Open access

A Review on Electronic Transport Properties of Individual Conducting Polymer Nanotubes and Nanowires

Written By

Yun-Ze Long, Zhaojia Chen, Changzhi Gu, Meixiang Wan, Jean-Luc Duvail, Zongwen Liu and Simon P. Ringer

Published: 01 February 2010

DOI: 10.5772/39488

From the Edited Volume

Nanowires Science and Technology

Edited by Nicoleta Lupu

Chapter metrics overview

5,149 Chapter Downloads

View Full Metrics

1. Introduction

Quasi-one-dimensional nanostructures, such as carbon nanotubes, inorganic semi-conducting nanotubes/wires, and conjugated polymer nanotubes/wires, have drawn considerable attention in the past 20 years due to their importance for both fundamental research and potential applications in nanoscale devices (Kuchibhatla et al., 2007, Xia et al., 2003, MacDiarmid, 2002). Since the electrical conductivity of conjugated polymers can be increased by many orders of magnitude from 10-10-10-5 to 103-105 S/cm upon doping (MacDiarmid, 2002), conducting polymer nanotubes and nanowires (e.g., polyacetylene, polyaniline (PANI), polypyrrole (PPY), and poly(3,4-ethylenedioxythiophene) (PEDOT), poly(p-phenylenevinylene) (PPV)), are promising materials for fabricating polymeric nanodevices such as field-effect transistors (Aleshin, 2006), actuators (Jager et al., 2000), bio- and chemical sensors (Huang et al., 2003, Ramanathan et al., 2004), nano light emitting diodes, electrochromic displays (Cho et al., 2005), artificial muscles, and solar cells, etc. (Zhang & Wang, 2006)

By now, conducting polymer nanotubes and nanowires can be prepared by various methods such as the template-guided synthesis (Martin, 1994), template-free method (Wan, 1999), interfacial polymerization (Huang et al., 2003), electrospinning (MacDiarmid, 2001), dilute polymerization (Chio & Epstein, 2005), reverse emulsion polymerization method (Zhang et al., 2006), etc. The template method of polymerization proposed by Martin et al. is an effective technique to synthesize polymer micro-/nanotubes and wires with controllable length and diameter (Cai & Martin, 1989, Cai et al., 1991, Parthasarathy & Martin, 1994, Martin, 1994& 1995). The disadvantage of this method is that a post-synthesis process is needed in order to remove the template. The template-free method developed by Wan et al. is a simple self-assembly method without an external template (Wan et al., 1999& 2001, Zhang et al., 2004, Huang et al., 2005, Wan, 2008& 2009, Long et al., 2003). By controlling synthesis conditions such as temperature and molar ratio of monomer to dopant, polyaniline and polypyrrole nanostructures can be prepared by in-situ doping polymerization in the presence of protonic acids as dopants. The self-assembled formation mechanism in this approach is that the micelles formed by dopant and/or monomer-dopant act as soft templates in the process of forming tubes. The interfacial polymerization method proposed by Kaner et al. (Huang et al., 2003, Huang & Kaner, 2004) involves step polymerization of two monomers or agents, which are dissolved respectively in two immiscible phases so that the reaction takes place at the interface between the two liquids. Interfacial polymerization has been used to prepare various polymers, such as polyaniline nanofibers and nanotubes. Electrospinning is an effective approach to fabricate long polymer fibers using strong electrostatic forces (MacDiarmid, 2001, Tan et al., 2008). For instance, submicron fibers of doped polyaniline blended with polyethylene oxide or pure polyaniline have been prepared by this technique (MacDiarmid et al., 2001, Cárdenas et al., 2007). It should be noted that various spectroscopic results have indicated that the polymer micro- and nanostructures produced by these methods are usually partially crystalline; in other words, the small metallic regions of aligned polymer chains are interspersed with amorphous regions where the chains are disordered. The crystalline fraction depends on synthesis methods and synthesis conditions. At present, fabrication of highly crystalline and metallic polymer nanotubes and nanowires is still a challenge.

In order to fulfill the potential applications of conducting polymer nanotubes and wires, it is necessary to understand the electronic transport properties of individual polymer tubes/wires. The electrical characterization of individual conducting polymer nanotubes/ wires has made significant progress during the last decade. There are several strategies for measuring the conductivity of the template-synthesized fibres. The easiest and usual way is to leave the fibres in the pores of the template membrane and measure the resistance across the membrane (Cai & Martin, 1989, Cai et al., 1991, Parthasarathy & Martin, 1994, , Martin 1995, Granström and Inganäs, 1995, Duchet et al., 1998, Mativetsky and Datars, 2002, Duvail et al., 2002& 2004). Provided the number and diameter of the fibres are known, the measured trans-membrane resistance can be used to calculate the conductivity of a single fiber. However, this method may result in huge uncertainties on values due to the unknown number of connected fibres. Another way is to measure the resistances of compressed pellet or films (membrane removed) of the polymer nanofibres (Parthasarathy & Martin, 1994, Spatz et al., 1994, Orgzall et al., 1996). In fact, all these approaches did not realized the conductivity measurement of an individual fiber directly. Recently, the conductivity measurement of single polyaniline or polypyrrole tube/wire was achieved based on a conductive tip of an atom-force microscope (Park et al., 2002, Park et al., 2003, Saha et al., 2004, Liu et al., 2006). In this two-probe geometry, the contact resistance can be minimized by applying a significant pressure of the tip onto the nanotube/wire. A common approach was generally realized by dispersing nanotubes/wires on patterned micro- or nano-electrodes prepared by photo-lithography, electron-beam lithography and focused-ion beam deposition, followed by the subsequent searching of nanofibers just lying on the two or four electrodes only (Kim et al., 1999, MacDiarmid et al., 2001, Park et al., 2001, Park et al., 2003, Lee et al., 2004, Aleshin et al., 2004, Samitsu et al., 2005, Kim et al., 2005, Joo et al., 2005, Aleshin, 2006, Gence et al., 2007& 2008, Callegari et al., 2009). Particularly, focused-ion beam assisted deposition technique has been employed to attach metal microleads on isolated nanotubes/wires directly (Long et al., 2003a, 2004b& 2005b , Zhang et al., 2006 , Huang et al., 2006, Long et al., 2006c, Duvail et al., 2007, Lu et al., 2007, Long et al., 2008a& 2009b). There are also reports demonstrating a directed electrochemical nanowire assembly technique for the fabrication and measurement of polymer nanowire arrays between pre-patterned electrodes (Ramanathan et al., 2004). All these recent investigations contribute significantly to identify and understand the specific electrical behaviour of conjugated polymer nanwires and nanotubes in comparison to the bulk materials. Though a lot of efforts have been done, there are still some key questions needed to be clarified, for example, the effects of the nanocontacts on the electrical measurements, the differences in electrical properties between that of polymer nanotubes/wires and that of their bulk counterparts, the possibility of tuning and controlling the electrical properties of individual nanotubes/wires and so on. These questions are very important to fabrication and characterization of nanodevices based on individual nanofibers.

In this chapter we provide a brief review of recent advances in the study of electronic transport properties (e.g., size effect in electrical conductivity, nonlinear current-voltage characteristics, small magnetoresistance effect, and nanocontact resistance effect) of individual conducting polyaniline, polypyrrole and PEDOT nanotubes/wires.

Advertisement

2. Experimental

2.1. Preparation and characterization

The results reported in this review have been measured for conducting polymer nanotubes and nanowires prepared by template-free self-assembly method and template-guided method.

The protonic acids doped polyaniline and polypyrrole nanotubes/wires were prepared by the template-free self-assembly method. The polyaniline nanotubes are chosen as an example to introduce the synthesis procedure. Aniline monomer was distilled under reduced pressure. Ammonium persulphate as an oxidant and camphor sulfonic acid (CSA) as a dopant were used without any further treatment. In a typical synthesis procedure, aniline monomer (0.002 mol) and CSA (0.001 mol) were mixed in distilled water (10 ml) with stirring. The mixture reacted and formed a transparent solution of CSA-aniline salt. Before oxidative polymerization, the solution was cooled in an ice bath. Then an aqueous solution of ammonium persulphate (0.002 mol in 5 ml of distilled water) cooled in advance was added slowly into the above cooled CSA-aniline salt solution. After all the oxidant was added, the mixture was allowed to react for 15 h in the ice bath. The precipitates were then filtered and washed with distilled water and ethanol for several times, and finally dried at room temperature in a dynamic vacuum for 24 h. It was noted that if the synthesis temperature was changed to -10 C, polyaniline microspheres would be obtained (Long et al., 2004a). p-toluene sulfonic acid and 8-hydroxyquinoline-5-sulfonic acid doped polypyrrole tubes/wires were carried out along similar lines (Huang et al., 2005). The PEDOT nanowires were prepared in templates of polycarbonate track-etched membranes (Duvail et al., 2002, 2004& 2008b). After the polymerization, polycarbonate (membrane template) was removed by dissolution with a flow of dichloromethane and the nanowires were dispersed onto a SiO2 wafer.

The resulting polyaniline, polypyrrole, and PEDOT nanotubes/wires were characterized by elemental analysis, field-emission scanning electron microscopy (SEM), transmission electron microscopy, infrared and Raman spectroscopy, x-ray diffraction, x-ray photoelectron spectra and electron spin resonance (Zhang et al., 2004, Huang et al., 2005, Duvail et al., 2002& 2004). Fig. 1 shows the SEM and TEM images of the obtained polyaniline and polypyrrole nanotubes. The outer diameter is about 80-400 nm for the self-assembled polyaniline and polypyrrole tubes/wires and 20-190 nm for the template-synthesized PEDOT wires. It was found that the polymer tubes/wires prepared by the template-free self-assembly method show a partially crystalline character according to the x-ray diffraction patterns. This partially crystalline feature was also proved by specific heat (Long et al., 2004b) and magnetic susceptibility (Long et al., 2006b) studies on polymer nanotubes/wires.

Figure 1.

Typical SEM images of self-assembled polyaniline nanotubes (a and b) and polypyrrole nanotubes (c and d). Typical TEM images of polyaniline nanotubes (e) and polypyrrole nanotubes (f).

2.2. Fabrication of Pt microleads and electrical measurements

The method used to attach Pt microleads on an isolated polymer nanotube/wire was described in previous publications (Long et al., 2003a& 2005b). First, polymer nanotubes/ wires were ultrasonically dispersed in ethanol for template-free prepared nanotubes/wires and in dichloromethane for template-prepared PEDOT nanowires. Then, a drop of solution was placed on an insulating SiO2/Si wafer. After the evaporation of the solvent, an electron microscope was used to find an appropriately isolated nanotube/wire on the wafer. At last, two pairs of Pt microleads typically 0.5 μm in width and 0.4 μm in thickness were fabricated by FIB deposition (Dual-Beam 235 FIB System from FEI Company, working voltage of the system is 5 kV for the electron beam and 30 kV for the focused-ion beam, respectively, current of the focused-ion beam is very small, 1-10 pA, to minimize the modification of the conjugated polymer). Fig. 2 shows the individual polymer nanotubes/wires and the attached Pt microleads. It was noticed that the resistance of Pt microleads (less than 1 kΩ) is negligibly small compared with that of a single polymer nanotube/wire (several tens or hundreds of kΩ typically).

Figure 2.

Typical SEM images of an isolated polyaniline nanotube/wire (a, b and c) and polypyrrole nanotube/wire (d), and the attached platinum microleads. (Long et al., 2003a& 2005b)

The electrical connection between the Pt microleads and the sample holder was made by highly conductive silver paste and gold wires and the electrical measurements of individual polymer nanotube/wire were carried out using a Physical Property Measurement System from Quantum Design and a Keithley 6487 picoammeter/voltage source, or a Keithley 236 source-measure unit in an Oxford helium gas flow cryostat covering a wide temperature range from 300 down to 2 K. The four-probe resistance was measured by applying a very small current in the linear part of the I-V characteristics. The two-probe resistance was determined under V bias = 0.02 V, while no rectifying behaviour has been measured in our samples. The I-V curves were obtained by scanning the voltage from -4 to 4 V with a step of 0.02 V. The dI/dV curves were numerically derived from the corresponding I-V curves. The same polymer nanotube/wire was used for the four-probe measurement first and then for the two-probe measurement. The resistance of the polymer nanotube/wire with a given diameter was measured at least two times, for example, under cooling and during heating with a good reproducibility. In addition, for nanotubes/wires with a given diameter, two or more individual nanowires were measured to check the reproducibility.

Advertisement

3. Electronic transport properties

3.1. Diameter and temperature dependent electrical conductivity

The dependence of electrical conductivity on the diameter of the polymer nanotubes/wires (prepared by the template method) at room temperature has been widely reported (Cai et al., 1991, Parthasarathy & Martin, 1994, , Martin et al.1995, Granström & Inganäs, 1995, Duchet et al., 1998). It was found that the room-temperature conductivities of nanotubes/wires of conducting polypyrrole, polyaniline et al. can increase from 10-1-100 to about 103 S/cm with the decrease of their outer diameters from 1500 to 35 nm. The possible reason can be ascribed to the enhancement of molecular and supermolecular ordering (alignment of the polymer chains). For PEDOT nanowires prepared by template method, the room-temperature conductivities of the nanowires with diameters of 190, 95-100, 35-40, and 20-25 nm are about 11.2, 30-50, 490-530, and 390-450 S/cm, respectively (Duvail et al., 2007&2008a). For polypyrrole nanotubes prepared by template-free method, as shown in Fig. 3, it was found that the polypyrrole tube with a 560-400 nm outer diameter is poorly conductive and the room-temperature conductivity is only 0.13-0.29 S/cm. When the outer diameter decreased to 130 nm, the conductivity of the single nanotube increased to 73 S/cm (Long et al., 2005b). Such conductivity dependence on diameter was observed not only for template-synthesized polymer tubes/wires but also for self-assembled polypyrrole tubes, which indicates that the polymer tubes/wires prepared by these two different methods have similar structural characteristic: the smaller the diameter, the larger the proportion of ordered polymer chains.

Figure 3.

Diameter dependence of room-temperature conductivity of individual polypyrrole micro-/nanotubes prepared by template-free self-assembly method. (Long et al., 2005b)

Since the electrical properties of conducting polymers are strongly influenced by the effect of disorder and temperature, three different regimes (namely, insulating, critical, and metallic regimes close to the metal-insulator transition) have sorted out based on the extent of disorder and conductivity dependence on temperature (Yoon et al., 1994; Menon et al., 1998, Heeger, 2002). In the insulating regime, for a three-dimensional system, the temperature dependent resistivity usually follows Mott variable-range hopping (VRH) model: ρ(T) = ρ0exp(T M/T)1/4. At lower temperatures, when the Coulomb interaction between charge carriers is significant, ρ(T) usually follows Efros-Shklovskii (ES) VRH: ρ(T) = ρ0exp(T ES/T)1/2. In the critical regime, for a three-dimensional system close to the metal-insulator transition, the resistivity follows the power-law dependence: ρ(T)T -, where lies within the range of 0.3<<1. In the metallic regime, the sample shows a positive temperature coefficient of the resistivity at low temperatures (for example, below 10-20 K for metallic polypyrrole films, Menon et al., 1998).

Figure 4.

The dependence of resistance on temperature of a single polyaniline nanotube (σRT = 47 S/cm) and a single polypyrrole nanotube (σRT = 0.8 S/cm) in the insulating regime of the metal-insulator transition: (a) and (c) plotted as ln R(T) versus T-1/4 ; (b) and (d) plotted as ln R(T) versus T-1/2; the temperature dependence of resistance follows 3D-VRH at higher temperatures and ES-VRH at lower temperatures. (Long et al., 2005b)

Long et al. reported temperature dependent resistivity of a single polyaniline nanotube with average outer and inner diameters of 120 nm and 80 nm, which falls in the insulating regime of metal-insulator transition (Long et al., 2005b). The tube’s room-temperature conductivity is 47 S/cm. It was found that the resistivity follows three-dimensional (3D) Mott-VRH above 66 K, and follows ES-VRH model below 66 K, as shown in Figs. 4a and 4b. Here it is noted that from the view point of electrons, the polymer tube/wire with an outer diameter of 120 nm is still three dimensional because the localization length of carriers (LC< 20 nm) is much smaller than the wall thickness or the diameter of the submicrotube. Similar smooth crossover from Mott-VRH to ES-VRH has also been observed in a single polypyrrole microtube (room-temperature conductivity, 0.8 S/cm) at around 96 K (as shown in Figs. 4c and 4d, Long et al., 2005b). However, the crossover temperature (Tcros ~ 66-96 K) and the characteristic ES temperature (TES ~ 316-780 K) of a single polymer tube/wire are much higher than those of a polyaniline pellet or a polypyrrole film (Tcros< 15 K and TES ~ 29-56 K), which could be possibly due to enhanced strong Coulomb interaction in polymer nanotubes/wires at low temperatures.

In addition, with the decrease of disorder or diameter of polymer nanotubes/wires, Long et al. found that a 130-nm polypyrrole nanotube with room-temperature conductivity of 73 S/cm is lying close to the critical regime of metal-insulator transition (Long et al., 2005b). Its resistivity follows the power-law dependence: ρ(T)T-, as shown in Fig. 5. The fit yields a value of 0.488. Duvail et al. reported that a 100-nm PEDOT nanowire (σRT ~ 50 S/cm) fell in the critical regime with a value of ~ 0.78 (Duvail et al., 2007). Furthermore, the 35-40 nm template-prepared PEDOT nanowire (σRT ~ 490 S/cm) displays a metal-insulator transition at about 32 K, indicating that the nanowire is lying in the metallic regime (Duvail et al., 2007). However, for a PEDOT nanowire with a diameter of 20-25 nm, though its conductivity is relatively high at room temperature (~ 390-450 S/cm), the nanowire shows very strong temperature dependence (R(10K)/R(300K) ~ 105) or insulating behavior at low temperature. This is possibly due to a confining effect since the value of the diameter (20-25 nm) becomes equal or close to the localization length of electrons (Lc ~ 20 nm). In such a case, localization of electrons induced by Coulomb interaction or small disorder must be taken into account for explaining this insulating behavior at low temperatures.

Figure 5.

The dependence of resistance on temperature (ln R(T) versus ln T plot) for a single 130-nm polypyrrole nanotube in a conductive state close to the critical regime of the metal-insulator transition. (Long et al., 2005b)

3.2. Nonlinear I-V characteristics

The current-voltage (I-V) characteristics of individual polymer nanowires/tubes such as the polyacetylene, polyaniline, polypyrrole, and PEDOT have been explored extensively in the past ten years (Park et al., 2001, Park et al., 2003, Kaiser et al., 2002, 2003& 2004; Aleshin et al., 2004, Long et al., 2005b). With lowering temperature, a transition from linear to nonlinear I-V characteristics is usually observed (Fig. 6a), and a clear zero bias anomaly (i.e., Coulomb gap-like structure) gradually appears on the differential conductance (dI/dV) curves (Fig. 6b). Similar transition has also been reported in carbon nanotubes (Kang et al., 2002) and inorganic compound nanowires such as CdS nanoropes (Long et al., 2005a& 2008b), K0.27MnO20.5H2O nanowires (Long et al., 2008c), ZnO (Ma et al., 2005) and SnO2 (Ma et al., 2004) nanowires.

Up to now, several theoretical models such as the space-charge limited current, fluctuation-induced tunneling (Kaiser et al., 2003& 2004, Kaiser and Park, 2005), Coulomb gap (Kang et al., 2002,

Figure 6.

I-V characterisitics (a) and the corresponding differential conductance (dI/dV) curves (b) of a single polypyrrole nanotube at low temperatues.

Ma et al., 2004, Long et al., 2005b& 2008c), Coulomb blockade (Saha, 2002, Aleshin et al., 2005, Long et al., 2008b), Lüttinger liquid (Aleshin et al., 2004), Wigner crystal (Rahman and Sanyal, 2007) models, etc. have been considered to explain the conduction mechanism of quasi-one dimensional nanofibers. Saha and Aleshin et al. reported single-electron tunneling or Coulomb-blockade transport in conducting polypyrrole and helical polyacetylene nanofibers (Saha, 2002, Aleshin et al., 2005) separately. In addition, power-law behaviors for both I-V characteristics and electrical conductance G(T) have been reported recently in polyacetylene fibers (Aleshin et al., 2004) and polypyrrole wires/tubes (Rahman and Sanyal, 2007), which are characteristics of one-dimensional systems composed of several Lüttinger liquids or Wigner crystals connected in series, owing to electron-electron interactions (repulsive short-range electron-electron interactions or long-rang Coulomb interactions). Particularly, Kaiser et al. (Kaiser et al., 2004, Kaiser and Park, 2005) recently proposed a generic expression (extended fluctuation-induced tunneling and thermal excitation model) for the nonlinear I-V curves based on numerical calculations for metallic conduction interrupted by small barriers:

G = I / V = G 0 × exp ( V / V 0 ) / { 1 + h [ exp ( V / V 0 ) 1 ] } E1

where G0, V0 and h are parameters: G0 is the temperature-dependent zero-bias conductance; V0 is the voltage scale factor, which strongly depends on the barrier energy. Kaiser et al. showed that this expression could give a very good description of the observed nonlinearities in polyacetylene nanofibers, vanadium pentoxide nanofibers, etc. Here, one question arises for the Kaiser expression: Is it still appropriate to fit the nonlinear I-V characteristics of individual polymer nanowires/tubes if the Coulomb interactions are strong and should be taken into account?

The Kaiser expression has been used by Long et al. (Long et al., 2009b, Yin et al., 2009) to numerically calculate the I-V characteristics of individual polyaniline nanotube, polypyrrole nanotubes, PEDOT nanowires, CdS nanorope, and K0.27MnO2 nanowire, as shown in Fig. 7a. The fitting results indicate that except at low temperatures and low bias voltages, the Kaiser generic expression can give a good description of the I-V characteristics of individual nanotubes/wires, because the Kaiser expression (extending the Sheng model or fluctuation-induced tunneling and thermal excitation model) has well included the microstructure feature and the conduction feature of conjugated polymer nanofibers (quasi-one-dimensional metallic conduction interrupted by small barriers). Apparent deviation from the Kaiser expression has been evidenced in the low-temperature I-V curves as shown in Fig. 7b).

Figure 7.

I-V characterisitics of single polypyrrole nanotube with fits to expression (2), at temperatures (a) ranging rom 300 K to 100 K and (b) from 80 K to 10 K. (Long et al., 2009b)

Figure 8.

Zero-field conductance versus temperature, where G0 is determined from the fitting data and G0‘ is determined from the experimental data. (Long et al., 2009b)

Particularly, we compare the values of zero-bias conductance determined from the fitting parameter (G0) with that determined from experimental measurements (G0‘, obtained from the I-V curve or the differential conductance). As shown in Fig. 8, the fitting parameter G0 decreases smoothly with temperature lowering, but the experimental value G0‘ sharply decreases below 80-100 K and deviates from G0, although it becomes superposable to G0 for temperature equal and larger to 100 K. We note that the deviation temperature (about 80 K) is close to and consistent with the crossover temperature (66-96 K) for the crossover from Mott-VRH to ES-VRH, as shown in Fig. 4.

We propose that one possible reason for the deviation is that the Kaiser expression does not include the contributions from the Coulomb-gap occurring in density of states near Femi level and/or enhanced Coulomb interactions due to nanosize effects, which become important at low temperatures and voltages. (Long et al., 2009b& 2005b, Yin et al., 2009)

3.3. Magnetoresistance

The magnetoresistance (MR, defined as MR=ΔR(H)/R(0)=[R(H)-R(0)]/R(0)) of bulk films of conducting polymers have been extensively studied in the past 20 years (Menon et al., 1998). For example, polyaniline, polypyrrole, PEDOT films, and polyaniline composites usually exhibit a positive magnetoresistance at low temperatures (T<10 K) and MR H2 (H is not very large). The mechanism generally involved is the shrinkage of localised wavefunctions of electrons in the presence of a magnetic field or electron-electron interactions (Menon et al., 1998). Whereas highly conductive polyacetylene films usually show a negative magnetoresistance at low temperatures, which is mainly attributed to the weak localization effects (Menon et al., 1998, Kozub et al., 2002). Up to date, only a few papers have reported the magnetoresistance of polymer nanotubes/wires (Kim et al., 1999, Park et al., 2001, Kozub et al., 2002, Long et al., 2006a, 2006c& 2009c).

Figure 9.

The magnetoresistance curves for different temperatures of (a) a single polyaniline nanotube and (b) a pellet of polyaniline nanotubes. (Long et al., 2006a)

Long et al. reported that the magnetoresistance of single polyaniline nanotube and single PEDOT nanowire is positive below 10 K and increases as H2 up to 9 T. Typically, a positive magnetoresistance is expected for hopping conduction, because applying a magnetic field results in a contraction of the overlap of the localized state wavefunctions and thus an increase in the average hopping length. This corresponds to a positive magnetoresistance at sufficiently low temperatures. The theory of positive magnetoresistance has been developed for two cases, without and with electron-electron interactions. In both cases the weak-field MR with strong temperature dependence can be expressed as ln(R(H)/R(0))H2T-3/4 (Shklovskii & Efros 1984). However, the magnetoresistance of a single nanotube/wire is much smaller than that of the nanotube/wire pellet at 9 T: MR<5% (2K) for the single nanotube/wire (Fig. 9a), and MR ~ 90% (3K) for the polyaniline nanotubes’ pellet (Fig. 9b, Long et al., 2006a). In addition, when the temperature increases, the magnetoresistance of the single nanotube/wire becomes smaller and close to zero. No evident transition from positive magnetoresistance to negative one was observed. In contrast to that of single nanotube/wire, pellets of polyaniline and polypyrrole nanotubes/wires show a relatively larger positive magnetoresistance at low temperatures. With temperature increasing, there is a transition from a positive magneto-resistance to a small negative magnetoresistance at about 60 K. The results indicate that the magnetoresistance in the bulk pellet samples made of polymer nanotubes/wires is dominated by a random network of inter-fibril contacts.

The small magnetoresistance effect in individual polymer nanotube/wire has been confirmed in other samples. For example, the low-temperature magnetoresistance (MR~0.1%) in a polyacetylene nanofiber network is rather smaller than that in a bulk polyacetylene film (Kim et al., 1999, Park et al., 2001). A single gold/polyaniline microfiber shows a small positive magnetoresistance (MR<4.1%) below 6 K (Long et al., 2006c). The reason for this weak magnetoresistance effect in individual polymer nanotube/wire is possibly due to the elimination of inter-nanotube/wire contacts, small size and, relatively high conducitivity of individual polymer nanotube/wire. (Park et al., 2001, Long et al., 2006a)

3.4. Nanocontact resistance

The contact resistance is often encountered when we study electronic transport in an individual polymer nanowire/tube or polymer nanofiber-based nano-devices. As we know, there are two major factors responsible for the contact resistance magnitude: geometry and insulating layers (potential barriers) between the contacting surfaces. The resistance of a contact is inversely proportional to its area, and is dependent on the surface stiffness and the force holding the two surfaces together.The insulating layers (potential barriers) between the polymer nanowire and the metal electrode are usually formed due to their different energy levels or work functions. A bad (insulating or semiconducting) electronic contact may possess a strongly temperature dependent contact resistance, and thus can seriously complicate or even dominate the measured resistance. In this section, we discuss two kinds of nanocontact resistance: between two crossed polymer nanowires/tubes and between polymer nanowire/tube and metal microlead.

Figure 10.

SEM images showing two crossed polyaniline nanotubes and their attached Pt microleads. (Long et al., 2003c)

The nanocontact resistance between two crossed polymer nanotubes/wires has been studied by Long et al. (Long et al., 2003c& 2009a). It was found that the inter-tubular junction resistance of two crossed polyaniline nanotubes (Fig. 10) is very large, about 500 kΩ at room temperature, which is nearly 16 times larger than the intra-tube (intrinsic) resistance of an individual PANI nanotube (about 30 kΩ, Long et al., 2003c). This result explains straightforwardly why an individual polyaniline nanotube has a much higher room-temperature conductivity (30.5 S/cm) than that of a pellet of polyaniline nanotubes where the measured resistance is dominated by the inter-fibril resistance (0.03 S/cm). For crossed PEDOT nanowires, the junction resistance (between the two nanowires) at room temperature can vary from 885-1383 kΩ for one sample and to 370-460 MΩ for another sample, which is respectively comparable or much larger than the intrinsic resistance of the PEDOT nanowires. In addition, the contact resistance shows a stronger temperature dependence (R(72K)/R(300K) is about 120 ~141) and could be fitted by a thermal fluctuation-induced tunneling (FIT) model (Long et al., 2009a). It should be noted that the nanojunction resistance is comparable to the intrinsic resistance of polymer nanotube/wire and shows large sample-sample variations. The possible reasons could be attributed to the contamination of the nanotube/wire surfaces (polycarbonate for template-prepared PEDOT nanowires, solvent impurities or water adsorption), the variation of the junction area between the two nanotubes/ wires, and the self-formation conditions of the junction. It has to be mentioned that no special effort was made to control the formation of the junction between the two crossed nanotubes/wires during fabrication.

Figure 11.

The temperature dependence of the four-probe resistance (R4P) and the two-probe resistance (R2P) of (a) an individual PEDOT nanowire with a diameter of 35 nm, which falls in the metallic regime of the metal-insulator transition, and (b) an individual PEDOT nanotube with a diameter of 190 nm, which falls in the insulating regime of the metal-insulator transion (Long et al., 2010)

The nanocontact resistance between a polymer PEDOT nanowire and a platinum microlead prepared by FIB deposition has also been studied by Long et al. (Long et al., 2010). It was found that the nanocontact resistance (determined from four-probe resistance and two-probe resistance of the same nanowire) is in the magnitude of 10 kΩ at room temperature and can reach 10 MΩ at low temperatures, which, in some cases, is comparable to the intrinsic resistance of the PEDOT nanowires. For a semiconducting polymer nanowire in the insulating regime of the metal-insulator transition, the four-probe resistance is quite close to the two-probe resistance because the contact resistance is much smaller than the intrinsic resistance of the polymer nanowire as shown in Fig. 11a (Long et al., 2008a& 2010). However, for a nanowire that falls in the metallic regime of the metal-insulator transition (for example, the 35 nm PEDOT nanowire as shown in Fig. 11b, Long et al., 2010), the metallic nature of the measured polymer fibers could be over shadowed by the two-probe measurement although the nanowire shows a relatively high electrical conductivity at room temperature (390-450 S/cm). It can be attributed to the nanocontact resistance is much larger than the intrinsic resistance of the nanowire especially at low temperatures. We note that, for individual RuO2 nanowires (Lin et al., 2008), the temperature dependence of the two-probe resistance indicates that the nanowire is semiconducting, whereas the four-probe resistance dependence of the same nanowire shows the measured nanowire is metallic. So, in order to explore the intrinsic electronic transport properties of individual nanowires, especially in the case of metallic nanowires, the four-probe electrical measurement is necessary because nanocontact resistance cannot be excluded in a two-probe measurement.

Advertisement

4. Conclusion

During the past 15 years, significant progress has been made in synthesis, structural and electrical characterizations, and applications of conducting polymer nanotubes/wires. In this chapter, a brief review of the recent advances in electronic transport properties of individual conducting polymer nanotubes/wires prepared by both the template-free self-assembly method and the template method is presented. Results with broad interest have been discussed. For example, it was found that the electrical conductivity of the individual polymer tubes/wires increases by several times of magnitudes with decreasing outer diameter (size effect in electrical conductivity). The crossover from Mott to Efros-Shklovskii variable-range hopping conduction was observed at a relatively high transition temperature in single nanotubes/wires (enhanced Coulomb interaction effect). The low-temperature magnetoresistance of a single polymer tube/wire is positive and quite smaller than those of the nanotube/wire pellets (small magnetoresistance effect). The intrinsic resistance of an individual nanotube is much smaller than the contact resistance of two crossed nanotubes (nanocontact resistance effect). In addition, individual polymer tubes/wires show obvious transition from linear to nonlinear I-V curves at low temperature, and a clear zero-bias anomaly with Coulomb gap-like structure appeared on the differential conductance curves at low temperatures. These results indicate that the electrical properties of isolated conducting polymer tubes/wires are different from those of bulk polymer pellets or films in some cases due to their nanoscale diameters. However, in order to eliminate nanocontact resistance and reveal the intrinsic electronic transport properties of an individual nanotube/ wire, it is still quite important to develop new or improved conductivity measurement approaches on a single nanofibre. Furthermore, due to the complicated microstructures of conducting polymers, there are still problems and challenges to fulfill their applications in nanoscale devices, such as whether fully metallic conducting polymer nanotubes/wires, which show metallic behavior from room-temperature down to low temperatures, can be prepared through improving their molecular or supramolecular ordering. In addition, reproducibility and/or controllability of individual polymer nanotubes/wires are also a problem, since their electrical properties are sensitive to many factors such as doping level, extent of disorder, diameter, temperature, aging effect, etc.

Advertisement

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (Grant Nos.: 10604038, 10374107, 50825206 and 10910101081) and the Program for New Century Excellent Talents in University of China (Grant No.: NCET-07-0472). Part of this work has been supported by the Communauté urbaine de Nantes, France. The authors also wish to acknowledge the support from the National Center for Nano Science and Technology of China.

References

  1. 1. Aleshin A. N. 2006 Polymer nanofibers and nanotubes: Charge transport and device applications. Adv. Mater., 18 1 17 27 , 0935-9648
  2. 2. Aleshin A. N. Lee H. J. Jhang S. H. Kim H. S. Akagi K. Park Y. W. 2005 Coulomb-blockade transport in quasi-one-dimensional polymer nanofibers. Phys. Rev. B, 72 15 153202 (4pp), 1098-0121
  3. 3. Aleshin A. N. Lee H. J. Park Y. W. Akagi K. 2004 One-dimensional transport in polymer nanofibers. Phys. Rev. Lett., 93 19 196601 (4pp), 0031-9007
  4. 4. Cai Z. H. Lei J. T. Liang W. B. Menon V. Martin C. R. 1991 Molecular and super-molecular origins of enhanced electronic conductivity in template-synthesized polyheterocyclic fibrils. 1. Supermolecular effects. Chem. Mater., 3 5 960 967 , 0897-4756
  5. 5. Cai Z. H. Martin C. R. 1989 Electronically conductive polymer fibers with mesoscopic diameters show enhanced electronic conductivities. J. Am. Chem. Soc., 111 4138 4139 , 0002-7863
  6. 6. Callegari V. Gence L. Melinte S. Demoustier-Champagne S. 2009 Electrochemically template-grown multi-segmented gold-conducting polymer nanowires with tunable electronic behavior. Chem. Mater., 21 18 4241 4247 , 0897-4756
  7. 7. Cárdenas J. R. de França M. G. O. de Vasconcelos E. A. de Azevedo W. M. da Silva. Jr E. F. 2007 Growth of sub-micron fibres of pure polyaniline using the electrospinning technique. J. Phys. D.: Appl. Phys., 40 1068 1071 , 0022-3727
  8. 8. Chiou N. R. Epstein A. J. 2005 Polyaniline nanofibers prepared by dilute polymerization. Adv. Mater., 17 1679 1683 , 0935-9648
  9. 9. Cho S. I. Kwon W. J. Choi S. J. Kim P. Park S. A. Kim J. Son S. J. Xiao R. Kim S. H. Lee S. B. 2005 Nanotube-based ultrafast electrochromic display. Adv. Mater., 17 2 171 175 , 0935-9648
  10. 10. Duchet J. Legras R. Demoustier-Champagne S. 1998 Chemical synthesis of polypyrrole: structure-properties relationship. Synth. Metals, 98 113 122 , 0379-6779
  11. 11. Duvail J. L. Dubois S. Demoustier-Champagne S. Long Y. Piraux L. 2008a Physical properties of magnetic metallic nanowires and conjugated polymer nanowires and nanotubes. Int. J. Nanotechnol., 5 Nos. 6/7/8, 838 850 , 1475-7435
  12. 12. Duvail J. L. Long Y. Z. Cuenot S. Chen Z. J. Gu C. Z. 2007 Tuning electrical propeties of conjugated polymer nanowires with the diameter. Appl. Phys. Lett., 90 10 102114 (3pp), 0003-6951
  13. 13. Duvail J. L. Long Y. Rétho P. Louarn G. Dauginet De Pra. L. Demoustier-Champagne S. 2008b Enhanced electroactivity and electrochromism in PEDOT nanowires. Mol. Cryst. Liq. Cryst., 485 835 842 , 1542-1406
  14. 14. Duvail J. L. Rétho P. Fernandez V. Louarn G. Molinié P. Chauvet O. 2004 Effects of the confined synthesis on conjugated polymer transport properties. J. Phys. Chem. B, 108 48 18552 18556 , 1520-6106
  15. 15. Duvail J. L. Rétho P. Garreau S. Louarn G. Godon C. Demoustier-Champagne S. 2002 Transport and vibrational properties of poly(3,4-ethylenedioxythiophene) nanofibers. Synth. Metals, 131 123 128 , 0379-6779
  16. 16. Gence L. Callegari V. Faniel S. Vlad A. Dutu C. Melinte S. Demoustier-Champagne S. Bayot V. 2008 Size related transport mechanisms in hybrid metal-polymer nanowires. Phys. Stat. Sol. (a), 205 6 1447 1450 , 1862-6300
  17. 17. Gence L. Faniel S. Gustin C. Melinte S. Bayot V. Callegari V. Reynes O. Demoustier-Champagne S. 2007 Structural and electrical characterization of hybrid metal-polypyrrole nanowires. Phys. Rev. B, 76 11 115415 (8pp), 1098-0121
  18. 18. Granström M. Inganäs O. 1995 Electrically conductive polymer fibres with mesoscopic diameters: 1. Studies of structure and electrical properties. Polymer, 36 15 2867 2872 , 0032-3861
  19. 19. Heeger A. J. 2002 Semiconducting and metallic polymers: the fourth generation of polymeric materials. Synth. Metals, 125 23 42 , 0379-6779
  20. 20. Huang J. X. Kaner R. B. 2004 A general chemical route to polyaniline nanofibers. J. Am. Chem. Soc., 126 3 851 855 , 0002-7863
  21. 21. Huang J. X. Virji S. Weiller B. H. Kaner R. B. 2003 Polyaniline nanofibers: facile synthesis and chemical sensors. J. Am. Chem. Soc., 125 2 314 315 , 0002-7863
  22. 22. Huang K. Wan M. X. Long Y. Z. Chen Z. J. Wei Y. 2005 Multi-functional polypyrrole nanofibers via a functional dopant-introduced process. Synth. Metals, 155 495 500 , 0379-6779
  23. 23. Huang K. Zhang Y. J. Long Y. Z. Yuan J. H. Han D. X. Wang Z. J. Niu L. Chen Z. J. 2006 Preparation of highly conductive, self-assembled gold/polyaniline nano-cables and polyaniline nanotubes. Chem. Eur. J., 12 5314 5319 , 0947-6539
  24. 24. Jager E. W. H. Smela E. Inganäs O. 2000 Microfabricating conjugated polymer actuators. Science, 290 1540 1545 , 0036-8075
  25. 25. Kaiser A. B. Iorns T. M. Park J. G. Kim B. Lee S. H. Park Y. W. 2002 Variation with temperature of the I-V characteristics of polyacetylene nanofibres. Curr. Appl. Phys., 2 4 285 288 , 1567-1739
  26. 26. Kaiser A. B. Rogers S. A. Park Y. W. 2004a Charge transport in conducting polymers: Polyacetylene nanofiberes. Mol. Cryst. Liq. Cryst., 415 115 124 , 1542-1406
  27. 27. Kaiser A. B. Park J. G. Kim B. Lee S. H. Park Y. W. 2004b Polypyrrole micro-line: current-voltage characteristics and comparison with other conducting polymers. Curr. Appl. Phys., 4 5 497 500 , 1567-1739
  28. 28. Kaiser A. B. Park Y. W. 2003 Comparison of tunnelling conduction in polyacetylene nanofibres, CDW and SDW systems. Synth. Metals, 135 Nos. 1-3, 245 247 , 0379-6779
  29. 29. Kaiser A. B. Park Y. W. 2005 Current-voltage characteristics of conducting polymers and carbon nanotubes. Synth. Metals, 152 181 184 , 0379-6779
  30. 30. Kang N. Hu J. S. Kong W. J. Lu L. Zhang D. L. Pan Z. W. Xie S. S. 2002 Consistent picture of strong electron correlation from magnetoresistance and tunneling conductance measurements in multiwall carbon nanotubes. Phys. Rev. B, 66 24 241403(R) (4pp), 1098-0121
  31. 31. Kim B. H. Park D. H. Joo J. Yu S. G. Lee S. H. 2005 Synthesis, characteristics, and field emission of doped and de-doped polypyrrole, polyaniline, poly(3,4-ethylenedioxy-thiophene) nanotubes and nanowires. Synth. Metals, 150 279 284 , 0379-6779
  32. 32. Kim B. K. Kim Y. H. Won K. Chang H. Choi Y. Kong K. J. Rhyu B. W. Kim J. J. Lee J. O. 2005 Electrical properties of polyaniline nanofibre synthesized with biocatalyst. Nanotechnology, 16 8 1177 1181 , 0957-4484
  33. 33. Kim G. T. Burghard M. Suh D. S. Liu K. Park J. G. Roth S. Park Y. W. 1999 Conductivity and magnetoresistance of polyacetylene fiber network. Synth. Metals, 105 207 210 , 0379-6779
  34. 34. Kozub V. I. Aleshin A. N. Suh D. S. Park Y. W. 2002 Evidence of magnetoresistance for nanojunction-controlled transport in heavily doped polyacetylene. Phys. Rev. B, 65 22 224204 (5pp), 1098-0121
  35. 35. Kuchibhatle S. V. N. T. Karakoti A. S. Bera D. Seal S. 2007 One dimensional nanostructured materials. Prog. Mater. Sci., 52 5 699 913 , 0079-6425
  36. 36. Lee H. J. Jin Z. X. Aleshin A. N. Lee J. Y. Goh M. J. Akagi K. Kim Y. S. Kim D. W. Park Y. W. 2004 Dispersion and current-voltage characteristics of helical polyacetylene single fibers. J. Am. Chem. Soc., 126 51 16722 16723 , 0001-4842
  37. 37. Lin Y. H. Chiu S. P. Lin J. J. 2008 Thermal fluctuation-induced tunneling conduction through metal nanowire contacts. Nanotechnology, 19 36 365201 (7pp), 0957-4484
  38. 38. Liu L. Zhao Y. M. Jia N. Q. Zhou Q. Zhao C. J. Yan M. M. Jiang Z. Y. 2006 Electrochemical fabrication and electronic behavior of polypyrrole nano-fiber array devices. Thin Solid Films, 503 241 245 , 0040-6090
  39. 39. Long Y. Z. Chen Z. J. Wang N. L. Ma Y. J. Zhang Z. Zhang L. J. Wan M. X. 2003a Electrical conductivity of a single conducting polyaniline nanotube. Appl. Phys. Lett., 83 9 1863 1865 , 0003-6951
  40. 40. Long Y. Z. Chen Z. J. Wang N. L. Zhang Z. M. Wan M. X. 2003b Resistivity study of polyaniline doped with protonic acids. Physica B, 325 208 213 , 0921-4526
  41. 41. Long Y. Z. Zhang L. J. Ma Y. J. Chen Z. J. Wang N. L. Zhang Z. Wan M. X. 2003c Electrical conductivity of an individual polyaniline nanotube synthesized by a self-assembly method. Macromol. Rapid Commun., 24 16 938 942 , 1022-1336
  42. 42. Long Y. Z. Chen Z. J. Ma Y. J. Zhang Z. Jin A. Z. Gu C. Z. Zhang L. J. Wei Z. X. Wan M. X. 2004a Electrical conductivity of hollow polyaniline microspheres synthesized by a self-assembly method. Appl. Phys. Lett., 84 12 2205 2207 , 0003-6951
  43. 43. Long Y. Z. Luo J. L. Xu J. Chen Z. J. Zhang L. J. Li J. C. Wan M. X. 2004b Specific heat and magnetic susceptibility of polyaniline nanotubes: inhomogeneous disorder. J. Phys.: Condens. Matter, 16 1123 1130 , 0953-8984
  44. 44. Long Y. Z. Xiao H. M. Chen Z. J. Wan M. X. Jin A. Z. Gu C. Z. 2004c Electrical conductivity of individual polypyrrole microtube. Chin. Phys., 13 11 1918 1921 , 1009-1963
  45. 45. Long Y. Z. Chen Z. J. Wang W. L. Bai F. L. Jin A. Z. Gu C. Z. 2005a Electrical conductivity of single CdS nanowire synthesized by aqueous chemical growth. Appl. Phys. Lett., 86 15 153102 153102 (3pp), 0003-6951
  46. 46. Long Y. Z. Zhang L. J. Chen Z. J. Huang K. Yang Y. S. Xiao H. M. Wan M. X. Jin A. Z. Gu C. Z. 2005b Electronic transport in single polyaniline and polypyrrole microtubes. Phys. Rev. B, 71 16 165412 (7pp), 1098-0121
  47. 47. Long Y. Z. Chen Z. J. Shen J. Y. Zhang Z. M. Zhang L. J. Huang K. Wan M. X. Jin A. Z. Gu C. Z. Duvail J. L. 2006a Magnetoresistance studies of polymer nanotube/wire pellets and single polymer nanotubes/wires. Nanotechnology, 17 24 5903 5911 , 0957-4484
  48. 48. Long Y. Z. Chen Z. J. Shen J. Y. Zhang Z. M. Zhang L. J. Xiao H. M. Wan M. X. Duvail J. L. 2006b Magnetic properties of conducting polymer nanostructures. J. Phys. Chem. B, 110 46 23228 23233 , 1520-6106
  49. 49. Long Y. Z. Huang K. Yuan J. H. Han D. X. Niu L. Chen Z. J. Gu C. Z. Jin A. Z. Duvail J. L. 2006c Electrical conductivity of a single Au/polyaniline microfiber. Appl. Phys. Lett., 88 16 162113 (3pp), 0003-6951
  50. 50. Long Y. Z. Duvail J. L. Chen Z. J. Jin A. Z. Gu C. Z. 2008a Electrical conductivity and current-voltage characterisitics of individual conducting polymer PEDOT nanowires. Chin. Phys. Lett., 25 9 3474-3477, 0025-6307X
  51. 51. Long Y. Z. Wang W. L. Bai F. L. Chen Z. J. Jin A. Z. Gu C. Z. 2008b Current-voltage characteristics of an individual helical CdS nanowire rope. Chin. Phys. B, 17 4 1389 1393 , 1674-1056
  52. 52. Long Y. Z. Yin Z. H. Chen Z. J. Jin A. Z. Gu C. Z. Zhang H. T. Chen X. H. 2008c Low-temperature electronic transport in single K0.27MnO20.5H2O nanowires: enhanced electron-electron interaction. Nanotechnology, 19 21 215708 (5pp), 0957-4484
  53. 53. Long Y. Z. Duvail J. L. Wang Q. T. Li M. M. Gu C. Z. 2009a Electronic transport through crossed conducting polymer nanowires. J. Mater. Res., 24 10 3018 3022 , 0884-2914
  54. 54. Long Y. Z. Yin Z. H. Li M. M. Gu C. Z. Duvail J. L. Jin A. Z. Wan M. X. 2009b Current-voltage characteristics of individual conducting polymer nanotubes and nanowires. Chin. Phys. B, 18 6 2514 2522 , 1674-1056
  55. 55. Long Y. Z. Duvail J. L. Chen Z. J. Jin A. Z. Gu C. Z. 2009c Electrical properties of isolated poly(3,4-ethylenedioxythiophene) nanowires prepared by template synthesis. Polym. Adv. Technol., 20 541 544 , 1042-7147
  56. 56. Long Y. Z. Duvail J. L. Li M. M. Gu C. Z. Liu Z. 2010 Electrical conductivity studies on individual conjugated polymer nanowires: two-probe and four-probe results.. Nanoscale Res. Lett., DOI: 10.1007/s11671-009-9471-y, 1931-7573 1931 7573
  57. 57. Lu G. W. Li C. Shen J. Y. Chen Z. J. Shi G. Q. 2007 Preparation of highly conductive gold-poly(3,4-thylenedioxythiophene) nanocables and their conversion to poly(3,4-ethylenedioxythiophene) nanotubes. J. Phys. Chem. C, 111 16 5926 5931 , 1932-7447
  58. 58. Ma Y. J. Zhang Z. Zhou F. Lu L. Jin A. Z. Gu C. Z. 2005 Hopping conduction in single ZnO nanowires. Nanotechnology, 16 6 746 749 , 0957-4484
  59. 59. Ma Y. J. Zhou F. Lu L. Zhang Z. 2004 Low-temperature transport properties of individual SnO2 nanowires. Solid State Commun., 130 313 316 , 0038-1098
  60. 60. Mac Diarmid. A. G. 2002 Synthetic metals: a novel role for organic polymers. Synth. Metals, 125 11 22 , 0379-6779
  61. 61. Mac Diarmid. A. G. Jones W. E. Jr Norris I. D. Gao J. Johnson A.T. Jr Pinto N.J. Hone J. Han B. Ko F. K. Okuzaki H. Llaguno M. 2001 Electrostatically-generated nanofibers of electronic polymers. Synth. Metals, 119 27 30 , 0379-6779
  62. 62. Martin C. R. 1994 Nanomaterials: A membrane-based synthetic approach. Science, 266 5193 1961 1966 , 0036-8075
  63. 63. Martin C. R. 1995 Template synthesis of electronically conductive polymer nanostructures. Accounts Chem. Res., 28 2 61 68 , 0001-4842
  64. 64. Mativetsky J. M. Datars W. R. 2002 Morphology and electrical properties of template-synthesized polypyrrole nanocylinders. Physica B , 324 191 204 , 0921-4526
  65. 65. Menon R. Yoon C. O. Moses D. Heeger A. J. 1998 In: Handbook of Conducting Polymers, Skotheim, T.A.; Elsenbaumer, R.L. & Reynolds, J.R. (Eds.), 2nd Edition, 85 136 , Marcel Dekker, 0-82470-050-3 York
  66. 66. Orgzall I. Lorenz B. Ting S. T. Hor P. H. Menon V. Martin C. R. Hochheimer H. D. 1996 Thermopower and high-pressure electrical conductivity measurements of template synthesized polypyrrole. Phys. Rev. B, 54 23 16654 16658 , 1098-0121
  67. 67. Rahman A. Sanyal M. K. 2007 Observation of charge density wave characteristics in conducting polymer nanowires: Possibility of Wigner crystallization. Phys. Rev. B, 76 4 045110 (6pp), 1098-0121
  68. 68. Park J. G. Kim B. Lee S. H. Park Y. W. 2003 Current-voltage characteristics of polypyrrole nanotube in both vertical and lateral electrodes configuration. Thin Solid Films, 438-439 , 118 122 , 0040-6090
  69. 69. Park J. G. Kim G. T. Krstic V. Kim B. Lee S. H. Roth S. Burghard M. Park Y. W. 2001 Nanotransport in polyacetylene single fiber: Toward the intrinsic properties. Synth. Metals, 119 53 56 , 0379-6779
  70. 70. Park J. G. Lee S. H. Kim B. Park Y. W. 2002 Electrical resistivity of polypyrrole nanotube measured by conductive scanning probe microscope: The role of contat force. Appl. Phys. Lett., 81 24 4625 4627 , 0003-6951
  71. 71. Park J. H. Yu H. Y. Park J. G. Kim B. Lee S. H. Olofsson L. Persson S. H. M. Park Y. W. 2001 Non-linear I-V characteristics of MEH-PPV patterned on sub-micrometer electrodes. Thin Solid Films, 393 129 131 , 0040-6090
  72. 72. Parthasarathy R. V. Martin C. R. 1994 Template-synthesized polyaniline microtubules. Chem. Mater., 6 10 1627 1632 , 0897-4756
  73. 73. Ramanathan K. Bangar M. A. Yun M. Chen W. Mulchandani A. Myung N. V. 2004 Individually addressable conducting polymer nanowires array. Nano Lett., 4 7 1237 1239 , 1530-6984
  74. 74. Saha S. K. 2002 Room-temperature single-electron tunneling in conducting polypyrrole nanotube. Appl. Phys. Lett., 81 3645 3647 , 0003-6951
  75. 75. Saha S. K. Su Y. K. Lin C. L. Jaw D. W. 2004 Current-voltage characteristics of conducting polypyrrole nanotubes using atomic force microscopy. Nanotechnology, 15 66 69 , 0957-4484
  76. 76. Samitsu S. Shimomura T. Ito K. Fujimori M. Heike S. Hashizume T. 2005 Conductivity measurements of individual poly(3,4-ethylenedioxythiophene)/ poly(styrenesulfonate) nanowires on nanoelectrodes using manipulation with an atomic force microscope. Appl. Phys. Lett., 86 23 233103 (3pp), 0003-6951
  77. 77. Spatz J. P. Lorenz B. Weishaupt K. Hochheimer H. D. Menon V. Parthasarathy R. Martin C. R. Bechtold J. Hor P. H. 1994 Observation of crossover from three- to two-dimensional variable-range hopping in template-synthesized polypyrrole and polyaniline. Phys. Rev. B, 50 20 14888 14892 , 1098-0121
  78. 78. Tan J. S. Long Y. Z. Li M. M. 2008 Preparation of aligned polymer micro/nanofibers by electrospinning. Chin. Phys. Lett., 25 8 3067 3070 , 0025-6307 X
  79. 79. Wan M. X. 2008 A template-free method towards conducting polymer nanostructures. Adv. Mater., 20 2926 2932 , 0935-9648
  80. 80. Wan M. X. 2009 Some issues related to polyaniline micro-/nanostructures. Macromol. Rapid. Commun., 30 963 975 , 1022-1336
  81. 81. Wan M. X. Huang J. Shen Y. Q. 1999 Microtubes of conducting polymers. Synth. Metals, 101 708 711 , 0379-6779
  82. 82. Wan M. X. Liu J. Qiu H. J. Li J. C. Li S. Z. 2001 Template-free synthesized microtubules of conducting polymers. Synth. Metals, 119 71 72 , 0379-6779
  83. 83. Xia Y. N. Yang P. D. Sun Y. G. Wu Y. Y. Mayers B. Gates B. Yin Y. D. Kim F. Yan H. Q. 2003 One-dimensional nanostructures: Synthesis, characterization, and applications. Adv. Mater., 15 5 353 389 , 0935-9648
  84. 84. Yin Z. H. Long Y. Z. Gu C. Z. Wan M. X. Duvail J. L. 2009 Current-voltage characterisitics in individual polypyrrole nanotube, poly(3,4-ethylene-dioxy- thiophene) nanowire, polyaniline nanotube, and CdS nanorope. Nanoscale Res. Lett., 4 1 63 69 , 1931-7573
  85. 85. Yoon C. O. Reghu M. Moses D. Heeger A. J. 1994 Transport near the metal-insulator transition: Polypyrrole doped with PF6. Phys. Rev. B, 49 16 10851 10863 , 1098-0121
  86. 86. Zhang D. H. Wang Y. Y. 2006 Synthesis and applications of one-dimensional nano-structured polyaniline: An overview. Mat. Sci. Eng. B- Solid, 134 1 9 19 , 0921-5107
  87. 87. Zhang L. J. Long Y. Z. Chen Z. J. Wan M. X. 2004 The effect of hydrogen bonding on self-assembled polyaniline nanostructures. Adv. Funct. Mater., 14 7 693 698 , 0161-6301 X
  88. 88. Zhang X. Y. Lee J. S. Lee G. S. Cha D. K. Kim M. J. Yang D. J. Manohar S. K. 2006 Chemical synthesis of PEDOT nanotubes.. Macromolecules, 39 470 472 , 0024-9297

Written By

Yun-Ze Long, Zhaojia Chen, Changzhi Gu, Meixiang Wan, Jean-Luc Duvail, Zongwen Liu and Simon P. Ringer

Published: 01 February 2010