Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-30T00:56:34.079Z Has data issue: false hasContentIssue false

Carbon Nanotube/Polyelectrolyte Composites as Novel Actuator Materials

Published online by Cambridge University Press:  15 March 2011

Debjit Chattopadhyay
Affiliation:
Nanomaterials Optoelectronics Laboratory, Department of Chemistry, Polymer Program, Institute of Materials Science, University of Connecticut, Storrs, CT 06269-3136
Izabela Galeska
Affiliation:
Nanomaterials Optoelectronics Laboratory, Department of Chemistry, Polymer Program, Institute of Materials Science, University of Connecticut, Storrs, CT 06269-3136
Ray Baughman
Affiliation:
Honeywell Co., 101 Columbia Road, Morristown NJ 07962-102
Anwar Zakhidov
Affiliation:
Honeywell Co., 101 Columbia Road, Morristown NJ 07962-102
Fotios Papadimitrakopoulos*
Affiliation:
Nanomaterials Optoelectronics Laboratory, Department of Chemistry, Polymer Program, Institute of Materials Science, University of Connecticut, Storrs, CT 06269-3136
*
*To whom the correspondence should be addressed. Email: papadim@mail.ims.uconn.edu. Tel.: (860) 486-3447, Fax: (860) 486-4547.
Get access

Abstract

Mechanical actuators that simultaneously provide high power densities and large force generation capacities are of great scientific and technological interest. Recently single wall carbon nanotube (SWNT) papers (“bucky” papers) were shown to possess significant promise as electrochemical actuators. Embedding polyelectrolytes, like Nafion™, within the nanotube matrix has the potential to address the limitations of SWNT bundling and tube slippage thus increasing force generation. In this paper two types of Nafion/SWNT composite actuators have been investigated depending on the method of fabrication. In the first case, infiltration of Nafion within SWNT paper matrix was followed by annealing at 150°C to invert Nafion's micellar structure and render it insoluble. This has resulted in a substantially exfoliated layer morphology that causes a reduction in both conductivity and actuation strain (c.a. 0.03%). In the second case, slow casting of a methanolic suspension of Nafion and SWNT soot, followed by annealing at 150°C, resulted in a more homogeneous structure. This composite, upon electrochemical cycling between -1 and +1 V in aqueous electrolytes, exhibited actuation strains (as high as 0.43%). However, these higher strains are accompanied by an order of magnitude reduction in modulus largely due to Nafion swelling.

Type
Research Article
Copyright
Copyright © Materials Research Society 2001

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

[1] Busch-Vishniac, I. J., Electromechanical Sensors and Actuators, 2 ed. New York: (Springer, 1999).Google Scholar
[2] Baughman, R. H., Synthetic Metals, 78, 339353, (1996).Google Scholar
[3] Baughman, R. H., Cui, C., Zakidov, A. Z., Iqbal, Z., Barisci, J. N., Spinks, G. M., Wallace, G. G., Mazzoldi, A., Rossi, D. D., Rinzler, A., Jaschinski, O., Roth, S., and Kertesz, M., Science, 284, 13401344, (1999).Google Scholar
[4] Treacy, M. M. J., Ebbesen, T. W., and Gibson, J. M., Nature, 381, 678680, (1996).Google Scholar
[5] Falvo, M. R., Curry, G. J., Taylor, R. M., Chi, V., Brooks, F. P., Washburn, S., and Superfine, R., Nature, 389, 582584, (1997).Google Scholar
[6] Ebbesen, T. W., Lezec, H. J., Hiura, H., Bennett, J. W., Ghaemi, H. F., and Thio, T., Nature, 382, 5456, (1996).Google Scholar
[7] Mojarrad, M. and Shahinpoor, M., “Ion exchange membrane-platinum composites as electrically controllable artificial muscles,” ICIM/ECSSM, Lyon, (1996).Google Scholar
[8] Salehpoor, K., Shahinpoor, M., and Razani, A., “Role of ion transport in actuation of ionic polymeric-platinum composite (IPMC) artificial muscles,” SPIE, San Diego, (1998).Google Scholar
[9] Shahinpoor, M., Bar-Cohen, Y., Simpson, J. O., and Smith, J., “Ionic-polymer-metal composites (IPMC's) as biomimetic sensors, actuators, and artificial muscles- a review,” Smart Materials and Structures, pp. R15–R30, (1998).Google Scholar
[10] Bar-Cohen, Y., Xue, T., Joffe, B., Lih, S. S., Shahinpoor, M., Simpson, J., Smith, J., and Willis, P., “Elctroactive polymers (EAP) low mass muscle actuator,” SPIE, (1997).Google Scholar
[11] Manley, D. S., Wiliamson, D. L., Noble, R. D., and Koval, C. A., Chem. Mater., 8, 25952600, (1996).Google Scholar
[12] Eisenberg, A. and Eager, H. L., Perfluorinated Ionomer Membranes, American Chemical Socicety, (1982).Google Scholar
[13] Rinzler, A. G., Liu, J., Dai, H., Nikolaev, P., Huffman, C. B., Rodriguez-Marcias, F. J., Boul, P. J., Lu, A. H., Heymann, D., Colbert, D. T., Lee, R. S., Fischer, J. E., Rao, A. M., Eklund, P. C., and Smalley, R. E., Applied Physics A, 67, 2937, (1998).Google Scholar
[14] Hone, J., Llaguno, M. C., Nemes, N. M., Johnson, A. T., Fischer, J. E., Walters, D. A., Casavant, M. J., Schimidt, J., and Smalley, R. E., Applied Physics Letters, 77, 666668, (2000).Google Scholar