Piezoelectric Polymeric Thin Films Tuned by Carbon

0 downloads 0 Views 3MB Size Report
properties of the PVDF material to be adjusted such that lower poling voltages can ... To compare the piezoelectric characteristics of the CNT-PVDF composite.
Piezoelectric Polymeric Thin Films Tuned by Carbon Nanotube Fillers Junhee Kim, Kenneth J. Loh, and Jerome P. Lynch* Dept. of Civil and Environmental Engineering, University of Michigan, 2328 G.G. Brown Building, Ann Arbor, MI USA 48109-2125 ABSTRACT Piezoelectric materials have received considerable attention from the smart structure community because of their potential use as sensors, actuators and power harvesters. In particular, polyvinylidene fluoride (PVDF) has been proposed in recent years as an enabling material for a variety of sensing and energy harvesting applications. In this study, carbon nanotubes (CNT) are included within a PVDF matrix to enhance the properties of PVDF. The CNTPVDF composite is fabricated by solvent evaporation and melt pressing. The inclusion of CNT allows the dielectric properties of the PVDF material to be adjusted such that lower poling voltages can be used to induce a permanent piezoelectric effect in the composite. To compare the piezoelectric characteristics of the CNT-PVDF composite proposed, scanning electron microscope (SEM) images were analyzed and ferroelectric experiments were conducted. Finally, the aforementioned composites were mounted upon the surface of a cantilevered beam to compare the voltage generation of the CNT-PVDF composite against homogeneous PVDF thin films. Keywords: energy harvesting; carbon nanotubes; PVDF; smart structures; piezoelectricity

1. INTRODUCTION Piezoelectric materials have been extensively adopted as sensors and actuators in a variety of smart structure applications due to their desirable electro-mechanical characteristics. Piezoelectric materials can serve as sensors since they produce electric voltages when mechanically deformed. They can also mechanically deform when an electrical field is conversely applied; hence, they can serve as actuators. Furthermore, the direct piezoelectric effect can be leveraged for energy harvesting purposes by generating current under ambient mechanical vibrations. With recent advances in wireless sensor technology, energy harvesting has been a topic of great interest since power harvesting can potentially overcome the limitations of conventional power supplies (i.e., replacing batteries on a regular basis). Among the surge of research focused on energy harvesting, piezoelectric-based power harvesting techniques have been at the forefront because of their various advantages including their simplicity, the possibility of unlimited lifetimes, and cost-effectiveness. However, there are still unresolved problems including the fact that piezoelectrics alone are insufficient to power most electronics [1]. Two different piezoelectric materials, i.e. lead zirconate titanate (PZT) and polyvinylidene fluoride (PVDF), have been conventionally used to generate electric power. Unlike PZT, which is a ceramic material, polymeric PVDF has various merits including high flexibility, low cost, high voltage output, and compatibility with other chemical species. These attributes render PVDF materials ideal for a variety of sensing applications. However, in spite of the various advantages of PVDF, such materials are limited in their ability to generate elastic stress waves within rigid structures. Hence, they are generally not used as actuators. Piezoelectricity in materials like PVDF results from polarization which is a separation of positive and negative charge at the molecular level of the material. For example, PVDF forms a dipole between fluorine and hydrogen along the length of the polymer chain when in the β crystalline phase (Figure 1). In pure PVDF, the β crystalline phase is attained by mechanical stretching. However, the copolymer of vinylidene fluoride (VDF) and trifluoroethylene(TrFE) (P(VDF-TrFE)) crystallizes into the β phase autonomously without a need for stretching it. That convenience has led researchers to fabricate and explore piezoelectric P(VDF-TrFE) films easily in the laboratory [2,3,4].

*

[email protected]; phone(734) 615-5290; http://www-personal.umich.edu/~jerlynch/ Sensors and Smart Structures Technologies for Civil, Mechanical, and Aerospace Systems 2008, edited by Masayoshi Tomizuka, Proc. of SPIE Vol. 6932, 693232, (2008) 0277-786X/08/$18 · doi: 10.1117/12.774256 Proc. of SPIE Vol. 6932 693232-1

2008 SPIE Digital Library -- Subscriber Archive Copy

qiboe WOL ueu

ax

20

cuJ

C

E

Fig. 1. Molecular Structure of PVDF.

CNTs have been proposed for inclusion within composite materials because to their notable physical and electrical properties [2,3,4,5,6,7,8,9]. On overcoming one of the major challenges of CNTs, namely the difficulty of uniformly dispersing CNTs into polymer matrices, CNTs can be ideal candidates for the reinforcement of a polymer-based composite from both electrochemical and electromechanical view points. Embedding CNTs into a polymer matrix can be a powerful means of mechanically reinforcing the composite. Furthermore, the conductive nature of CNT can be leveraged to tune the electrical properties of the final composite. In this study, CNTs are embedded in a PVDF copolymer to tune the piezoelectric properties of the resulting composite material. This paper presents early work in the fabrication of CNT-PVDF composites as well as the experiments used to quantify their piezoelectric properties.

2. BACKGROUND AND LITERATURE REVIEW Piezoelectricity arises from small changes in the internal polarization of a dielectric material under applied mechanical changes. Traditionally, piezoelectricity had been considered a characteristic of ceramic materials due their crystal lattice structure. One widely used ceramic piezoelectric material is PZT. However, since Kawai [10] discovered the considerable piezoelectric properties of PVDF, a number of researchers have been explored the piezoelectricity of different electro-active polymers (EAPs), such as polyamide, polyvinylidene chloride (PVC), nylon, among others [11]. The polarization mechanism of dielectric materials, especially polymeric piezoelectrics, consists of dipolar polarization and interfacial polarization. Dipolar polarization occurs when an electric field is applied. The dipoles inherent to the molecular structure rotate and align parallel to the applied field [12]. By its unique molecular structure (polar form, alltrans conformation), each unit cell of PVDF has a large dipole moment of 6.9 x 10-30 cm normal to the carbon chain axis as shown in Figure 1. However, there exists no net charge in the bulk material due to the random orientation of each polymer chain at the macro-scale. However, dielectric materials enjoy the property of ferroelectricity; upon a sufficiently high electric field, the materials can be polarized spontaneously. Termed “polling”, the piezoelectric properties of piezoelectric materials are usually enhanced by placing the material within an extremely high electric field (in the kV range). Therefore, dipole polarization can be generated at the macro-scale through poling of the PVDF or its copolymer materials (e.g., P(VDF-TrFE)). The tendency for polarization can be enhanced by adjusting the permittivity or dielectric constant of the piezoelectric polymer by incorporating a conductive phase within the material. This enhancement can be explained by the concept of interfacial polarization. Interfacial polarization arises for electrically heterogeneous materials, i.e. composites of different conductive materials such as CNTs-polymer composites (i.e., a nonconductive polymer matrix with conductive CNT inclusions) [8]. Mobile charge within the conductive phase (CNTs) can facilitate polarization of the composites. Therefore interfacial polarization results in an increase in the dielectric constant[9]. Research exploring interfacial polarization and the dielectric characteristics of CNT-PVDF composites have been recently reported [2,3,4]. Ramarantnam, et al.[2] fabricated CNT/P(VDF-TrFE) composite sensors by solvent casting methods and demonstrated an increase in sensing capability by adding CNTs. Even though their film’s piezoelectricity is lower than that of

Proc. of SPIE Vol. 6932 693232-2

conventional PVDF films, they showed the possibility of CNT-based sensors. Wang, et al.[3] studied the dependence of dielectric properties of untreated multiwall carbon nanotube(MWNT)/PVDF composites with different volume fractions of MWNT. They found low percolation threshold (2.0 vol % MWNT) of the MWNT/PVDF composites. Finally, they illustrated that the dielectric properties of MWNT/PVDF composites may be improved without the need for chemical functionalization of the CNTs. Dang, et al.[4] observed significant decreases in the dielectric properties of MWNT/PVDF composites before and after stretching. They explained their finding by the rearrangement of MWNT along the tensile-strain direction inducing a decrease in the dielectric constant. As previously mentioned, polarization can be induced by poling procedures The material’s piezoelectricity is directly correlated to the degree of polarization achieved by poling. Theoretically, the higher the electric field, the greater the piezoelectric properties achieved. However, for the piezoelectric polymer, the value of the electric field needed for polling is limited by the dielectric breakdown of the material. Practically, the dielectric breakdown of crystalline PVDF and its copolymers is 150 V/µm [14]. Furthermore, it is difficult to apply high electric fields to PVDF film fabricated in the laboratory due to problems associated with film texture, including surface impurities, cracking, and pinholes [15].

3. SWNT/P(VDF-TRFE) COMPOSITE FILMS Three different SWNT/P(VDF-TrFE) nanocomposite thin films with different concentrations of SWNT were fabricated by solvent casting and melt pressing. The nanotube percentages shown in Table 1 were chosen based on the possibility of poling the final SWNT/P(VDF-TrFE) composite. Previous work exploring the percolation threshold of SWNT/P(VDF-TrFE) nanocomposites revealed a threshold of 0.4 wt. % of SWNT with respect to the weight of P(VDFTrFE). This value is lower than that of Dang, et al. [3] suggesting excellent dispersion of SWNT in the P(VDF-TrFE) matrix. However, a high dielectric constant of the composite slightly below percolation would make poling difficult due to dielectric breakdown in local regions where CNTs are adjacent to each other. Thus, lower SWNT percentages, i.e. 0.05 and 0.1 wt. %, were selected in this study. 3.1 Solvent Casting The main purpose of solvent casting is to fabricate homogeneous SWNT/P(VDF-TrFE) thin film composites. Two separate solutions were needed to make the SWNT/P(VDF-TrFE) composites. First, 1g P(VDF-TrFE) pellets (65/35 wt.%, Kteck Corp.) were added into 2ml N,N-dimethylacetamide (DMF) solution (Fisher Scientific). The pellets were fully dissolved after placing them in a warm (60 ºC) sonication bath for 3 hours. The specified amounts (as shown in Table 1) of SWNTs (Carbon Nanotechnologies, Inc.) were dispersed in 2 ml DMF by 1 hour sonication in the bath. Upon mixing the two solutions at specific mass ratios, high powered probe sonication is conducted for 5 minutes. Finally, three P(VDF-TrFE) viscous solutions of different SWNT concentrations are obtained. Nanocomposite films were formed by thermally evaporating as-prepared SWNT/P(VDF-TrFE) solutions on top of aluminum foil as shown in Figure 2a. It took 12 hours to evaporate the DMF solvent with the help of elevated temperatures generated by the hot plate. Through complete evaporation, P(VDF-TrFE) thin films tuned by CNT inclusions were obtained. 3.2 Melt Pressing It is very difficult to control the film thickness and polymer crystallinity by the solvent casting procedure above. Furthermore, due to the high boiling point (128-136 ºC) of the DMF solvent, some solvent might still remain in the composite even after thermal evaporation. Additional melting and solidification can be applied to effectively solve these problems. The melting work was done using the Heated Presser 4386 (Carver, Inc.), which is controlled by digital thermometers (Figure 2b). Because the melting point of P(VDF-TrFE) varies from 150 to 170 ºC, depending on the content of TrFE, high temperature (180 ºC) well above the melting point was applied. Solidification is attained under the control of pressure and a specific cooling rate. Slow cooling under high pressure increases crystallinity, crystal size, and the ferroelectric and piezoelectric response of the final materials [13]. Specifically, 5 hours cooling to room temperature under a pressure of 2000 N/cm2 also anneals the composite. Each film’s thickness is measured using a micrometer. It should be noted that the thicknesses of three different films (Table 1) show minimal variance, thereby making quantitative comparisons of piezoelectricity of each film more reliable.

Proc. of SPIE Vol. 6932 693232-3

Table 1. Fabricated P(VDF-TrFE) nanocomposites.

Film Composition

SWNT (mg) 0.0 0.5 1.0

Pure P(VDF-TrFE) SWNT/P(VDF-TrFE) SWNT/P(VDF-TrFE)

SWNT (wt. %) 0% 0.05 % 0.1 %

Film Thickness (µm) 26 26 27

$ ':'n fli'

Sc

S

(a)

(b)

Fig. 2. SWNT/P(VDF-TrFE) film fabrication: (a) solvent casting; (b) melt pressing.

3.3 Specimen Preparation and Poling Each film is precisely cut into small specimens (10 x 25 mm2) with electrical contacts (15 x 15 mm2) made using colloidal silver paint (Ted Pella, Inc.). Poling was conducted by submerging each film in an insulating fluid, i.e. silicon oil (Fisher Scientific), to prevent arcing. A constant electric field (1500 V) generated by a DC power supply (E3642A, Agilent) and voltage amplifier (623B, Trek) was applied to each composite for 20 minutes at an elevated temperature (42 ºC to increase dipole mobility). Furthermore, the increased volume induced by the elevated temperature helped the film to conform to the volume change associated with orientation of molecular dipoles.

4. EXPERIMENTAL RESULTS AND DISCUSSION 4.1 Morphology of the P(VDF-TrFE) Composites The crystalline morphology, which affects the material’s overall piezoelectricity, was characterized by analyzing scanning electron microscope (SEM) images of the films. Due to the slow cooling rate and the high pressure applied during solidification, high crystallinity can be confirmed for P(VDF-TrFE) composites. Randomly distributed crystalline lamellae, shown as rod-like objects in Figure 3, prove high crystallinity and confirms other researchers’ results[16,17]. The crystalline lamellae are isolated packs of chain-folds of P(VDF-TrFE) molecules entangled without any special orientation. Slight differences in the surface texture between pure P(VDF-TrFE) and SWNT/P(VDF-TrFE) composite can be recognized with respect to crystallinity (Figure 3b). The width and length of rod-like crystalline lamellae are reduced and a rough surface is obtained in the SWNT composite film as compared to the pure P(VDF-TrFE) film. This difference can be thought to be due to the existence of CNTs. Further study is needed to investigate the effect of CNTs

Proc. of SPIE Vol. 6932 693232-4

l p

(a) Pure P(VDF-TrFE) film

(b) 0.1% SWNT/P(VDF-TrFE) Composite film Fig. 3. Scanning electron microscope (SEM) images of various P(VDF-TrFE) composite films.

upon the crystallization of the polymer composite; detail experiments, i.e. density measurement, wide-angle X-ray scattering (WAXS), and differential scanning calorimetry (DSC) are needed to determine the degree of crystallinity. 4.2 Polarization vs. Electric Field Curve Measurement The measurement of the polarization vesus electrical field curve, herein referred to as the PE curve, is needed. The PE curve is a hysteresis plot showing the polarization developed by an applied electric field at a certain frequency under a certain temperature. Experimental derivation of the PE curve is the most crucial step to prove the material’s ferroelectricity, since it represents the reversal of spontaneous polarization. Furthermore, based on some characteristic parameters, such as remanent polarization (Pr), saturation polarization (Ps), and coercive electric field (Ec) acquired from a PE curve, the appropriate poling voltage can be determined and the piezoelectric performance of the material can be optimized. Since its development by Sawyer and Tower in 1930 [18], the Sawyer-Tower circuit has been used as a convenient method for measuring PE curves. However, there are some difficulties inherent to Sawyer-Tower circuit, such as the parasite effect of the test material with low capacitance [21], and lack of adaptability with high impedance oscilloscopes [19,20,22,23]. In this research, the modified Sawyer-Tower circuit which includes two operational amplifier (LF412, National Semiconductor) (Figure 4) was used to measure the PE curve. Accurate PE curve measurements were conducted with a function generator (33120A, HP), AC voltage amplifier (623B, Trek), oscilloscope (54621D, Agilent), and modified Sawyer-Tower circuit. AC sinusoidal signals of 50Hz with three different amplitudes (1000, 1300, and 1500 V) were applied to the films at the same temperate (43 ºC). Figure 5 depicts the overlay of the measured PE curves for each voltage level. Each film’s characteristic values, i.e. coercive electric fields (Ec) which marks the point where the loop intersects the electric field axis, remanent polarization (Pr) which

Proc. of SPIE Vol. 6932 693232-5

C

Aok(sae 2SWbIS

02c11102cobG

EflUC!OL

S

CGUGLSIOL

a _t tJ —

(a)

(b)

Fig. 4. PE curve measurement: (a) experimental setup; (b) schematic of modified Sawyer-Tower circuit [21].

marks the point where the loop intersects the polarization axis, and saturation polarization (Ps) which marks the point where the tangential line at the loop edge intersects the polarization axis, are tabulated in Table 2. Under low electric fields (1000 V), all PE curves show needle-like loops, with an almost linear relation between polarization and electric field (Figure 5b). The slope of this linear line equals the permittivity (ε0ε); therefore, the dielectric constant (relative permittivity, ε) (Table 2) is calculated from the slope using the vacuum permittivity constant (ε0 = 8.8542 pF/m). The dielectric constant of the pure P(VDF-TrFE) film was compared with reference values [21,24], which consequently confirmed the quality of the fabricated films. It should be noted here that slight differences among each film’s dielectric constant proves that interfacial polarization does not generally affect the total polarization potential when compared under a relatively low electric field. Inclined narrow hysteresis loops for the 1300 volt electric field (Figure 5c) indicate the response of a lossy capacitor [25]. The dissipation of energy due to the inertia of the moving charge produces the evident phase difference between the charge and voltage signals [12]. It is apparent that the area within the loop, which is proportional to the loss of tangent (dissipation factor), depends on the CNTs concentration. At the 1500 volt electric field, both SWNT/P(VDF-TrFE) composites show rectangular hysteresis loops that prove the reversal of the spontaneous polarization. However, SWNT/P(VDF-TrFE) still shows some lossy capacitor behavior due to the absence of interfacial polarization. There is the distinct trend that the areas of each loop are proportional to the content of CNT in the composite. To explore the ferroelectricity more precisely, high voltage up to the full saturation level which makes the hysteresis loop rectangular should be applied. However, as mentioned in Section 2, even the 1500 volt electric field was challenging in this study. Based on the PE curve results, the poling voltages of each film was also fixed at the same value, i.e. 1500 V. 4.3 Energy Harvesting Test from Structural Vibration To validate and explore the piezoelectricity of the SWNT/P(VDF-TrFE) composites, tests exploring voltage generation due to vibration was conducted after poling each film. Three films were mounted with epoxy onto a cantilevered thin insulated plate (0.8mm - 31 x 18 cm2). A commercially poled PVDF film cut to the same size as the other specimens (28 µm thick from Measurement Specialties, Inc.) was also mounted to compare its performance as shown in Figure 6. One of the ends of the plate was fixed by a vice and the other end was connected to a modal exciter (Modal 110, MB Dynamics), which is employed to actuate the plate at a constant sinusoidal frequency. Two sinusoidal inputs with 40 Hz Table. 2. Characteristic properties of fabricated P(VDF-TrFE) composites.

Film Composition Pure P(VDF-TrFE) 0.05% SWNT/P(VDF-TrFE) 0.1% SWNT/P(VDF-TrFE)

Dielectric Constants (ε) 13.8 15.7 15.4

Remanent Polarization, Pr (mC/m2) 6.9 17.6 26.3

Saturation Polarization, Ps * (mC/m2) 17.3 24.9 32.5

Coercive electric Field, Ec * (MV/m) 24.3 36.6 44.1

Note : * Since the hysteresis loops are not saturated for 1500 volt, the true values are higher than those shown here.

Proc. of SPIE Vol. 6932 693232-6

bOI9L!9POL p P(VDF-TFE) 0.05% SWNT-P(VDF-TFE)

0 0.1% SWNT-P(VDF-TFE)

b

E

EI6CILIC E!61C

(a) Conceptual drawing and parameters

(b) Applied voltage : 1000V

(c) Applied voltage : 1300V

(d) Applied voltage : 1500V

Fig. 5. PE curves of SWNT/P(VDF-TrFE) composites at 50Hz sinusoidal electrical fields and 43℃

\\ (a)

(b)

Fig. 6. Experimental setup of cantilever plate with actuator: (a) SWNT/P(VDF-TrFE) films affixed to insulated plate; (b) side view.

and 110 Hz were selected for excitation. Under the same excitation, data acquisitions of each film’s response was accomplished by using an oscilloscope (54621D, Agilent) with a 0.2 msec sampling step.

Proc. of SPIE Vol. 6932 693232-7

11

j

p P(VDF-TFE)

— 0.05% SWNT-P(VDF-TFE) — 0.1% SWNT-P(VDF-TFE)

(a) Responses to 40 Hz sinusoidal excitation p P(VDF-TFE)

— 0.05% SWNT-P(VDF-TFE) — 0.1% SWNT-P(VDF-TFE)

(b) Responses to 110 Hz sinusoidal excitation Fig. 7. Measured voltage outputs and their frequency components.

Figure 7 depicts the overlay of measured output voltages as the thin aluminum plate is mechanically actuated at a frequency of 40 and 110 Hz; frequency components were also acquired through the use of a Fourier transformation. It should be noted here that clear time domain signals with small high frequency noise show the typical response of the properly poled piezoelectric materials. It is apparent that the piezoelectric performance is proportional to the content of CNTs. This can be explained by the fact that there is an increase of dielectric tendency induced by interfacial polarization due to CNTs during the high voltage poling stage. Furthermore, the 0.1% SWNT/P(VDF-TrFE) thin film generates almost the same magnitude of output voltage with respect to the commercially-obtained PVDF film, even though it was not poled up to the full saturation polarization level. Such a capable performance as compared to commercial film needs further investigation with the mechanism of interaction between CNTs and P(VDF-TrFE) more thoroughly explored.

5. CONCLUSIONS In this research, SWNT/P(VDF-TrFE) nanocomposites were fabricated by solvent casting, melt pressing, and annealing. Different concentrations of CNTs were added to a P(VDF-TrFE) matrix to produce a superior piezoelectric composite. The piezoelectric characteristics of each composite were explored by experimentation. Through analyzing SEM images, it was suggested that the existence of CNTs might hinder crystallization of the P(VDF-TrFE) matrix during solidification. Less crystallinity may result in a reduced dielectric constant. However, results from measurement of the polarization versus electric field curve show SWNT/P(VDF-TrFE) composites demonstrated higher dielectric constants and ferroelectricity depending on the weight fraction of CNT. Finally, through comparing generated voltages from

Proc. of SPIE Vol. 6932 693232-8

nanocomposites induced by vibration of a host structure, increased piezoelectricity of SWNT/P(VDF-TrFE) was demonstrated with respect to pure P(VDF-TrFE) under the same poling voltage. These increases of dielectric constant and piezoelectricity can be explained by the concept of interfacial polarization between the embedded CNTs phase and the polymer matrix.

6. ACKNOWLEDGEMENTS The authors would like to express their gratitude to the National Science Foundation (NSF) for sponsoring this research under grant number CMMI-0528867 (Program Manager Dr. S.C. Liu). The authors would like to acknowledge the assistance of Prof. Nick Kotov in the Department of Chemical Engineering, University of Michigan; his help was vital to the success of this research.

REFERENCES [1] [2] [3] [4]

[5] [6] [7]

[8] [9] [10] [11] [12] [13]

[14] [15] [16] [17] [18]

H. A. Sondano, D. J. Inman, and G. Park, “A Review of Power Harvesting from Vibration Using Piezoelectric Materials”, The Shock and Vibration Digest, Vol. 36, No. 3, pp.197-205, 2004 A. Ramaratnam, and N. Jalili, “Reinforcement of Piezoelectric Polymers with Carbon Nanotubes: Pathway to NextGeneration Sensors”, Journal of Intelligent Material Systems And Structures, Vol. 17, No. 3, pp. 199-208, 2006 Z. Dang, S. Yao, and H. Xu, “Effect of Tensile Strain on Morphology and Dielectric Property in Nanotube/polymer Nanocomposites”, Applied Physics Letters, Vol. 90, Iss. 1, 012907, 2007 L. Cai, H. Qu, C. Lu, S. Ducharme, P. A. Dowben, and J. Zhang, “Surface Structure of Ultrathin Copolymer Films of Ferroelectric Vinylidene fluoride (70%) with Trifluoroethylene (30%) on Graphite”, Physical Review B, Vol. 70, 155411, 2004 E. T. Thostenson and T. W. Chou, “Aligned Multi-Walled Carbon Nanotube-Reinforced Composites: Processing and Mechanical Characterization”, Journal of Physics D: Applied Physics, Vol. 35, No. 16, pp. 77-80, 2002 G. Decher, “Fuzzy Nanoassemblies Toward Layered Polymeric Multicomposites”, Science, Vol. 277, pp. 12321237, 1997 K. Loh, J. Kim, J. Lynch, N. W. S. Kam, and N. A. Kotov, “Multifunctional Layer-by-Layer Carbon NanotubePolyelectrolyte Thin Films for Strain and Corrosion Sensing”, Smart Materials and Structures, Vol. 16, pp.429-438, 2007 Z. Ounaies, C. Park, K. Wise, E. J. Harrison, “Electrical Properties of Single Wall Carbon Nanotube Reinforced Polyimide Composites”, Composites Science and Technology, Vol. 63, Iss. 11, pp. 1637-1646, 2003 Z. Ounaies et al., “Sensing/actuation Materials Made from Carbon Nanotube Polymer Composites And Methods”, US Patent(US 2006/0084752 A1), 2006 H. Kawai, “The Piezoelectricity of Poly(vinylidene Fluoride)”, Japanese Journal of Applied Physics, Vol. 8, pp. 975-976, 1969 E. Fukada, “History and Recent Progress in Piezoelectric Polymer Research”, Proc. of IEEE Ultrasonic Symposium, pp. 597-605, 1998 M. W. Barsoum, “Fundamentals of Ceramics”, Institute of Physics Publishing, Bristol and Philadelphia, 2003 M. C. Christie, J. I. Scheinbeim, and B. A. Newman, “Ferroelectric and Piezoelectric Properties of a Quenched Poly(Vinylidene Fluoride-Trifluoroethylene) Copolymer”, Journal of Polymer Science: Part B: Polymer Physics, Vol. 35, Iss. 16, pp.2671-2679, 1997 B. Dickens, E. Balizer, A. Dereggi, and S. Roth, “Hysteresis Measurements of Remanent Polarization and Coercive Field in Polymers”, Journal of Applied Physics, Vol. 72, Iss. 9, 4285, 1992 J. D. Foster, and R. M. White, “Electrophoretic Deposition of The Piezoelectric Polymer P(VDF-TrFE)”, Proc. of 201st Meeting of The Electrochemical Society, Philadelphia, 2002 L. M. P. Pinherio, M. R. Chaud, and R. Gregorio, “Miscibility and Morphology of PVDF/P(VDF-TrFE) Blends”, Materials Science Forum, Vol. 403, pp. 83-88, 2002 G. Zhu, J. Xu, Z. Zeng, L. Zhang, X. Yan, and J. Li, “Electric Field Dependence of Topography in Ferroelectric P(VDF/TrFE) films”, Applied Surface Science, Vol. 253-260, pp. 2498, 2006 C. B. Sawyer, and C. H. Tower, “Rochelle Salt as A Dielectric”, Physical Review, Vol. 35, pp. 269-273, 1930

Proc. of SPIE Vol. 6932 693232-9

[19] [20] [21] [22] [23] [24] [25]

G. Orenelas-Arciniega, and J. Reyes-Gomez, “A New Modification to the Sawyer-Tower Ferroelectric Hysteresis Loop Tracer”, Journal of the Korean Physical Society, Vol. 32, pp.380-382, 1998 T. Yoshimura, and N. Fujimura, “Polarization Hysteresis Loops of Ferroelectric Gate Capacitors Measured by Sawyer-Tower Circuit”, Japanese Journal of Applied Physics, Vol. 42, pp. 6011-6014, 2003 T.T. Wang, J.M. Herbert, and A.M. Glass, “The Application of Ferroelectric Polymers”, Chapman and Hall, New York, 1988 T. Furukawa, M. Date, and E. Fukada, “Hysteresis Phenomena in Polyvinylidene Fluoride under High Electric Field”, Journal of Applied Physics, Vol. 51, pp. 1135-1141, 1980 T. Furukawa, and G.E. Johnson, “Measurement of Ferroelectric Switching Characteristics in Polyvinylidene Fluoride”, Applied Physics Letters, Vol. 38, Iss. 12, 1027, 1981 Y. Kubouchi, Y. Kumetani, T. Yagi, T. Masuda, and A. Nakajima, “Structure and Dielectric Properties of Vinylidene Fluoride Copolymers”, Pure And Applied Chemistry, Vol. 61, pp.83-90, 1989 M. G. Cain, M. Stewart, M. G. Gee, G. J. Hill, and D. A. Hall, “Electronic Property Measurements for Piezoelectric Ceramics: Technical Notes”, NPL Report CMMT(A)98, Center for Materials Measurement and Technology of National Physical Laboratory(NPL), UK, 1998

Proc. of SPIE Vol. 6932 693232-10