensors and Actuators A: Physical, 115:79—90, 2004.
[12] C. Bonomo, L. Fortuna, P. Giannone, S. Graziani, and S. Strazzeri. A model for ionic
polymer metal composites as sensors. Smart Materials and Structures, 15:749—758,
2006.
[13] C. Bonomo, L. Fortuna, P. Giannone, S. Graziani, and S. Strazzeri. A nonlinear model
for ionic polymer metal composites as actuators. Smart Materials and Structures,
16:1—12, 2007.
[14] C. Bonomo, C. Del Negro, and L. Fortuna S. Graziani. Characterization of IPMC strip
sensorial properties: Preliminary results. In Proceedings of the 2003 International
Symposium on Circuits and Systems, volume 4, pages IV—816— IV—819, 2003.
[15] Frdric Boyer, Mathieu Porez, and Wisama Khalil. Macro-continuous computed
torque algorithm for a three-dimensional eel-like robot. IEEE Transactions on
Robotics, 22(4):763-775, August 2006.
[16] Paola Brunetto, Luigi Fortuna, Salvatore Graziani, and Salvatore Strazzeri. A model
of ionic polymer-metal composite actuators in underwater operations. Smart Materi-
als and Structures, 17(2): 1—12, 2008.
[17] G. Del Bufalo, L. Placidi, and M. Porfiri. A mixture theory framework for modeling
the mechanical actuation of ionic polymer metal composites. Smart Materials and
Structures, 17:04501011—14, June 2008.
[18] Z. Chen, Y. Shen, N. Xi, and X. Tan. Integrated sensing for ionic polymer—metal com-
posite actuators using PVDF thin films. Smart Materials and Structures, 16(2):SZ62—
8271, 2007.
[19] Z. Chen, X. Tan, and M. Shahinpoor. Quasi-static positioning of ionic polymer-metal
composite (IPMC) actuators. In Proceedings of the IEEE/ASME International Con-
ference on Advanced Intelligent Mechatronics, pages 60—65, Monterey, CA, 2005.
[20] Z. Chen, X. Tan, A. Will, and C. Ziel. A dynamic model for Ionic Polymer-Metal
Composite sensors. Smart Materials and Structures, 16(4): 1477—1488, 2007.
[21] Zheng Chen and Xiaobo Tan. A control-oriented and physics-based model for ionic
polymer-metal composite actuators. IEEE/ASIWE Transactions on Mechatronics,
l3(5):519—529, 2008.
[22] C.K. Chung, P.K. Fung, Y.Z. Hong, M.S. Ju, C.C.K. Lin, and TC. Wu. A novel fabri-
cation of ionic polymer-metal composites (IPMC) actuator with silver nano-powders.
Sensors and Actuators B: Chemical, 117:367—375, November 2006.
[23] R. V. Churchill. Operational Mathematics. McGraw-Hill, New York, 3rd edition,
1972.
203
[24] R W. Clough and J. Penzien. Dynamics of Structures. McGraw-Hill, 1993.
[25] B. E. Conway. Electrochemical Supercapacitors Scientific Fundamentals and Tech-
nological Applications. Springer, 1999.
[26] H. Darcy. Les Fontaines Publiques de la Ville de Dijon. Dahnont, Paris, 1856.
[27] P. G. de Gennes, K. Okumura, M. Shahinpoor, and K]. Kim. Mechanoelectric effects
in ionic gels. Europhysics Letters, 50(4):513—518, 2000.
[28] Deniz Dogruer, Joonsoo Lee, Woosoon Yim, KwangJ Kim, and Doyeon Kim. Fluid
interaction of segmented ionic polymerCmetal composites under water. Smart Mate-
rials and Structures, 165220—8226, 2007.
[29] JJ. Dosch, D.J. Inman, and E. Garcia. A self-sensing piezoelectric actuator for col-
located control. Journal of Intelligent Material Systems and Structures, 3:166—185,
1992.
[30] Y. Fang, T. J. Pence, and X. Tan. Nonlinear elastic modeling of differential expansion
in trilayer conjugated polymer actuators. Smart Materials and Structures, 172065020,
2008.
[31] Y. Fang, X. Tan, Y. Shen, N. Xi, and G. Alici. A scalable model for trilayer con-
jugated polymer actuators and its experimental validation. Materials Science and
Engineering: C, 28(3):421—428, 2008.
[32] K. Farinholt and DJ. Leo. Modeling of electromechanical charge sensing in ionic
polymer transducers. Mechanics of Materials, 36:421—433, 2004.
[33] Kevin M. F arinholt. Modeling and Characterization of Ionic Polymer Transducers
for Sensing and Actuation. PhD thesis, Virginia Polytechnic Institute and State Uni-
versity, 2005.
[34] Guo-Hua Feng and Ri-Hong Chen. Improved cost-effective fabrication of arbitrar-
ily shaped uIPMC transducers. Journal of Micromechanics and Microengineering,
18(1):015016:1—9, January 2008.
[3 5] F. E. Fish and G. V. Lauder. Passive and active flow control by swimming fishes and
mammals. Annual Review of Fluid Mechanics, 38: 193—224, 2006.
[36] G. F. Franklin, J. D. Powell, and A. Emami-Naeini. Feedback Control of Dynamic
Systems. Pearson Education, Inc., Upper Saddle River, NJ, 5th edition, 2006.
[3 7] J. M. Gere and SP. Timoshenko. Mechanics of Materials. PWS Publishing Company,
Boston, MA, fourth edition, 1997.
[38] RE. Grimshaw, J.H. Nussbaum, and AJ. Grodzinsky. Kinetics of electrically and
chemically induced swelling in polyelectrolyte gels. Journal of Chemical Physics,
93:4462—4472, 1990.
204
[39] S. Guo, T. Fukuda, and K. Asaka. A new type of fish-like underwater microrobot.
IEEE/ASAE Transactions on Mechatronics, 8(1):]36—141, 2003.
[40] H. Hu, J. Liu, 1. Dukes, and G. Francis. Design of 3D swim patterns for autonomous
robotic fish. In Proceedings of the 2006 IEEE/RSJ International Conference on Intel-
ligent Robots and Systems, pages 2406—241 1, Beijing, China, 2006.
[41] J .-H. Jeon, I.-K. Oh, J .-H. Han, and S.-W. Lee. Development of bio-mimetic patterned
IPMC actuators with multiple electrodes. Key Engineering Materials, 334-335: 1005—
1008, 2007.
[42] Jin-Han Jeon and Il-Kwon Oh. Selective grth of platinum electrodes for MDOF
IPMC actuators. Thin Solid Films, 517:5288—5292, 2009.
[43] Jin-Han Jeon, Sung-Won Yeom, and Il-Kwon Oh. Fabrication and actuation of ionic
polymer metal composites patterned by combining electroplating with electroless
plating. Composites Part A: Applied Science and Manufacturing, 39:588—596, 2008.
[44] Norihiro Kamamichi, Masaki Yamakita, Kinji Asaka, and Zhi-Wei Luo. A snake-like
swimming robot using IPMC Actuatorl/ Sensor. In Proceedings of the IEEE Confer-
ence on Robotics and Automation, pages 1812—1817, Orlando, Florida, May 2006.
[45] R. Kanno, A. Kurata, S. Tadokoro, T. Takamori, , and K. Oguro. Characteristics and
modeling of ICPF actuator. In Proceedings of the Japan- USA Symposium on Flexible
Automation, pages 219—225, 1994.
[46] R. Kanno, S. Tadokoro, T. Takamori, and M. Hattori. Linear approximate dynamic
model of IPMC (ionic conducting polymer gel film) actuator. In Proceedings of the
IEEE International Conference on Robotics and Automation, pages 219—225, Min-
neapolis, MN, 1996.
[47] B. Kim, D. Kim, J. Jung, and J. Park. A biomimetic undulatory tadpole robot using
ionic polymer-metal composite actuators. Smart Materials and Structures, 14: 1579—
1585, 2005.
[48] B. Kim, J. Ryu, Y. Jeong, Y. Tak, B. Kim, and J.-O. Park. A ciliary based 8-legged
walking micro robot using cast IPMC actuators. In Proceedings of IEEE International
Conference on Robots and Automation, pages 2940—2945, 2003.
[49] K. J. Kim and M. Shahinpoor. Ionic polymer-metal composites: II. manufaturing
techniques. Smart Materials and Structures, 12:65—79, 2003.
[50] Kwang J. Kim and Mohsen Shinpoor. A novel method of manufacturing three-
dimensional ionic polymer-metal composites (IPMCs) biomimetic sensors, actuators,
and artificial muscles. Polymer, 43:797—802, November 2002.
[51] Seong Jun Kim, In Taek Lee, and Yong Hyup Kim. Performance enhancement
of IPMC actuator by plasma surface treatment. Smart Materials and Structures,
16(1):N6—N1 1, January 2007.
205
[52] G. V. Lauder, E. J. Anderson, J. Tangorra, and P. G. A. Madden. Fish biorobotics:
Kinematics and hydrodynamics of self-propulsion. Journal of Experimental Biology,
210:2767—2780, 2007.
[53] G. V. Lauder and E. G. Drucker. Morphology and experimental hydrodynamics of fish
fin control surfaces. IEEE Journal of Oceanic Engineering, 29(3):556—571, 2004.
[54] G. V. Lauder and P. G. A. Madden. Learning from fish: Kinematics and experimental
hydrodynamics for roboticists. International Journal of Automation and Computing,
4:325—335, 2006.
[55] K.-M. Lee, Z. Wei, and Z. Zhou. Modeling by numerical reduction of modes for mul-
tivariable control of an optical-fiber draw process. IEEE Transactions on Automation
Science and Engineering, 3:1 19—130, 2006.
[56] M. J. Lighthill. Aquatic animal propulsion of high hydromechanical efliciency. Jour-
nal of Fluid Mechanics, 44:265—301 , 1970.
[57] Pin Lu and K.H. Lee. An alternative derivation of dynamic admittance matrix of
piezoelectric cantilever bimorph. Journal of Sound and Vibration, 266:723—735,
2003.
[58] E. Mbemmo, Z. Chen, S. Shatara, and X. Tan. Modeling of biomimetic robotic fish
propelled by an ionic polymer-metal composite actuator. In Proceedings of the IEEE
International Conference on Robotics and Automation, pages 689—694, 2008.
[59] S. McGovern, G. l Alici, V. T. Truong, and G. Spinks. Finding NEMO (Novel Elec-
tromaterial Muscle Oscillator): A polypyrrole powered robotic fish with real-time
wireless speed and directional control. Smart Material and Structures, 18:09500921—
10, 2009.
[60] Measurement Specialties. Piezo Film Sensors Technical Manual. Hampton, VA,
1998.
[61] P. Meissner, B. Kogel Amd F. Riemenschneider, H. Halbritter, S. Jatta, M. Maute,
and M. C. Amann. Tunable long-wavelength VCSELs using a moveable mirror mem-
brane. In Proceedings of the 18th Annual Meeting of the IEEE Lasers and Electro-
Optics Society, pages 324—325, 2005.
[62] Kristi A. Morgansen, Benjamin 1. Triplett, and Daniel J. Klein. Geometric methods
for modeling and control of free-swimming fin-actuated underwater vehicles. IEEE
Transactions on Robotics, 23(6):]184—1199, December 2007.
[63] S. Nemat-Nasser. Micromechanics of actuation of ionic polymer-metal composites.
Journal of Applied Physics, 92(5):2899—2915, 2002.
[64] S. Nemat-Nasser and J. Li. Electromechanical response of ionic polymer-metal com-
posites. Journal of Applied Physics, 87(7):3321—333 1 , 2000.
206
[65] K. Newbury. Characterization, Modeling, and Control of Ionic Polymer Transducers.
PhD thesis, Virginia Polytechnic Institute and State University, Blacksburg, Virgina,
2002.
[66] K. M. Newbury and D. J. Leo. Electromechanical modeling and characterization
of ionic polymer benders. Journal of Intelligent Material Systems and Structures,
13:51-60, 2002.
[67] Kenneth M. Newbury and Donald J Leo. Linear electromechanical model for ionic
polymer transducers - part 1: model development. Journal of Intelligent Material
Systems and Structures, 14:333—342, 2003.
[68] K. Oguro, Y. Kawami, and H. Takenaka. Bending of an ion-conducting polymer film-
electrode composite by an electric stimulus at low voltage. Journal of Micromachine
Society, 5:27-30, 1992.
[69] J. J. Pak, J.Kim, S. W. Oh, J. H. Son, S. H. Cho, S.-K. Lee, J.-Y. Park, and B. Kim.
Fabrication of ionic-polymer-metal-composite (IPMC) micropump using a commer-
cial nafion. In Y. Bar-Cohen, editor, Smart Structures and Materials 2004 : Electroac-
tive Polymer Actuators and Devices (EAPAD), Proceeding of SPIE, pages 272—280,
2004.
[70] R. Pelrine, R. Kombluh, Q. Pei, and J .Joseph. High-speed electrically actuated elas-
tomers with strain greater than 100%. Science, 287(5454):836—839, February 2000.
[71] S. D. Peterson, M. Porfiri, and A. Rovardi. A particle image velocimetry study of
vibrating ionic polymer metal composites in aqueous environments. IEEE/ASME
Transactions on Mechatronics, 2009. In press.
[72] E. Popov. Mechanics of Materials. Prentice-Hall, Englewood Cliffs, NJ, 2 edition,
1976.
[73] Maurizio Porfiri. Charge dynamics in ionic polymer metal composites. Journal of
Applied Physics, 104(10): 104915, November 2008.
[74] A. Punning, M. Kruusmaa, and A. Aabloo. Surface resistance experiments with IPMC
sensors and actuators. Sensors and Actuators A, 2007. in press.
[75] R. C. Richardson, M. C. Levesley, M. D. Brown, J. A. Hawkes, K. Watterson, and
P. G. Walker. Control of ionic polymer metal composites. IEEE/ASAE Transactions
on Mechatronics, 8(2):245—253, 2003.
[76] John Elie Sader. Frequency response of cantilever beams immersed in viscous flu-
ids with applications to the atomic force micrscope. Journal of Applied Physics,
84(1):64—76, 1998.
207
[77] Takahisa Sakakibara, Hiroaki Izu, Tatsuya Kura, Wataru Shinohara, Hiroshi Iwata,
Seiichi Kiyama, and Shinya Tsuda. Development of high-voltage photovoltaic micro-
devices for an energy supply to micromachines. In Proceedings of IEEE 5th Inter-
national Symposium on Micro Machine and Human Science, pages 71-76, Nagoya,
Japan, 1994.
[78] M. Shahinpoor and K. J. Kim. The effect of surface-electrode resistance on the perfor-
mance of ionic polymer-metal composite (IPMC) artificial muscles. Smart Materials
and Structures, 9:543—55 1, 2000.
[79] M. Shahinpoor and K.J. Kim. Ionic polymer-metal composites: 1. Fundamentals.
Smart Materials and Structures, 10:819—833, 2001 .
[80] M. Shahinpoor and K.J. Kim. Ionic polymer-metal composites: IV. Industrial and
medical applications. Smart Materials and Structures, 14:197—214, 2005.
[81] Y. Shen, U. C. Wejinya, N. Xi, M. S. Buckley, C. A. Pomeroy, and D. N. Arhosti.
Force measurement and mechanical characterization of living Drosophila embryos
for human medecal study. Journal of Engineering in Medicine, 221(2):99—112, 2007.
[82] Y. Shen, N. Xi, U.C. Wejinya, and WJ. Li. High sensitivity 2-D force sensor for
assembly of surface MEMS devices. In Proceedings of IEEE/RSI International Con-
ference on Intelligent Robots and Systems, pages 3363—3368, Sendai, Japan, 2004.
[83] M. Siripong, S. Fredholm, Q.A. Nguyen, B. Shih, J. Itescu, and J. Stolk. A cost-
effective fabrication method for ionic polymer-metal composites. In V. Bharti, Y. Bar-
Cohen, Z.Y. Cheng, Q. Zhang, and J. Madden, editors, Electroresponsive Polymers
and Their Applications, volume 889 of MS Symposium Proceedings, pages 0889—
WO4—03, Warrendale, PA, 2006.
[84] E. Smela. Conjugated polymers for biomedical applications. Advanced Materials,
15(6):481—494, 2003.
[85] S. Tadokoro, S. Yamagarni, T. Takamori, , and K. Oguro. Modeling of nafion-pt com-
posite actuators (ICPF) by ionic motion. In Yoseph Bar-Cohen, editor, Smart Struc-
tures and Materials 2000: Electroactive Polymer Actuators and Devices (EAPAD),
Proceedings of SPIE, volume 92-102, pages 92—102, 2000.
[86] X. Tan, D. Kim, N. Usher, D. Laboy, J .Jackson, A.Kapetanovic, J. Rapai, B. Sabadus,
and X. Zhou. An autonomous robotic fish for mobile sensing. In Proceedings of the
IEEE/RSJ International Conference on Intelligent Robots and Systems, pages 5424—
5429, Beijing, China, 2006.
[87] J. Tangorra, P. Anquetil, T. F ofonofi‘, A. Chen, M. Del Zio, and I. Hunter. The applica-
tion of conducting polymers to a biorobotic fin propulsor. Bioinspiration & Biomimet-
ics, 2:S6—S 17, 2007.
208
[88] Michael S. Triantafyllou and George S. Triantafyllou. An efficient swimming ma-
chine. Scientific American, 272:64, 1995.
[89] S. Tung, B. Maines, F. Jiang, and T. Tsao. Transonic flow separation control by
MEMS sensors and actuators. In Proceedings of 3 9th AIAA Aerospace Sciences Meet-
ing & Exhibit, Reno, NV, 2001. AIAA Paper 2001-0123.
[90] Steve Tung, Scott R Witherspoon, Larry A Roe, Al Silano, David P Maynard, and
Ned Ferraro. A MEMS-based flexible sensor and actuator system for space inflatable
structures. Smart Materials and Structures, 10(6):]230—1239, December 2001.
[91] E. Voltera and E. C. Zachmanoglou. Dynamics of Vibrations. Charles E. Merrill
Books, Inc., 1965.
[92] Y. Xiao and Bhattacharya. Modeling electromechanical properties of ionic polymers.
In Smart Structures and Materials 2001: Electroactive Polymer Actuators and De-
vices (EPAD), Proceedings of SPIE, volume 4329, pages 292—300, 2001.
[93] Sung-Weon Yeom and Il-Kwon Oh. A biomimetic jellyfish robot based on ionic
polymer metal composite actuators. Smart Materials and Structures, 1820850021—
10, June 2009.
[94] W. Y11‘Il, J. Lee, and K. J. Kim. An artificial muscle actuator for biomimetic underwa-
ter propulsors. Bioinspiration & Biomimetics, 2:S31—S4l, 2007.
[95] Junzhi Yu, Long Wang, and Min Tan. Geometric optimization of relative link lengths
for biomimetic robotic fish. IEEE Transactions on Robotics, 23(2):382—386, April
2007.
[96] Side Zhao and Junku Yuh. Experimental study on advanced underwater robot control.
IEEE Transactions on Robotics, 21(4):695—703, August 2005.
[97] J. W. L. Zhou, H.-Y. Chan AndT. K. H. To, K. W. C. Lai, and W. J. Li. Polymer
MEMS actuators for underwater micromanipulation. IEEE/ASME Transactions on
Mechatronics, 9(2):334—342, 2004.
[98] Kemin Zhou. Essentials of Robust Control. Prentice Hall, Upper Saddle River, New
Jersey, 1998.
209
IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII
111111111111111111111
3
629
PVDF {LA‘> R,
-Q(S)I - R
1 R39
———l::l——
! _
l
r
1
Figure 6.2: Design of the charge amplifier.
Basically, the charge Q(s) generated by the PVDF is proportional to the bending dis-
placement Z (s) of the beam [82]:
Q(s) = GZ(S)7 (6-2)
where the constant G depends on the transverse piezoelectric coefficient d3] , the geometry
of the composite beam, and the Young’s moduli of individual layers. By combining (6.1)
and (6.2), one can obtain the transfer function from Z (s) to V0 (s). A laser displacement
sensor (OADM 2016441/Sl4F, Baumer Electric) is used for both calibration of the PVDF
119
sensor and validation of the sensing approach. In order to test the charge amplifier cir-
cuit, the IPMC/PVDF beam with one end fixed is tapped and then the laser sensor is used
to detect the damped vibration of the beam. The measured vibration frequency is 42 Hz,
which is much higher than the cutofl’ frequency of the charge amplifier. Fig. 6.3(a) shows
the charge output of PVDF corresponding to the damped vibration, and Fig. 6.3(b) demon-
strates that the charge signal is almost linear with respect to the bending displacement.
These experimental results have validated the performance of the charge amplifier circuit.
6.1.2 Impact of the stiffening effect and design of the insulator thick-
BESS
The additional PVDF and insulating layers make the composite beam stiffer than the IPMC
layer itself. It is of interest to understand the impact of this stiffening effect on the bending
performance since this will be useful for the optimal IPMC/PVDF structure design. The
investigation is conducted by combining analytical modeling, finite element computation,
and experiments. Design optimization here is concentrated on the thickness of the insulat-
ing layer, but the approach can be used for the design of other parameters, such as the type
of material for the insulating layer and the dimensions for IPMC and PVDF.
Fig. 6.4 illustrates the schematic of the IPMC/PVDF structure and the used notation in
the following discussion. The beam stiffness can be characterized by its spring constant
K = —, (6.3)
zmax
where F is a quasi-static transverse force applied at the free end of the cantilever beam
and zmax is the corresponding displacement at the acting point. The spring constant can be
calculated analytically using composite beam theory [37]. In Fig. 6.4, the position of the
120
0.016 3
0.015 * i
0.014 '
Charge (uC)
O
3
O)
0.012 r
0.01 1 *
0.01 ‘ ‘ r
0 0.05 0.1 0.15
Time (s)
(a)
0.016
0.015 -
0.014 *
Charge (uC)
O
S
c»
0.012 ~
0.011 * ‘
0.01
5 5.5 6
Displacement (mm)
(b)
Figure 6.3: (a) Charge output of the PVDF corresponding to the damped vibration; (b)
charge output versus the bending displacement.
121
mechanical neutral axis of the composite beam is given by:
2,11 Elm-Ci
ho = .
2:, EH.-
(6.4)
Here E1, E2 and E3 are the Young’s moduli of IPMC, insulating layer, and PVDF, respec-
tively. H1, H2 and H3 are the thicknesses of those layers. C1, C2 and C3 are the positions of
the central axes of the layers, which can be calculated as:
C1%H1/2,C2 2H1 +H2/2,C3 =H1+H2+H3/2. (6.5)
The distance between the central axis and the neutral axis can be written as:
di=ICi-ho|, for i-—-l,2,3. (6.6)
The moment of inertia of each layer is:
I,-=éWIL-3+WH,-d,2, for i=1,2,3. (6.7)
From the moment balance equation [37],
where p(x) is the radius of curvature. For small bending, the radius of curvature can be
given by:
1 dzz
36‘; = a}? (6-9)
where Z(x) denotes the deflection of the beam along the length x. With the boundary con-
122
dition 2(0) = 0 and 2(0) = 0, one gets
F sz x3
—— —-——). (6.10)
2?:1Ei1i( 2
200 = 6
Evaluating 2 at x = L yields the expression of spring constant
F 321 15.1,-
K = __ = ___,_1 , 6.11
w
e 1
Neutral hO IPMC H1
aXIS .................................................... .- l +
Figure 6.4: Bending of IPMC/PVDF composite beam.
Experiments are conducted to measure and compare the spring constants of the IPMC
and IPMC/PVDF beams. The IPMC or IPMC/PVDF beam is clamped at one end and is
pushed by the tip of a calibrated micro-force sensor which is mounted on a linear actuator.
The sensitivity of the micro-force sensor is 9.09 mV/uN :l: 6.5% and its spring constant
is 0.264 N/m. A laser displacement sensor measures the bending displacement of the
beam 2m under the pushing force F. A 20X microscope (FS60, Mitutoyo) is used to
monitor the experiments. Fig. 6.5(a) illustrates the diagram of the experimental setup, while
123
Fig. 6.5(b) shows the actual picture. Measurements are conducted for an IPMC beam and
two IPMC/PVDF beams which have insulating layers in different thickness (IPMC/PVDF]
and IPMC/PVDFZ). Detailed beam dimensions can be found in Table 6.1. Fig. 6.6 shows
the measured displacement versus force data and the linear approximations, from which
the spring constants can be calculated. From the experimental data, the Young’s moduli
of individual layers can be identified using (6.11): E1 = 0.571 GPa, E2 = 0.73 GPa, E3 =
1.96 GPa. These values are within the ranges reported in the literature [83, 77].
Table 6.1: Dimension and spring constant of different beams.
Beams IPMC IPMC/PVDF] IPMC/PVDF2
W(mm) 7.3 8.2 7.6
L (m) 37.2 36.0 33.0
H1 (pm) 355 340 350.0
H2 (pm) 30.0 100.0
H3 (11111) 30.0 30.0
Km (N/m) 0.906 2.283 4.676
KFEA(N/m) 0.908 2.286 4.647
To validate the linear analysis above, more accurate finite-element computation is con-
ducted using CoventorWare, where the identified parameters are used together with the
given geometric dimemsions. The spring constants are calculated based on the free-end
deflection of beams when they are subjected to an external force F = 20 pN at the tip.
Table 6.1 lists the spring constants obtained through experimental measurement (Kmea) and
finite element analysis (KFEA), for the different beams. The close agreement between Kmea
and KFEA validates the model and analysis.
As shown in Table 6.1, the thicker the insulator, the stiffer the IPMC/PVDF structure. In
order to optimize the bending performance of the IPMC/PVDF structure, one should select
the elastic insulating layer as soft and thin as possible. However, thinner insulating layer
may result in stronger electrical feedthrough coupling. In our design, the thickness of the
insulating layer is chosen to be 30 um to achieve tradeoff between the two considerations.
124
(— Linear actuator
Clamp Micro Force
(— Sensor
IPMC/PVDF
(— Laser Sensor {7
Computer _) |:l
=
Figure 6.5: (a) Experimental setup for spring constant measurement; (b) picture of the
experimental setup.
125
* Experimental data
()3 Linear approximation *
‘5 025* IPMC/PVDF1
3 0.2 IPMC
«’5 4
..s \
0.15 .
i
0.1- fl 1,
_ it a x
°-°5‘ * IPMC/PVDFZ
I I . . ,
0 200 400 600 800
Force (uN)
Figure 6.6: Spring constant of IPMC/PVDF beams.
6.1.3 Electrical feedthrough coupling and model-based real-time com-
pensation
The feedthrough coupling effect
Since the PVDF film is closely bonded to the IPMC with a very thin insulating PVC
film, the coupling capacitance between the IPMC and the PVDF results in the electrical
feedthrough effect during simultaneous actuation and sensing. When the actuation signal
is applied to the IPMC, the actuation voltage generates coupling current going through the
insulating layer and then induces coupling charge on the PVDF. As a result, the charge
amplifier gathers both the sensing and coupling charges from the PVDF. The presence
of feedthrough coupling is illustrated by applying a 0.4 Hz square-wave actuation input
(peak-to-peak 1.4 V). In the experiment, the humidity is 34% and the temperature is 23 °C.
Fig. 6.7(a) shows the bending displacement detected by the laser sensor, while Fig. 6.7(b)
126
shows the output from the charge amplifier. The spikes in the PVDF sensor output arise
from the capacitive coupling between the IPMC and PVDF layers when the actuation volt-
age jumps.
Modeling of the coupling effect
A complete circuit model of the IPMC/PVDF structure is developed to understand and
capture the feedthrough coupling dynamics. As shown in Fig. 6.8, the model includes the
equivalent circuits for individual layers and their natural couplings. Due to the nonneg-
ligible resistances resulted from the porous surface electrodes of the IPMC, the voltage
potential is not uniform along the IPMC length. A distributed transmission-line type model
is thus proposed. The overall circuit model is broken into discrete elements along its length
for parameter identification and simulation purposes. In this work, the circuit model is
chosen to have four sections of identical elements. The surface resistance of IPMC is rep-
resented by RS] , while other key electrodynamic processes (e.g., ionic transport, polymer
polarization, and internal resistances) are reflected in the shunt element consisting of resis—
tor Rcl and capacitor Cpl. The polymer resistance is described by Rpl- In the circuit model
of the insulating layer, R p2, sz, R62 are resistances and capacitances between the IPMC
and PVDF. In the circuit model of the the PVDF, R53 is the surface resistance of PVDF and
R p3, Cp3 represent the resistance and capacitance between the electrodes of the PVDF.
In order to identify the circuit parameters, the impedances are measured at multiple fre-
quencies. The impedances of each layer are nonlinear fiinctions of the resistances and ca-
pacitances involved. The parameters are identified using the Matlab command nl inear f it,
which estimates the coeflicients of a nonlinear function using least squares. Table 6.2 lists
the identified parameters.
The proposed circuit model will be validated by comparing its prediction of the feedthrough
coupling signal with experimental measurement. We first explain a simple method for mea-
suring the coupling signal. We observe that, due to the low surface resistance of PVDF (see
127
0.08
P .° P
o o o
c N h c»
Displacement (mm)
I
P
o
N
-0.04 -
-0.06 ‘ *
Time (s)
(a)
spike due
to coupling '
0.3L s '. . ,
r / ‘. .~
PVDF sensor output (V)
O
0 2 4 6
T lme (s)
(b)
Figure 6.7 : (a) Bending displacement detected by the laser sensor; (b) sensing output from
the PVDF, showing the spikes fiom electrical feedthrough.
128
i Levelt Level 4
'(3) +1: ;_.:
+ R81
R1
]p
H
H
4F—
] IPNK:
Fist “ICp1
J: .2
vmn E RcL
J—l
- pd-—---H----—d---
O
O
r 0
vii
‘ -----— -----------
L:
r
I
‘4
r1“.
1
4
’i,
r
'1."
3",; ——E —|i
l
: . .i
§Rp2 [1E 092: a Insulating layer
_i.... _
.117
To charge 5 RS3
amplifier
“‘+—L1
§Rp3 PVDF
. :45“ . . 4 47$"; 4 5: . I“ «L2
3 R33 : 9
Figure 6.8: Circuit model of the IPMC/PVDF structure.
Table 6.2: Identified parameters in the circuit model.
IPMC layer Insulating Layer PVDF layer
R51 17 Q sz 500 M!) RS3 0.1 Q
RC1 30 Q sz 42 pF Rp3 600 M9
Cpl 3 mF R02 4.5 M!) C13 290 pF
RP] 25 KS2
129
Table 6.2), the electrode layer L1 in Fig. 6.8 shields the coupling current from reaching the
electrode layer L2. This means that the feedthrough coupling signal does not exist in V _,
which is related to the charge fiom the layer L2. This statement is supported by the mea-
surement, shown in Fig. 6.9(a), where spikes only appear in Vp+. Since only Vp+ has the
coupling component while the sensing components in Vp- and Vp+ have a phase shifl of
180°, the coupling signal is obtained as:
Vc=Vp++V-. (6.12)
Fig. 6.9(b) shows the extracted coupling signal.
Fig. 6.10 compares the Pspice simulation results based on the circuit model with exper-
imental results when a 1 Hz square-wave actuation voltage is applied. Good agreement is
achieved for both the actuation current in IPMC (Fig. 6.10(a)) and the coupling voltage Vc
(Fig. 6.10(b)).
The transfer function from the actuation voltage to the coupling voltage can be derived
fi'om the circuit model. Since there are 14 capacitors in the circuit model, the transfer
firnction will be 14th-order, which is not easy to implement in real time. After an order-
reduction process, the transfer function of the coupling dynamics can be approximated by
a 5th-order system:
T _ —(509s4+72s3+ 1.5 x104sz+2203s)
C _ s5 +9525s4 +1.5 x10483 +2.9 x10582 +4.5 x105s+6 x104.
(6.13)
To further verify the coupling model, a sequence of sinusoidal voltage signals with
frequency ranging from 0.01 Hz to 20 Hz are applied to the IPMC. Actuation voltages are
measured and coupling signals are effectively extracted from Vp+ and Vp— for the purpose
of obtaining the empirical Bode plots of coupling dynamics. Fig. 6.11 shows that the Bode
plots of the derived transfer function (6.13) match up well with the measured Bode plots.
130
0'2 Spike due *1“
" ‘ ~to coupling
Time (s)
(a)
0.08
0.06 :
0.04 -
0.02 .
ch
-0.02 i
-0.04 -
-0.06 ' ‘
0 2 4 6
Time (s)
(b)
Figure 6.9: (a) Vp+, Vp- sensing signals; (b) extracted coupling signal Vc.
131
— Experimental data
- - - Simulation data
2'."
E .
E
2
I-
3
O
C
.2
iii
3
‘6
< I
0 1 2 3 4
Time(s)
(a)
0.06 7 -— Experimental data
5 - - - Simulation data
0.04. ........ .......... ..... ..
E ' .
5’ r
5 l
3 .
> 01 _______ ...
U) I
s .
Q .
0 l
-01“. ......... - .......... . ........
l _ t ' :
I q I
-0.06 #
0 1 2 3 4
Time(s)
(b)
Figure 6.10: Comparison of model prediction and experimental measurement. (a) Actua-
tion current; (b) coupling signal.
132
‘20 t7: 7 7* * *7 ‘‘‘‘‘ ff?
6 f”. ,3;
3 :25:
82-40’. ...__. . .::I:::I . Iiilifll j
2 5f . :572225 +EXPerimentaldatai iii???
; g ; ';i;i§i; ---Simulationdata
_6O -2 L 1111.1;1-1 L 111111110 1 IZZZTITI1 I I :1211112 2 . . .1. 3
10 1O 10 10 10 10
300 7 7”??? 7 777???? . if???
a . :..‘::.L. I i. i: I ..:.:..I
3 250-
8 9
2 200’
0- :j
150 . ....1 “.1 ....l ...i ....
10'2 10‘1 10° 101 102 103
w(rad/s)
Figure 6.11: Verification of the coupling dynamics.
Real-time compensation in simultaneous actuation and sensing
There are several possible schemes to get rid of the coupling signal. Inserting another
conductive layer between the IPMC and PVDF to shield the feedthrough coupling is one
potential solution, but at the cost of increased stiffness and fabrication complexity. Another
solution is to just use Vp— as the sensing signal, but this single-mode sensing scheme is
sensitive to the common-mode noise in practice. Since the coupling dynamics has high-
pass characteristics, one might also try to eliminate the coupling component with low-
pass filtering. However, the relatively low cut-off frequency of the coupling dynamics,
comparing to the actuation bandwidth (See Fig. 6.12), makes this approach infeasible.
In this reseach, a model-based real-time compensation scheme is proposed to remove
the feedthrough coupling component. The coupling charge is calculated from the coupling
circuit model (6.13). By subtracting it from the measured charge of the PVDF, the sensing
charge can be extracted. Fig. 6.13 illustrates the compensation scheme. Fig. 6.14 compares
the displacement measurement obtained from the PVDF sensor with that from the laser
133
Mag (dB)
<1)
01
l
.h
01
—+—Bode plotofcoupling dynamics 3
j i 1.3.};3- 0 -Bode plotofactuation dynamics g g 1.33;;
_55 L..;;;;;i ;;;;;;;;, 2;;;;;;;; ;;;;;;;;; '
10'2 10‘ 100 101 10
I
01
O
A ALLA
(0 (rad/s)
Figure 6.12: Bode plots of coupling dynamics and actuation dynamics.
sensor when a 0.4 Hz square-wave actuation input is applied. It is seen that the spike related
to the electrical coupling is removed by the compensation scheme. Although there is about
12% error shown in Fig. 6.14, the amplitudes and the phases agree well. Investigation is
under way to further improve the measurement accuracy of the PVDF sensor.
6.1.4 Application to micro-injection of living drosophila embryos
The developed IPMC/PVDF sensori-actuator is applied to the micro-injection of living
Drosophila embryos. Such operations are important in embryonic research for genetic
modification. Currently this process is implemented manually, which is time-consuming
and has low success rate due to the lack of accurate control on the injection force, the
position, and the trajectory. The IPMC/PVDF structure is envisioned to provide accurate
force and position control in the micro-injection of living embryos, and thus to automate
this process with a high yield rate. In this research, an open-loop injection experiment with
134
. IPMC/PVDF
structure
Actuation signal
? i ,9 sensing slggpl
- ' I sensing + + E + -
Feedthrough , i
Figure 6.13: Diagram of the real-time compensation.
0.08
9 9
Displacement (mm)
-0.06 :
-0.08
O
I I
p p
c c
-h N
coupling dynamics :
5*; Feedthrough :
5 coupling model g, _____ Coupling
I : prediction
----------------------------
—— Measured by laser sensor
- - - Measured by PVDF
1‘”
I
I I
I I
l
\. 3
L
4
Time (s)
Figure 6.14: Comparison of displacement measurements by laser sensor and PVDF sensor.
135
the IPMC/PVDF sensori-actuator is conducted, and the process of the injection behavior is
captured by the PVDF sensor.
The developed IPMC/PVDF micro-force injector is illustrated in Fig. 6.15. A mi-
cropipette with an ultra-sharp tip (1.685 pm in diameter and 2.650 in angle), is mounted
at the end point of a rigid tip attached to the IPMC/PVDF structure. The Drosophila
IPMC/PVDF
Mounting l E
Connector Z
\
'\ Micro
/ ii" Force
Micropipette
Ineedle X
Figure 6.15: Illustration of the IPMC/PVDF micro-force injector.
embryos are prepared as described in [81]. The dimensions of the embryos are variable
with an average length of 500 um and a diameter of about 180 um. Fig. 6.16(a) shows
the diagram of the experimental setup for embryo injection, while Fig. 6.16(b) shows the
photo. A 3-D precision probe station (CAP-945, Signatone), which is controlled by a 3-D
joystick, moves the needle close to an embryo and then a ramp voltage, which starts from 0
V and saturates at 2 V, is applied to the IPMC. The IPMC drives the beam with the needle
to approach the embryo. After the needle gets in contact with the membrane of the embryo,
136
Microscope
Lens
L—A Micropipette/n
- Micropipette eedle (20X
Halocarbon 700 011 i Ineedle image)
Adhesive tape
' 3-D
Glass Slrde 3-D
Probe _
\ Station Joystick
Micro-Force
Sensor
(8)
Laser sensor
(b)
Figure 6.16: (a) Diagram of experimental setup for embryo injection; (b) picture of exper-
imental setup.
137
the latter will be deformed but not penetrated due to its elasticity. At this stage, the needle
is still moving until the reaction force between the needle and embryo reaches the pene-
tration force. The needle stops at the penetration moment for a while (about 0.2 ms) due
to temporary force balance. After that, the embryo membrane is penetrated and the needle
moves freely into the embryo.
Fig. 6.l7(a) shows the snapshots of the successful injection progress. Fig. 6.l7(b)
shows both the displacement of the needle detected by the laser sensor and by the com-
pensated PVDF sensing signal. It is concluded that the predicted displacement reflects the
movement of the needle, and the process of injection can be monitored by the PVDF.
138
0.12 - -
-— Measured by laser sensor
- - - measured by PVDF , V
0.1 i z
E 0.08- , -
T: I
C I
0 I
E 0 06 , .
O I
o I
4g .
.2 0'04 - a- “ 5 ’ 1
n I ‘ ’
a I
0.02 - ,’ Injection -
r “ f ’
"o 6
Time (s)
(b)
Figure 6.17: (a) Snap shots captured during the embryo injection; (b) bending displacement
during the injection measured by both the laser sensor and the integrated PVDF sensor.
139
6.2 IPMC/PVDF Differential-Mode Sensory Actuator and
Its Validation in Feedback Control
6.2.1 Design of IPMC/PVDF differential-mode sensory actuator
Integrated differential-mode sensor for bending output
Fig. 6.18 illustrates the design of the integrated differential mode bending sensor for an
IPMC actuator. Two complementary PVDF films, placed in opposite poling directions, are
bonded to both sides of an IPMC with insulating layers in between. In our experiments, we
have used 30 um thick PVDF film from Measurement Specialties Inc., and 200 pm thick
IPMC fiom Environmental Robots Inc. The IPMC uses non-water-based solvent and thus
operates consistently in air, without the need for hydration. Scrapbooking tape (double-
sided adhesive tape,70 um thick) fi'om 3M Scotch Ltd. is used for both insulating and
bonding purposes. A picture of a prototype is shown at the bottom of Fig. 6.18. Since we
are focused on demonstrating the proof of the concept in this work, the materials used are
chosen mainly based on convenience. However, the models to be presented later will allow
one to optimize the geometry design and material choice based on applications at hand.
The differential charge amplifier, shown in Fig. 6.19, is used to measure the PVDF
sensor output. In particular, the inner sides of two PVDF films are connected to the common
ground, while the outer sides are fed to the amplifiers. Let Q1(s) and Q2(s) be the charges
generated on the upper PVDF and the lower PVDF, respectively, represented in the Laplace
domain. The signals Vp+ and Vp‘ in Fig. 6.19 are related to the charges by
R15
— —mQ2(S),
iPMC VDF
Eiectrode—+ Insulating
layers
PVDF
Figure 6.18: Design of the IPMC/PVDF composite structure for sensing of bending output
(force sensor not shown).
and the sensor output V0 equals
_ R1R3S
Vo(S) - m(Q1 (S) — Q2(S))- (6-14)
Let the bending-induced charge be Q(s) for the upper PVDF, and the common noise-
induced charge be Q,,(s). If the sensor response is symmetric under compression ver-
sus tension (more discussion on this in Section 6.2.2), one has Q1(s) = Q(s) + Qn(s),
Q2(s) = —Q(s) + Qn(s), which implies
2R1R3S
Vo(S) = mg”, (6-15)
and the effect of common noises (such as thermal drift and electromagnetic interference)
is eliminated fiom the output. The charge amplifier (6.15) is a high-pass filter. To accom-
modate the actuation bandwidth of IPMC (typically below 10 Hz), the R1 and C] values
in the circuit are properly chosen so that the cutoff frequency of the charge amplifier is
sufficiently low. By picking R1 = 5000 MD, C1 = 1350 pF and R2 = R3 = 10 k9, a cutoff
141
frequency of 0.023 Hz is achieved.
:1 R
Q1 C —i.'_}:l——
PVDF1 I ]P°""9
direction
PVDF—$133231... mam
-Q2 R1 R39
IF _
91
Figure 6.19: Differential charge amplifier for PVDF sensor.
A model is developed for predicting the sensitivity of the bending sensor in terms of the
design geometry and material properties. Refer to Fig. 6.20 for the definition of geometric
variables. Suppose that the IPMC/PVDF beam has a small uniform bending curvature with
tip displacement 21; without external force, the force sensor beam attached at the end of
IPMC/PVDF appears straight with tip displacement 22. One would like to compute the
_Q_
22 ’
effector displacement 22. With the assumption of small bending for IPMC/PVDF beam,
sensitivity where Q represents charges generated in one PVDF layer given the end-
the curvature can be approximated by [19]
1 z — (6.16)
p
where p represents the radius of curvature. As H3 << 0.5H1 +H2, we assume the stress
inside the PVDF to be uniform and approximate it by the value at the center line of this
layer:
0.5H H 0.5H
032E3£=E3 ‘+ p” 3 (6.17)
where E3 is the Young’s modulus of the PVDF. The electric displacement on the surface of
142
PVDF is
Ds=d3103. (6.18)
where d3] is the transverse piezoelectric coefficient. The total charge generated on the
PVDF is then
Q=/DSdS=DsL1W1. (6.19)
With (6.16), (6.17), (6.18) and (6.19), one can get
_ 2d31E3W1(0.5H1+H2 +0.5H3)Z1
6.20
Q L] ( )
The end-effector displacement 22 is related to 21 by
22:21+Lzsin arctan(a) zz1(l+& , (6.21)
L1 L1
Combining (6.20) and (6.21), one can get the sensitivity
S: Q = 2d31E3W1(0.5H1+H2+0.5H3). (6.22)
22 L1 '1' 2L2
Table 6.3 lists the parameters measured or identified for our prototype. The sensitivity
is predicted to be 1830 pC/mm, while the actual sensitivity is characterized to be 1910
pC/mm using a laser distance sensor (OADM 2016441/Sl4F, Baumer Electric). With the
charge amplifier incorporated, the sensitivity 2:3 at frequencies of a few Hz or higher is
measured to be 2.75 V / mm, compared to a theoretical value of 2.71 V/mm.
Force Sensor for End-effector
The structure of the force sensor is similar to that of IPMC/PVDF sensory actuator. As
illustrated in Fig. 6.2], two PVDF films are bonded to the both sides of a relatively rigid
beam. In our experiments, we have used 200 pm thick Polyester from Bryce Corp. for the
143
Table 6.3: Parameters identified for the IPMC/PVDF sensory actuator prototype (including
force sensor).
W1 L1 H1 H2 H3
10 mm 40 mm 200 um 65 um 30 um
W2 L2 h] h2 113
6 mm 30 mm 200 um 65 um 30 um
Er Ez E3 6131
5 GPa 0.4 GPa 2 GPa 28 pC/N
Net-Lital axis “ 91PMC ’ “ a"?
O Y
., -e.-. “-
,-., ' .
insulating layer
L1
Z ' IPMC/PVDF
PVDF force sensor
-------------.
Figure 6.20: Geometric definitions of IPMC/PVDF sensory actuator.
144
beam. An end-effector, e.g., a glass needle in microinjection applications, is bonded the
tip of the force sensor. An external force experienced by the end—effector will cause the
composite beam to bend, which produces charges on the PVDF films. Another differential
charge amplifier as in Fig. 6.19 is used for the force sensor. The whole force-sensing beam
is attached to the front end of the IPMC/PVDF beam.
P
Bottom View
Figure 6.21: Design of the force sensor for the end-effector.
Refer to Fig. 6.22. The sensitivity model for force sensing, QFI, is provided below.
Here Q f represents the charges generated in one PVDF in response to the force F exerted
by the end-effector. The beam curvature can be written as
1 F L —
_ = #1 (6.23)
P (x) 24:0 EiIi
where p(x) denotes the radius of curvature at x, E], E2, E3 are the Young’s moduli of the
Polyester film, the bonding layer, and PVDF respectively. 11, 12 and 13 are the moments of
inertia for those layers, which are given by
1 3
11 — l—z‘thla
1
I2 = -6-W2hg+
l
13 = 6W2h§+
W2h2(h1+h2)2
2
,
W2h3(h1 + 2h2 + h3)2
145
The stress generated in the PVDF is approximately
h] + 2112 + h3
0307) = E383 (x) = E3 (6-24)
210 (x)
With (6.18), (6.23) and (6.24), one can get the electric displacement in PVDF,
h] +2112 +h3 F(L2 -x)
D3 (3‘) = 6131030?) = E36131 - (6-25)
2 23:05:]:
The total charge generated in the PVDF can be written as
L2 d E W L2 11 2h h
QfZ/ D3(x)W2dx= 3‘ 3 2 2(3 1+ 2+ 3)F (6.26)
0 42,-:oE11i
Then the sensitivity of the force sensor is
d E W L2 h 2h h
szgiz 313 2 2( 1+ 2+ 3). (6.27)
F 42:05:11
Relevant parameters for the force sensor in our prototype can be found in Table 6.3.
Theoretical value of S f is computed to be 0.456 pC/IIN, which is close to the actual value
0.459 pC/IIN from measurement. With the charge amplifier circuit, the sensitivity of the
overall force sensor @- at high fiequencies (several Hz and above) is characterized to be
0.68 mV/uN, compared to the model prediction of 0.67 mV/IIN.
The integrated IPMC/PVDF sensory actuator and the charge sensing circuits are placed
in conductive plastic enclosures (Hammond Manufacturing) to shield electromagnetic in-
terference (EMI) and reduce air disturbance and thermal drift. A slit is created on the
side of the shielding box enclosing IPMC/PVDF so that the end-effector protrudes out for
manipulation purposes. Fig. 6.23 shows the picture of the overall system.
146
Figure 6.22: Geometric definitions of the PVDF sensor.
.,_Jii .: __ Conductive boxes“
. Sensing circuit ’ ‘
l for position sensor
Q 4 —. 4. "1—~:.~——
Sensing circuit
for force sensor ; - ,i
,
\ .
i .‘ :.;"" ’ VA ‘-
Isl; 48.»:\.-~
my: :.$ ,- 7.. ‘
Figure 6.23: IPMC/PVDF sensory actuator and sensing circuits in shielding enclosures.
147
6.2.2 Experimental verification of sensor robustness
In this section we experimentally verify the robustness of the proposed sensory actuator
with respect to the following undesirable factors: 1) feedthrough of actuation signal, 2)
thermal drift and other environmental noises, and 3) asymmetric PVDF sensing responses
during compression versus tension. The discussion will be focused on the PVDF sensor
for IPMC bending output, since the problems associated with the PVDF force sensor are
similar and actually simpler (no need to worry about actuation feedthrough).
Feedthrough Coupling
Close proximity between IPMC and PVDF results in capacitive coupling between the two.
Fig. 6.24 illustrates the distributed circuit model for the composite IPMC/PVDF beam.
Suppose an actuation signal V,-(s) is applied to IPMC. If one connects both sides of a single
PVDF film to a differential charge amplifier, as done typically in Section 6.1, the output will
pick up a signal that is induced by the actuation signal via electrical coupling. Fig. 6.7 in
Section 6.1 illustrates the traditional feedthrough problem. While one can attempt to model
the feedthrough coupling and cancel it through feedforward compensation, the complexity
of such algorithms and the varying behavior of coupling make this approach unappealing
to real applications.
In the new charge sensing scheme proposed in this design, the inner sides of the two
PVDF sensors are connected to a common ground (see Fig. 6.19). Since the surface elec-
trode resistances of PVDF films are very low (< 0.152), the inner layers L2 and L3 in
Fig. 6.24 will effectively play a shielding role and eliminate the feedthrough coupling sig-
nals. This analysis is verified experimentally, where a square-wave actuation voltage with
amplitude 2 V and frequency 0.1 Hz is applied to the IPMC. Fig. 6.25(a) shows that the
charge amplifier output V0 contains no feedthrough-induced spikes. The definitions for V0,
V}, and Vp‘ in the figure can be found in Fig. 6.19. Furthermore, the bending displacement
148
PVDF1
Insulating
layer
To charge
amplifier
—_l-_
4)
IPMC
k ‘
insulating
layer
PVDF2
Section 1 Section n
Figure 6.24: Distributed circuit model of IPMC/PVDF beam.
149
obtained from the PVDF output V0 correlates well with the actual bending displacement
measured by the laser distance sensor, as shown in Fig. 6.25(b). Note that the PVDF output
V0 is related to the bending displacement 21 through the charge amplifier dynamics (6.15)
and the proportional relationship (6.20). Since (6.15) represents a high-pass filter, at rela-
tively high frequencies (determined by the cut-off fiequency), the correlation between V0
and 21 can be approximated by a constant; however, at lower fiequencies (including the
step input, in particular), the dynamics (6.15) has to be accommodated to obtain the dis-
placement trajectory from the raw PVDF signal V0. The latter has been adopted throughout
the chapter, whether the inverse of the charge amplifier dynamics is implemented digitally
to retrieve 21.
Thermal Drift and Environmental Noises
PVDF sensors are very sensitive to ambient temperatures and electromagnetic noises. Such
environmental noises could significantly limit the use of PVDF bending/force sensors, es-
pecially when the operation frequency is low (comparing with the fluctuation of ambient
conditions). Refer to Fig. 6.19. Let noise-induced charges be Q"l and Q"2 for PVDF1
and PVDF 2, respectively. Suppose that no actuation signal is applied, and thus bending-
induced charge Q(s) = 0. The voltage signals can then be expressed as
+ _ R15
_ Rls
Vp (S) —an2(S), (6-29)
_ R1R3S
V00) — R20 +R1Cls)(Q..(s) —Qn.(s)). (6.30)
Inside a conductive shielding enclosure, thermal and EMI conditions are relatively steady
and uniform. This implies Q"I (s) z Qn2(s) and the influence of environmental noises on
the sensor output V0 is negligible.
Two experiments have been conducted to confirm the above analysis. In order to isolate
150
—— V
o
3 t - - . Vp+
- - - V
2 p-
2 1 ’ .1
a II 5 I '
g o -’_ "'I‘ is, ,’ u ,s ’I " I”
g 1 r \i ’ ‘ I, : ‘ I!" ‘1 ’
‘ i \ .I I " I 1
’r\ ‘1‘" 'I \ ‘I‘: I, K \‘ i
-2 4 l ‘. ‘ \ I ‘. \ ‘ ‘ , \i
N ' ‘ \\ I ‘. \ '
\ I s i ‘ '
..3 i- \' ~ ‘ I 4
0 5 10 15 20 25 30
Time (s)
(a)
Measured by laser sensor
; - - - Predicted by PVDF sensor
Bending displacement (mm)
0
_2 l
20 5 10 15 20 25
E . -
E , ,
8 ;
t
m -2 1 1 1 1
0 5 10 15 20 25
Time (s)
(b)
Figure 6.25: Experimental results showing elimination of feedthrough signal. (a) Raw
PVDF sensing signals under a square-wave actuation input (2V, 0.1 Hz); (b) Comparison
between the bending displacements obtained from the PVDF sensor and from the laser
sensor.
151
the effect of noises, no actuation signal is applied. In the first experiment, the IPMC/PVDF
beam was exposed to ambient air flows and electromagnetic noises. Fig. 6.26(a) shows the
complementary sensing outputs Vp+ and VP“ and the resulting charge amplifier output V0.
From (6.28)-(6.30), the discrepancy between Vp+ and VP“ indicates that the noise-induced
charges Q”l and Q"2 on the two PVDF layers can be significantly different, leading to
relatively large sensing noise in V0. In contrast, Fig. 6.26(b) shows the results from the sec-
ond experiment, where the IPMC/PVDF sensory actuator was placed inside the conductive
shielding enclosure. In this case, while V; and Vp‘ could still vary over time individually,
their trajectories are highly correlated and close to each other. Consequently, V0 remained
under 1 mV, compared to about 20 mV in the first case. These experiments have confirmed
that the proposed differential sensing scheme, together with the shielding enclosure, can
effectively minimize the effect of thermal drift and other common noises.
Asymmetric Sensing Response during Extension versus Compression
Because of its compliant nature, a single PVDF film does not produce symmetric charge
responses when it is under tension versus compression. In particular, it is diflicult to ef-
fectively introduce compressive normal stress into the flexible film. As a result, the charge
response of a PVDF layer under extension can faithfully capture the beam motion while
the response under compression cannot. This is illustrated by experimental results shown
in Fig. 6.27(a): each of the sensing signals Vp+ and Vp‘ fiom the two PVDF layers is asym-
metric under a symmetric, sinusoidal actuation input. With the differential configuration
of two PVDF films, however, the asymmetric responses of individual PVDF films combine
to form a symmetric output V0, as seen in Fig. 6.27(a). This is because when one film is
in compression, the other is in tension. Fig. 6.27(b) shows that the bending displacement
obtained based on the PVDF signal V0 agrees well with the laser sensor measurement. We
have further examined the performance of the proposed integrated sensing scheme under
other types of actuation inputs, including the step inputs. From Fig. 6.28, the bending
152
Voltage (V)
0 5 10 15 20 25 30
Time (s)
(a)
0.5
0 4 ‘5 .f’c . _
—.q’.:'\"‘ |Pi ~..~_T_f:.~~...
"ti/II 1i “ " " " ~ ::=’s‘5‘~‘r
E 0.3 ~ ’ 1 t .J
o i
a:
i -—Vo
> 0.2 ! - - -V J
p+
- -.Vp_
0.1 ‘
WW
0 L 1 J
0 5 10 15 20 25 30
Time (s)
(b)
Figure 6.26: (a) Sensing noise when IPMC/PVDF placed in open field; (b) sensing noise when
IPMC/PVDF placed inside conductive shielding enclosure.
153
trajectory under a step input (2V) can be captured well by the PVDF sensor.
Another advantage of adopting two complementary PVDF films is that it alleviates
the effect of internal stresses at bonding interfaces. When bonding a single PVDF to
IPMC, mismatch of internal stresses at the PVDF/IPMC interface could lead to delami-
nation and/or spontaneous creep of the composite beam. While this problem could be less—
ened by using appropriate bonding technologies, it was found that the proposed scheme
can effectively maintain the structural stability of the composite beam, without stringent
requirements on bonding.
6.2.3 Feedback control based on the integrated sensor
The practical utility of the proposed IPMC/PVDF sensory actuator has been demonstrated
in feedback control experiments. Trajectory tracking experiments are first performed,
where no tip interaction force is introduced. Simultaneous trajectory tracking and force
measurement are then conducted to examine both integrated bending and force sensors.
Feedback control of bending displacement
Fig. 6.29 illustrates the closed-loop system for the control of IPMC bending displacement.
Here P(s) represents the actuation dynamics for the IPMC/PVDF composite structure, H (s)
is the bending sensor dynamics, K (s) is the controller, r is the reference input, u is the ac-
tuation voltage, and 22 is the bending displacement of the end-effector. In experiments data
acquisition and control calculation are performed by a dSPACE system (DSl 104, dSPACE
Inc.); for real applications such tasks can be easily processed by embedded processors, e. g.,
rnicrocontrollers. A laser sensor is used as an external, independent observer for verifica-
tion purposes.
In general K (s) can be designed based on a nominal model of the plant P(s) and various
objectives and constraints. An example of H00 control design can be found in [20], where a
physics-based, control-oriented model is also developed for IPMC actuators. Since IPMC
154
“V
0
--IV
2 13+ _
---V
p—
2 1"
a)
a)
g 0»
o
> ¢\ ¢' ’a¢‘\ " _-\
I 3 .\ 'I. I“, " ' —
I - '
-1r I i v" I .' " \ ‘
I I \ ‘ ' .1
\ 1 ' \ I . I /\ 1.“
I I l ‘ ’ \ ‘
\ \ , ‘ , \ l -I \ I \ -\
’ ' ' 1 I ‘3. ' \
’ \. I \ \ I \ -
*2‘ 'I \ " \l \, I! \
v v ‘I V
0 5 10 15 20
Time(s)
(a)
Measured by laser sensor 4
- - - Predicted by PVDF sensor
0.5 ................ ................ . .............................
Bending displacement (mm)
C
-0.5
E 05. ............. ................................. .1
..E, 3 -j
8 3 i
LU : g
0 5 10 15 20
Time (s)
(b)
Figure 6.27: Self-compensation of asymmetric tension/compression sensing response. (a)
Raw PVDF sensing signals under a sinusoidal actuation signal (0.2 Hz, amplitude l V); (b)
Comparison between the bending displacements obtained fi'om the PVDF sensor and from
the laser sensor.
155
A 1 r
E
:7 0.8~
r:
o
E 0.6 p . 1
§ 04 p - .. ~ Measuredbylasersensor 1‘
'8'. ' -— Predicted by PVDF sensor
:6 ‘ : .
or 0.2
.E
.2 0 ......
0 i 1 1 1 '1
m 0 10 20 30 40 50 60
A 1 I V 1 I I
E
E .
r 0,,— -
8
0 10 20 30 40 50 60
Time (s)
Figure 6.28: Comparison between the bending displacements obtained from the PVDF
sensor and fi'om the laser sensor, when a 2V step input is applied.
Bending
r + K(s) u P(S) displacemen:
3 ' z2
PVDF
sensing signal
H (s)
H "‘(s) 4, ,
IPMC/PVDF” ------------------------ =
Figure 6.29: Closed-loop system for control of IPMC bending displacement.
156
modeling and control design are not the focus of this chapter, we have identified the plant
model P(s) empirically and used a simple proportional-integral (PI) controller for K (s) to
validate the integrated sensing scheme. In particular, the empirical frequency response of
the IPMC/PVDF sensory actuator has been obtained by applying a sequence of sinusoidal
actuation inputs (amplitude 0.2 V, frequency 0.01 Hz to 10 Hz) and measuring the corre-
sponding bending response. It has been found that the measured dynamic behavior could
be approximated by a second order system, the parameters of which have been further
determined using the Matlab command “fitsys”. The resulting P(s) is
P(s) _ 2.73 +20
_ 1000(82 + 33.4S + 18.9) '
The sensing model is obtained from (6.15) and (6.20):
181505
H =———.
(S) 6.57s+1
The following reference trajectory is used: r(t) = sin(0.37tt) mm. Based on the mod-
els and the reference, a PI controller K (s) = 1000 (40 + ?) is designed to achieve good
tracking performance while meeting the constraint |u| < 2 V. Fig. 6.30(a) shows the ex-
perimental results of tracking the bending reference. It can be seen that the PVDF sensor
output tracks the reference well; firrthermore, the actual bending displacement, as observed
by the laser sensor, has close agreement with the PVDF output. The actuation voltage u,
shown in Fig. 6.30(b), falls within the limit {—2, 2] V.
Feedback bending control with simultaneous force measurement
It is desirable in many applications to have both displacement and force feedback. With the
proposed IPMC/PVDF sensory actuator, one can perform feedback control of the displace-
ment while monitoring the force output, as well as perform feedback control of the force
157
—- Displacement from laser sensor
,4 1.5- '- - -- Displacement from PVDF sensor
E - - - Reference
E 1 -
0
i
g 0.5 ~
Q
.2
'0
.E .
“g .
(an) -0.5
-1 . ’
0 5 10 15 20
Time (s)
(a)
2
1.5 -
1 _
E 0.5 -
O
> -0.5~
-1 .
-1.5
..2 1 1 .
0 5 10 15 20
Time (s)
(b)
Figure 6.30: Experimental results on feedback control of bending displacement using inte-
grated PVDF sensor. (a) Bending displacement; (b) actuation voltage.
158
output while monitoring the displacement. In the following experiment we will demon-
strate the feedback bending control with simultaneous force measurement.
To mimic the force level often encountered in bio and micromanipulation applications,
we have attached a sharp glass needle as an end-effector at the tip of force-sensing beam
and used it to pierce soap bubbles. Fig. 6.3 1(a) shows the experimental setup. A number of
bubble-penetrating experiments were conducted to get an estimate of the rupture force by
moving a bubble manually toward the needle until it breaks, when no actuation voltage was
applied. Fig. 6.31(b) shows the force sensor response during a typical run. It can be seen
that the response first rises from zero to a peak value, and then starts decayed oscillations.
Since the PVDF sensor measures essentially the bending of the passive beam, its output
can be interpreted as an interaction force only when the end-effector is in contact with a
foreign object. Thus for the response in Fig. 6.3l(b), only the first rising segment truly
represents the force, after which the membrane ruptures and the beam starts oscillating.
Hence we take the peak value of such responses as the penetration force. Fig. 6.32 shows
the penetration force measured in 26 independent experiments. Overall the measurements
are consistent with an average of 11 MN. The variation is believed due to the randomly
created bubbles that might have different thicknesses. Note that for many real applications,
such as microinjection of embryos or cells [18], the end-effector will maintain contact with
the object under manipulation, in which case the output of PVDF force sensor would truly
represent the interaction force at all times.
A feedback bending control experiment with force monitoring has been conducted,
where the reference for the end-effector displacement r(t) = 0.2 sin(0.47rt) mm. During
the experiment, the end-effector penetrated two soap bubbles at t = 9.32 and t = 15.72
seconds, respectively. Fig. 6.33(a) shows the estimated end-effector displacement based on
the integrated PVDF bending sensor (sandwiching IPMC), under the assumption that the
force-sensing PVDF beam is not deflected. The estimated displacement trajectory follows
closely the reference, with slight perturbations at the moments when penetrations occur,
159
Penetration force
10-
’z‘
3 5- A .
E U m- . V A A A A A. A A.
a - , vvvvvvv
_5_ True force ‘
-10-
_.,= 1 ‘ Beam oscillations _
"’0 0.5 1 1.5 2 2.5 3
Time(s)
(b)
Figure 6.31: Measurement of the micro force in piercing soap bubbles. (a) Experimental
setup; (b) PVDF sensor response during and afier penetration.
160
.15 ' T ' ' m
10"
Penetration force of bubble p. N)
5 r .
o 1 1 1 1 1
0 5 10 15 20 25 30
Samples
Figure 6.32: Measured forces during penetration of soap bubble membranes.
indicating that the feedback control was in effect. Fig. 6.33(b) shows the output of the
integrated force sensor, where the two penetrations were captured clearly. Note that, as
explained earlier and illustrated in Fig. 6.31(b), only the first rising segment of the trajectory
during each penetration truly represents the interaction force, while the remaining portion
of the signal arises from oscillations following penetration. The control output (actuation
voltage) is shown in Fig. 6.34, where one can see that feedback is in action to suppress the
disturbance caused by penetration.
Note that the end-effector displacement 22 predicted by the PVDF bending sensor alone
(Fig. 6.33(a)) does not capture the true displacement d when the end-effector interacts with
objects. To obtain the true displacement, one can combine the bending sensor output 22 and
the force sensor output F:
d = 22 +F/k, (6.31)
where k is the stiffness of the force-sensing beam. For our prototype, k = 0.067 N/m.
Fig. 6.35 compares the end-effector displacement obtained from (6.31) and that observed
by the laser sensor, which shows that indeed the end-effector position can be monitored by
161
I; l
- - - Reference
03 . '- - -- Measured by PVDF sensor
0.2 -
0.1"
Bending displacement (mm)
O 5 10 15 20
Time (s)
(a)
15
10"
Tip force ()1 N)
O
-10..
_15 1 1 1
0 5 10 15 20
Time (s)
(b)
Figure 6.33: Experimental results on bending feedback control with tip force measurement.
(a) Displacement of the end-effector estimated based on the integrated PVDF bending sen-
sor alone; (b) PVDF force sensor output.
162
0.5
Control output (V)
O
0 5 10 15 20
Time (s)
Figure 6.34: Actuation voltage generated by the feedback controller.
combining the integrated bending and force sensors.
6.3 Chapter Summary
In this chapter, a novel scheme was proposed for implementing integrated sensors for an
IPMC actuator, to achieve sensing of both the bending displacement output and the force
output. In the first design, an IPMC is bonded with a PVDF sensing film in a single-
mode sensing configuration. The stiffening effect and the electrical feedthrough coupling
are investigated. In the second design, two thin PVDF films are bonded to both sides of
an IPMC beam to measure the bending output, while a passive beam sandwiched by two
PVDF films is attached at the end of IPMC actuator to measure the force experienced by
the end-effector. The differential configuration adopted in both sensors has proven critical
in eliminating feedthrough coupling, rejecting sensing noises induced by thermal drift and
EMI, compensating asymmetric tension/compression responses, and maintaining structural
stability of the composite beams. For the first time, feedback control of IPMC has been
163
0.3 ’
0.2 * ~ -' , l-
0.1 i l l
-0.1 ~
-0.3r
-0.4 -
-0.5 ‘ ‘ '
0 5 10 15 20
Time (s)
End-effect displacement (mm)
— Laser sensor
- - - Integrated sensors"
Figure 6.35: Estimation of true end-effector displacement by combining the integrated
bending and force sensors, and its comparison with the laser sensor measurement.
successfully demonstrated using only integrated sensors, showing that one can simulta-
neously regulating/tracking the bending displacement and monitoring the force output (or
vice versa).
164
Chapter 7
Monolithic Fabrication of Ionic Polymer
Metal Composite Actuators Capable of
Complex Deformations
This chapter is organized as follows. The fabrication process for MDOF IPMC actuators is
presented in Section 7.1. In Section 7.2 we investigate the change of stiffness and swella-
bility of Nafion films caused by the ion-exchange process. The performance of fabricated
artificial pectoral fins is characterized in Section 7.3. Finally, concluding remarks are pro-
vided in Section 7.4.
7.1 FabricationProcess
In this section, we outline the overall fabrication flow first and then discuss the individual
steps in more details. Fabrication of an artificial pectoral fin is taken as an example. As
illustrated in Fig. 7.1, the major process steps include:
o (a): Create an aluminum mask on Nafion with e-beam deposition, which covers the
intended IPMC regions;
165
o (b): Etch with argon and oxygen plasmas to thin down the passive regions;
0 (c): Remove the aluminum mask and place the sample in platinum salt solution to
perform ion-exchange. This will stiffen the sample and make the following steps
feasible;
0 ((1): Pattern with photoresist (PR), where the targeted IPMC regions are exposed
while the passive regions are protected;
0 (e): Perform the second ion-exchange and reduction to form platinum electrodes in
active regions. To further improve the conductivity of the electrodes, 100 nm gold is
sputtered on the sample surface;
0 (t): Remove PR and lift off the gold on the passive areas. Soften the passive regions
with HCL treatment (to undo the effect of step (c));
o (g): Cut the sample into a desired shape.
7.1.1 Aluminum mask deposition
Since plasma will be used to selectively thin down the passive areas of the Nafion film,
the first step is to make an aluminum mask to protect the active areas from being etched.
Two shadow masks made of transparency films are used to cover both sides of the Nafion
film, in such a way that the passive areas are covered and the active areas are exposed. The
sample is then put into an e-beam deposition system, where aluminum can be deposited
at room temperature. Aluminum fihns of 200 nm thick are deposited on both sides of the
Nafion film. When the transparency masks are removed, the aluminum masks stay on the
active regions and the sample is ready for plasma etching.
166
Nafion Transparency mask
9 1
¢ o 4
E-beam aluminum deposition
Cross-section view Top view
(a) Deposit aluminum mask on both sides of Nafion film;
Plasma etching
:21...) an m
(c) Remove aluminum mask and perform
(b) Thin down passive area
ion—exchange to make Nafion stiffer;
with plasma etch;
((1) Deposit PR and then pattern (e) Perform another ion-exchange and
PR through lithography; electroless plating of platinum to create
IPMC electrodes; 2
Top view
(0 Remove PR and perform (g) Cut the patterned IPMC into a
final treatment; fin shape.
Cl N afion - Transparency Aluminum
Photoresist (PR) - Platinum - Gold
Figure 7.1: The process flow for monolithic fabrication of an MDOF IPMC actuator.
167
7.1.2 Plasma etching
Plasma treatment has been used for roughening the Nafion surface to increase the capaci-
tance of IPMC [51]. In this research, we use plasma to thin down the passive areas in the
MDOF IPMC actuator. The plasma etching system used in this research is Plasmaquest
Model 357. Experiments have been conducted to study the etching rates with different
recipes of gas sources. The etching rates with different recipes are shown in Table 7.1.
Table 7.1: Plasma etching with different recipes.
No. Ar 02 RF power Microwave power Etching rate
R1 20 sccm 30 sccrn 70 W 300 W 0.28 um/ min
R2 0 seem 50 sccm 70 W 300 W 0.22 um/ min
R3 50 sccm 30 sccm 70 W 300 W 0.51 um/ min
It can be seen that the combination of oxygen and argon plasmas can achieve a higher
etching rate than using the oxygen plasma alone. The oxygen plasma performs chemical
etching, which can oxidize the polymer and break Nafion into small molecules. The higher
oxygen flow rate, the higher etching rate but also the higher temperature on the Nafion film.
We have found that too strong oxygen plasma will damage the film because of overheating.
The argon plasma performs physical etching with high-speed, heavy argon ions. It can
remove the small molecules created by the oxygen plasma and roughen the Nafion surface,
which creates larger contact area for the oxygen plasma and thus accelerates the chemical
etching. The combination of oxygen and argon plasmas can achieve a high etching rate
with a low resulting temperature, which is critical to maintaining the original properties of
Nafion. In this research, the recipe R3 is adopted.
After several hours of plasma etching, the passive areas are thinned down to the desired
value. Then the sample is boiled in 2 N hydrochloride acid solution at 90 °C for 30 minutes
to remove the aluminum mask and to remove impurities and ions in Nafion. After that, the
membrane is further boiled in deionized (DI) water for 30 minutes to remove acid. Fig. 7.2
shows the Scanning Electron Microscope (SEM) picture of a selectively etched Nafion
168
sample, where the thinnest region is 48 um (down from the original 225 pm). The etched
sidewall is almost vertical (the angle is 88°). To roughen the active areas, the sample is
treated with plasma for 5 minutes. This roughening process can enlarge the metal-polymer
contact areas, which enhances the actuation performance of IPMC [51].
300 Olim
I
15.0kV15.3mm x180 SEiMl7/15f20091314 SOOum
Figure 7.2: SEM picture of a plasma-etched Nafion film.
7.1.3 Stiffening treatment
It is difficult to perform lithography on a pure Nafion film because Nafion swells in the de-
veloper, which usually contains water and organic solvent, and the swelling force can eas-
ily destroy the photoresist pattern. To address this challenge, we perform an ion-exchange
process to impregnate Nafion with large platinum complex ions (Pt(NH3)i+). Usually,
this ion-exchange process is used to absorb platinum complex ions for electroless platinum
plating [78]. However, we have discovered that, after the ion-exchange, the Nafion film
becomes stiffer and virtually non-swellable in water or acetone, which makes lithography
and the subsequent patterned electroless plating possible.
169
To perform ion-exchange, 25 ml of aqueous solution of tetraammine-platinum chloride
[Pt(NH3)4]C12 (2 mg Pt/ml) is prepared with 1 ml of ammonium hydroxide solution (5%)
to neutralize. The sample is immersed into the solution at room temperature for one day.
The formula of reaction is
[Pt(NH3 )4]2+ + 20H” + 2H+ (Nafion) 2°—°‘§ [Pt(NH3 )4]2+ (Nafion) + 2H20.
The impact of stiffening treatment will be studied in detail in Section 7.2.
7.1.4 Patterning of Nafion surface
AZ 9260 positive photoresist is selected to create thick PR patterns to protect the passive
areas of the actuator from electroless platinum plating. We spin-coat PR at 1000 rpm to
get 17 pm thick film and then bake it in oven at 90 °C for 2 hours. Since the Nafion film
alone is not rigid enough for spin-coating and it is easy to deform when baked in oven, an
aluminum frame is used to support and fix the Nafion film (25 mm by 25 mm), as shown
in Fig. 7.3. UV light with power density of 20 mw/cm2 is used to expose the sample with
the pattern mask for 105 seconds. After the sample is exposed, the PR is developed with
AZ 400K developer.
7.1.5 Electroless platinum plating and gold sputtering
The electroless platinum plating process is used to create thick platinum electrodes on the
active areas, which results in strong bonding between the metal and the polymer. Three
steps are taken for this process. First, another ion-exchange is performed to absorb more
platinum complex ions. Second, the sample is put into a bath with DI water at 40 °C.
Third, we add 1 ml of sodium boronhydride solution (5 wt% NaBH4) into the bath every
10 minutes and raise the bath temperature up to 60 °C gradually. After 30 minutes of
reduction, about 10 pm thick platinum electrodes will grow on the surface of the active
170
Sprew Aluminpm frame
o 0‘.
Q Nafion film 0
(25 mm by 25 mm)
Top view
I I l :1 1 Picture
Cross section view
Figure 7.3: Nafion film fixed by an almninum frame.
areas. To further improve the electrode conductivity, 100 nm thick gold is sputtered on
both sides of the sample. The surface resistance can be reduced by half with this gold-
sputtering step. Since PR patterns are still on the surface, the passive areas are protected.
When the PR is removed with acetone, the gold on the passive areas will be lified off.
7.1.6 Final membrane treatment
After electroless plating, the Pt complex ions in active regions are reduced to platinum
metal. But the Pt complex ions in the passive regions are still there, which makes the
passive regions stiff. To facilitate 3-D deformation, we need to undo the eflect of step (c)
to replace the Pt complex ions with H+. This can be achieved by simply boiling the sample
in 2N hydrochloride (HCL) acid [50]. The formula of reaction is
[Pt(NH3)4]2+(Nafion) + 2H+ 1&0 [Pt(NH3)4]2+ + 2H+(Nafion).
171
After the sample becomes flexible, it is put into sodium or lithium ion solution (1 N)
for one day to exchange H+ with Na+ or Li+ ions, to enhance actuation of IPMCs. A
fabricated IPMC membrane is shown in Fig. 7.4.
7.2 Impact of Stiffening Treatment
Stiffening treatment is an important step in fabrication of MDOF IPMC actuators. In this
section, we study the impact of the ion-exchange process on the stiffness and swellability
of Nafion films, and its implication in lithography and the overall fabrication process.
7.2.1 Stiffness change
The experimental setup for the stiffness measurement is shown in Fig. 7.5. The film is
fixed at one end by a clamp. A load cell (GSO-10, Transducer Techniques) is mounted
on a moving stage (U-SVRB-4, Olympus), which can be manipulated by hand to generate
smooth horizontal motion. When the stage is moved, the load cell pushes the Nafion film
to bend, and the restoring force at the tip of the film is measured. A laser displacement
sensor (OADM 2016441/Sl4F, Baumer Electric) is used to measure the corresponding tip
displacement. The resolution of the laser sensor is 20 um and the resolution of the load
cell is 0.05 mN. The setup is placed on an anti-vibration table (LW3048B-OPT, Newport).
A dSPACE system (DS1104, dSPACE) is used for data acquisition. The spring constant of
the Nafion film is calculated as
k:—
d,
where F and d are the tip force and the tip displacement, respectively.
We have measured the stiffness of Nafion—l 17 (183 um) and Nafion-l 110 (240 um)
before and after the ion-exchange process. The ion-exchanged Nafion fihns are dried in
air before the experiments. Fig. 7.6 shows the force and displacement data for each case,
172
Platinum electrode
10 >1 Ill
Passive area
Active area
300 0pm
l‘fi—F—r—l—‘lfi—lfi—W—‘I
15.0kV15.3mm x180 SE(M) 7/1 512009 13:34 300um
(e)
Figure 7.4: Patterned IPMC membrane. (a) Top view picture; (b) planar dimensions of the
membrane; (c) SEM picture of the cross section.
173
Di.
Laser sensor '
Moving direction
m *“r
Horizon indicator
Load cell
Moving stage
Anti-vibration table
Figure 7.5: Experimental setup for measuring the stifliiess of a Nafion fihn.
together with the results using linear fitting. The spring constant of the cantilever film is
_YWfl
_ 4L3’
where Y, W, L, and T are the Young’s modulus, width, length, and thickness of the Nafion
film, respectively. Based on the measured spring constant and dimension parameters, we
can calculate the Young’s modulus. Table 7.2 shows the spring constant, dimensions, and
Young’s moduli of the Nafion films. It can be seen that the stiffness of an ion-exchanged
Nafion film increases by about 2-3 times, compared to that of pure Nafion.
Table 7.2: Spring constant, dimensions, and Young’s moduli of the Nafion films in stiffness
testing. The ion-exchanged films are identified with an asterisk.
k W L T Y
Nafion—II7 1.7N/m 26mm 20mm 183 pm 355 Mpa
Nafion—117* 6.1N/m 26mm 19mm 183 pm 1.05 Gpa
Nqfion —N1110 4.5 N/m 25 mm 20 mm 254 pm 354 Mpa
Nafion—~N1110* 10.7 N/m 25 mm 21 mm 254 pm 966 Mpa
174
100 . . i i . .
90 - * Experimental data for Naflon-N1110' a
80 _ ' ' " Simulation data for Nafion-N1110' _
D Experimental data for Naflon-Nf110 ’ .
70 - - - - Simulation data for Nafion-N1110 , ’ -
2 , ’
E 60 - t ’ -
v ’ I
g 50 b ’ I I ‘1
.9 ’i‘
E 40 ’ ’ ’ a; "
30 ,. I i a ‘_ a “a
I ' an
I ’ a
20 i- ’ ’ :- * ' I’ ’ ‘. -l
I rr
10’- r’it’ .. '0‘ flélwca ‘
2’2“ ‘ '
or 1 J 1 1 1 1
0 1 2 3 4 5 6 7
Tip displacement (mm)
(a)
60 r L ,
* Experimental data for Naflon-N117' ‘
50 _ " ' ' Simulation data for Naflon-N117' I I ’ _
:1 Experimental data for Nation-11 7 4’
- - - Simulation data for Naflon-N117 I ’
A 40 - I -
Z I
e ,«’
g 30 * fi’ ’* "i
l- 20 . 1" I a
I
* I ’ ’ a
I a
10 _ I ’ a .. er I T T U 4
I Q— I
’ 'FW 9 f]
at i ..U- " ’3'
w "Lr ? ’ l l l l l l l
0 1 2 3 4 5 6 7 8 9
Tip displacement (mm)
(b)
Figure 7.6: Results of stiffness measurement. (a) Nafion—Nl l 10; (b) Nafion-l 17.
175
7.2.2 Swelling capacity change
We have further investigated the swelling capacity of Nafion before and after the ion-
exchange step. Four samples as listed in Table 7.3 have been tested. Measurements are
taken in the following steps. First, the surface areas of the samples in the dry condition
are measured. Second, the samples are immersed in water for 5 minutes and then taken
out for surface area measurement. Third, the samples are dried with paper towel and in air
before being immersed in acetone. After 5 minutes, the samples are taken out of acetone
and their surface areas are measured again. Table 7.3 shows the percentages of surface area
change comparing to the original size, for the four samples, after being soaked in water and
in acetone, respectively.
Table 7.3: Surface area changes of Nafion films in water and acetone. The ion-exchanged
films are identified with an asterisk.
With water With acetone
Nafion — 11 7 +19.1% +59.4%
Nafion — I 1 7* +07% +08%
Nafion —N1110 +20.1% +55.9%
Nafion—N1110* +1.2% +1.6%
It can be seen that while pure Nafion can expand by about 20% and 60% when soaked in
water and in acetone, respectively, ion-exchanged Nafion experiences only 1-2% expansion
under the same conditions. In other words, ion-exchanged Nafion has very low swellability
in solvents, which is important in lithography based patterning of the film. One possible
reason for the significantly reduced swelling capacity is that the film becomes stiffer and
the swelling force is unable to enlarge the volumn of the film. But the precise explanation
of the phenomenon requires further study.
176
7.2.3 Impact of ion-exchange on lithography
The impact of ion-exchange process on lithography has been investigated. First, we per—
formed lithography on pure Nafion. When the sample was put into the developer, the fihn
swelled and the PR patterns were destroyed by the swelling force, as shown in Fig. 7.7(a).
Then we performed lithography on ion-exchanged Nafion. The patterning result was sharp,
as shown in Fig. 7.7(b). It thus has demonstrated that the lithography of Nafion can be
dramatically improved by the stiffening treatment using ion-exchange.
7.3 Characterization of Fabricated Artificial Fins
7.3.1 Characterization method
We have characterized the performance of fabricated MDOF IPMC actuators on producing
sophisticated shape change. To capture the deformation, one may use multiple laser sen-
sors to detect the bending displacement of active areas [43]. However, when the actuator
generates large, complex deformation, some laser sensors can lose measurements. This
approach is also expensive. Another approach is to use a CCD camera to capture the video
of the actuator movement and then use image processing to extract the actuator movement
[17].
The experimental setup we use to characterize MDOF IPMC actuators is shown in
Fig. 7 .8. The fabricated IPMC membrane is cut into a pectoral fin shape and its base is
fixed by a multi-electrode clamp. The dimensions of the pectoral fin is shown in Fig. 7.8.
To minimize the contact resistance, gold foils (0.1 mm thick) are used to make the contact
electrodes. A CCD camera (Grasshopper, Point Grey Research) is oriented toward the
edge of the fin. Sinusoid voltage inputs of the same amplitude and frequency but different
phases are generated by the dSPACE system and applied to individual IPMC regions. The
tip bending displacements of active areas are detected using the image edge detector (Vision
177
- 'rmi
(b)
Figure 7.7: Lithography results. (a) With pure Nafion; (b) with ion-exchanged Nafion.
178
Assistant 8.5, National Instruments).
40mm:
'<-—>n
dSpace
Computer
NI Vision
Assistant 8.5
h ---------------------
Figure 7.8: Experimental setup for characterizing MDOF IPMC actuators.
7.3.2 Demonstration of twisting and cupping
Actuation experiments are first conducted in air. The phase differences between the actu-
ation signals applied to the three IPMC regions could be arbitrary. Since the goal here is
to demonstrate sophisticated deformation and not to optimize the control input, we have
restricted ourselves to the following particular class of phase patterns: the top IPMC leads
the middle IPMC in phase by 4), while the bottom IPMC lags the middle IPMC in phase by
4). For such a phase pattern, we call it phase «1: in short.
With all three IPMCs receiving inputs of the same phase (i.e., 0° phase), the fin gen-
erates bending. As this is not surprising (a single IPMC produces bending), we will not
present the detailed results on bending here. An example of twisting is shown in Fig. 7.9,
where the artificial fin has the same Nafion thickness, 85 pm, in the active and passive
areas. The voltages applied are sinusoidal signals with amplitude 3.0 V, frequency 0.3 Hz,
179
and phase 90°. The actuator clearly demonstrates a twisting motion. In order to quan-
tify the twisting deformation, we define the twisting angle 0, which is formed by the line
connecting the tips of top and bottom IPMCs with the vertical line m-n, as illustrated in
Fig. 7.10(a). Note that the tip displacements d1, d2, d3 of the IPMC regions are extracted
from the video. Fig. 7.10(b) shows the time trajectories of the displacements and the corre-
sponding twisting angle. The twisting angle achieved is 16° peak-to-peak, showing promise
in robotic fish applications.
We have also verified the actuator’s capability to generate the cupping motion. Fig. 7.1 1(a)
illustrates the definition of the cupping angle a, formed by the two lines connecting the tip
of the middle IPMC to the tips of the top and bottom IPMCs. Fig. 7.11(b) shows the tra-
jectory of the cupping angle for the same sample mentioned above, where the actuation
voltages have amplitude 3.0 V, frequency 0.3 Hz, and phase 180°.
7.3.3 Impact of the thickness in passive and active areas
To study the effects of thicknesses in active areas and passive areas on the actuation perfor-
mance, we have fabricated 5 samples with different thicknesses in active areas and passive
areas. Table 7.4 shows the thicknesses of the MDOF IPMC actuators. All actuators have
the same planar dimensions as specify in Fig. 7.4(b). To compare the actuation perfor-
mance of actuators with different thicknesses in the active areas, we have fabricated three
samples (SI, SZ, S5), where each sample has the same thickness in its active and passive
areas. They are fabricated from Nafion-l 110, Nafion—l 17, and Nafion-1135, respectively.
To study the effects of different thickness in passive areas, we have fabricated three samples
(S2, S3, S4) with the same thickness in the active areas (170 pm) but different thicknesses
in the passive areas.
The twisting angles generated by different samples with 0.3 Hz, 3 V and 90° phase volt-
age signals are shown in Fig. 7.12. From Fig. 7.12(a), for samples with uniform thickness,
the thinner the sample is, the larger the deformation. This can be explained by that, under
180
Figure 7.9: Snapshots of an actuated MDOF IPMC actuator, demonstrating the twisting
motion.
181
E
E
‘ ........
F/p1 .3
o
,
9 x’ 0 Y
1"]. l g
“d2 p2 X 8’
: E
x 1
p3 " g ‘3:
4.-.s E
.1 d3 s E
n5 .2
.2
Front cross section view of pectoral fin Time (s)
(a) (b)
Figure 7.10: (a) Definitions of the tip displacements and the twisting angle; (b) trajectories
of the tip displacements and the twisting angle corresponding to the voltage inputs as in
Fig. 7.9.
181 . . f .
180
’5?
/ 2
~x/pr 3179- ~~
1' O 3
_ y 2
l g 178 -
a
m x E177 . ,
x _ P2 :3;
/ 176-
/Pa 175 i
0 1 2 3 4 5 6
Front cross section view of pectoral fin Time (s)
(a) (b)
Figure 7.11: (a) Definition of the cupping angle; (b) the trajectory of the cupping angle
for a sample with 85 um thickness in both active and passive areas (voltages: 3 V, 0.3 Hz,
180°).
182
Table 7.4: Thicknesses of MDOF IPMC actuators.
Thickness in active area Thickness in passive area
S1 240 pm 240 um
S2 170 pm 170 um
S3 170 pm 120 um
S4 170 pm 60 um
SS 85 pm 85 um
the same voltage, a thinner sample experiences higher electrical field, and that a thinner
sample is more compliant. From Fig. 7 . 12(b), it can be seen that for samples with the same
thickness in active areas, with thinner passive regions, the deformation gets larger under
the same voltage inputs. This has thus provided supporting evidence for our approach of
modulating mechanical stiffness through plasma etching.
7.3.4 Impact of actuation signals
We have further experimented with actuation signals with different phases and amplitudes.
Table 7.5 shows all the actuation signals we have used in the experiments. We have se-
lected SI and S5 as the test samples. All the signals have the same frequency 0.3 Hz. To
study the effects of phase on the actuation performance, the voltage signals from Control
#1 to Control #5 have the same amplitude (3 V) but different phases. To understand the
performance of an MDOF IPMC actuator under different voltage levels, the voltage signals
from Control #5 to Control #7 have the same phase (90°) but different amplitudes.
Fig. 7.13(a) shows the twisting angle of SI actuated under different voltage levels but
with the same phase and frequency. While as expected, the higher the voltage, the larger the
twisting, the gain in deformation does not appear to be linearly growing with the voltage
level. Such nonlinearities will be examined in our future work. Fig. 7.13(b) shows the
twisting angle of S5 actuated by the same voltage amplitude but different phases. From the
figure, 900 appears to be the best phase to generate the twisting motion, the reason of which
183
Twisting angle (degree)
Time (s)
(a)
Twisting angle (degree)
-6 1
Time (s)
Figure 7.12: Twisting angles generated by different samples with 0.3 Hz, 3 V and 90° phase
sinusoid voltage signals. (a) MDOF IPMC actuators with different thickness in both active
areas and passive areas; (b) MDOF IPMC actuators with the same thickness in active areas
(170 pm) but different thicknesses in passive areas.
184
Table 7.5: Actuation signals.
Frequency Amplitude Phase
Control #1 0.3 Hz 3 V 180°
Control #2 0.3 Hz 3 V 120°
Control #3 0.3 Hz 3 V 60°
Control #4 0.3 Hz 3 V 0°
Control #5 0.3 Hz 3 V 90°
Control #6 0.3 Hz 4 V 90°
Control #7 0.3 Hz 8 V 90°
will be explored in our future modeling work.
7.3.5 DPIV study on MDOF IPMC actuation in water
In the interest of robotic fish applications, we have also conducted preliminary study of
underwater operation of the MDOF IPMC actuators. Digital Particle Image Velocimetry
(DPIV) system is used to observe fluid motion generated by the actuator. In a DPIV system,
small particles are dispersed in a fluid and a laser sheet is created in the fluid to illuminate
the particles. Processing of images taken in quick successions can reveal the movement of
particles and provide information about the flow fluid. We have tested sample S5 in water,
by applying voltage signals (4 V, 0.3 Hz, and 90° phase) to the actuator. Fig. 7.14(a) shows
the snapshots of the MDOF IPMC actuation in water. Fig. 7.14(b) shows the velocity field
of the water around the actuator. It demonstrates that the MDOF IPMC actuator can make
3-D deformation in water and can generate some interesting flow patterns around it. The
connections between the flow patterns and the actuator deformations are a subject of future
investigation.
185
2 . . r
1.5_ '---‘V=3Volt
- - -V=4 Volt
’ \
if 1- 1’ 7'“ —V=8 Volt ,
2 I \ . ' '
8) ’ \\ I
‘o 0.5' I ' \ ’ ’
V I I ‘ ~"'s, \ "r
2 l ." ‘ ‘ .”
g o» ’ ‘ ,
(u I \s “.a
C) ‘ v - '
s —o.5~ ‘\ ' .’
ZS \ I
E ‘1 \\ ’ I ’
-1.5-
-2 I
0 1 2 3 4
Time (s)
(a)
8 . . . "'""Phase=0°
' ,.\ "“Ph - °
‘1 .‘ ase-60 ~"\
6’ L' ' ‘1 ""‘Phase=90° ! ' -
' ‘ _ o .I
3 4_ r p 1‘. Phase—120 !
2 i. \_ Phase=180° v
a) .
a)
3
.03
or
r:
to
c»
E
E
E
3
Time (s)
(b)
Figure 7.13: (a) SI actuated by voltage signals with difl‘erent amplitudes but the same
frequency (0.3 Hz) and phase (90°); (b) S5 actuated by voltage signals with the same am-
plitude (3 V) and frequency (0.3 Hz) but difierent phases.
186
(a)
Figure 7.14: DPIV study of MDOF IPMC actuator operating in water. (a) Actuation of the
actuator in water (the edge of the actuator is highlighted); (b) Velocity field of the fluid.
187
7.4 Chapter Summary
In this chapter we have presented a new process flow for monolithic, batch-fabrication of
MDOF IPMC actuators. The methodology effectively incorporates standard techniques
for IPMC fabrication and lithography-based micromachining processes. The actuator con-
sists of multiple IPMC islands coupled through a passive membrane. The size and shape
of each IPMC island are defined through photolithography and thus the approach is scal-
able. A key innovation in the developed process is to stiffen the Nafion film through an
ion-exchange step, which virtually eliminates swelling in solvents and enables success-
ful patterning. Tailoring the thickness and thus the rigidity of the passive area is another
novel aspect of the proposed fabrication method. The fabrication method has been applied
to manufacture prototypes of biomimetic fins, with demonstrated capability of producing
complex deformation modes such as twisting and cupping. I
The prototypes fabricated in this work have dimensions of centimeters because of our
interest in biomimetic actuation for robotic fish applications. The developed fabrication
process, however, can be directly applied to produce microscale devices since it is based
on microelectromechanical systems (MEMS) processes. We also note that, while commer-
cially available Nafion films have been used as the base material in our work, the approach
can be easily extended to Nafion membranes obtained through casting, and to other types
of ion-exchange membranes (e.g., Flemion).
Future work includes understanding and modeling of MDOF IPMC actuators, by ex-
tending the authors prior work [21] on modeling of IPMC beams. The interactions between
the active IPMC regions and the passive regions will be a focus of study. We will also in-
vesti gate systematically the hydrodynamics in underwater operation of such MDOF IPMC
actuators using combination of analytical modeling, computational fluid dynamics (CFD)
modeling, and DPIV studies, and explore the use of these actuators in biomimetic robotic
fish.
188
Chapter 8
Conclusions and Future Work
8.1 Conclusions
In this work, a systems perspective has been taken to address the challenges in realizing
IPMC-based smart microsystems. To obtain a faithful and practical mathematical model for
IPMC, we have developed a physics-based, control—oriented modeling approach for IPMC
sensors and actuators. The effects of distributed surface resistance have been considered
in the models. The models are amenable to model reduction, geometrically scalable, and
practical in real-time control. The actuation model of IPMC has been applied to the model-
ing of robotic fish propelled by IPMC, which has further validated the modeling approach.
Since traditional sensors cannot be embedded easily into bio/micro systems, we have de-
veloped a compact sensing scheme for IPMC. We have further developed a process for
monolithic fabrication for IPMCs capable of generating complex deformations. Artificial
pectoral fins are fabricated and characterized, which have demonstrated bending, twisting,
and cupping motions.
189
8.2 Future Work
Future work can be pursued in the following three directions. First, the fabrication of the
artificial pectoral fins can be improved through enhanced lithography and electrode plating
techniques. The analysis of DPIV results for the pectoral fins will be another integral part
of this work, which can lead to the control strategies for the fin in robotic fish applica-
tions. Second, it will be interesting to extend the developed planar fabrication technology
to 3-dimensional fabrication technology which can be used in microfabricating more com-
plex bio-inspired IPMC materials, such as artificial lateral lines. The lateral line will have
multiple micro IPMC sensing hairs standing on the substrate to sense fluid flow. Third,
nonlinear models are necessary when a relatively high voltage is applied and a large de-
formation is generated. Many nonlinearities such as hysteresis, nonlinear elasticity of the
polymer, and nonlinear capacitance of IPMC are involved in IPMC. Model-based nonlin-
ear control should be explored to effectively control IPMC in applications involving large
deformations.
190
1
Appendix A
Appendix for Chapter 2
A.1 Derivation of impedance model (2.31)
From (2.28) and (2.29),
i(z s) : ¢(hh,(zs) sWKey(s)(s+K)
p ’ )(SS+Ktanh(1’()))
___ (1:3) /—z,rsdr——IP(ZS)) (S),
where the second equality is from (2.27) and (2.34). This results in
ip(z,s) 2 (VS) ——/Oz-;—l/is (T,s)dt') T123275.
From (2.26) and (2.30), one obtains
4) (h,Z,S) _ 4) (_hizas) +2lp(Z,S)l'§/W
Rip/W
2d) (h, z ,s)+2ip(z, s)r’2/W
2 213:0? Warm),
ik (275) :
191
(Al)
(A2)
where the last equality is from (2.27).
Combining (2.25), (A.l), (A.2), one gets
_8is (2,5) : A(s)V(s) _
T __2_ 3(5) [0‘ is(t,s)d1', (A3)
where A(s) and B(s) are as defined in (2.32) and (2.33).
Eq. (A3) is an integro-difierential equation for is. To solve this equation, we intro-
duce the unilateral Laplace transform for functions of the length coordinate z. The new
Laplace variable will be denoted as p since 3 has already been used for the transform of
time functions. For instance, the transform of is(z,s) will be defined as
Is(p,s)§/0 is(z,s)e-pzdz.
Now perform the Laplace transform with respect to the 2 variable on both sides of (A.3).
Using properties of Laplace transforms, one gets
p1. ( q. (s) q2(s> )
with
1
‘11 (S) = )3 q2(S
2\/:3( 2‘ / 8(5)
The surface current is(z, s) is then obtained from (A.6) using the inverse Laplace transform
OfIs (p,S)I
. _ . A (S) V (S) .
13 (2,5) _ 1s (O,s)cosh( B(s)z) —— m)— smh( B(s)z) . (A.7)
Using the boundary condition iS(L,s) = 0, one obtains:
V (3)/1 (s) tanh (ML)
is (0’s) 2 2 B(s)
(A.8)
193
Appendix B
Appendix for Chapter 4
B.1 Derivation of M(L,s), FC(L,s), Fd(z,s) in Section 4.2.3
Based on the principle of replacement in Section 4.2.3, one gets, for 0 g 2 S L,
L
Mme (23S) = / E: (r) (r -z> dr +F.(L,s> [<14 Malawian/ch] .111...) 1
M’ + [(1), (Lo) kb + (P1(L0)kc] W (Lo,S)
i=1
197
(3.8)
(13.9)
(B.10)
(13.11)
(13.12)
and with (4.32), the slope can be written as
MH215) = Hld(5)V(3) -
i Mm; (Lo)Q1 (S) [