Sunteți pe pagina 1din 17

View Online

FEATURE ARTICLE

www.rsc.org/materials | Journal of Materials Chemistry

The oxidation of aniline to produce polyaniline: a process yielding many different nanoscale structures
Henry D. Tran,*a Julio M. DArcy,b Yue Wang,b Peter J. Beltramo,b Veronica A. Strongb and Richard B. Kaner*b
Received 16th August 2010, Accepted 27th October 2010 DOI: 10.1039/c0jm02699a The number of different nano- and micro-scale structures produced from the chemical oxidation of aniline into polyaniline is rivaled by few other organic materials. Nanoscale structures such as bers, tubes, aligned wires, owers, spheres and hollow spheres, plates, and even those resembling anatomical organs, insects, and sea animals have been observed for the products produced when aniline is oxidized. This feature article examines these different structures and the small and subtle changes in reaction parameters that result in their formation. These changes can often result in drastic differences in the polymers nanoscale morphology. Because a nanomaterials properties are highly dependent on the type of morphology produced, understanding polyanilines propensity for forming these structures is crucial towards tailoring the material for different applications as well as improving its synthetic reproducibility. The different approaches to commonly observed polyaniline nanostructures are presented in this article along with some of the highly debated aspects of these processes. The article ends with our approach towards resolving some of these contentious issues and our perspective on where things are headed in the years to come.

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

1. Introduction
In the eld of nanoscience, there are many inorganic materials known for their ability to adopt a surprising number of different nanoscale sizes and shapes. This is now true of at least one material within the eld of inherently conducting polymers polyaniline. In recent years, variations on the chemical oxidation of aniline have produced polyaniline possessing a wide variety of micro- and nanoscale structures.15 Nanostructured materials have been a rapidly growing eld of research in recent years in part due to the realization that old, well-studied materials can display new and sometimes surprising properties at the nanoscale.6,7 Not only can new properties be introduced into old materials, but existing properties can also become magnied or enhanced when moving towards nanostructured counterparts. As a result, a variety of useful applications have been envisioned and demonstrated with nanostructured materials that are otherwise difcult or impossible to achieve with conventional bulk materials.8 Of particular importance to the eld is understanding the structure/property relationships of nanoparticles since many physical properties can be tuned by adjusting the materials size and/or shape. A classical example in this regard is the color variations observed for quantum dots of varying diameters.7 The degree of control that chemists and material scientists have achieved in manipulating nanoparticle structure has grown rapidly during the past two decades. This is especially
a Fibron Technologies Inc., 419 Hindry Ave Ste E, Inglewood, CA, USA. E-mail: htran@cal.berkeley.edu; henry.tran@brontech.com; Fax: +1 888 280 5715; Tel: +1 310 258 2411 b Department of Chemistry and Biochemistry, California NanoSystems Institute, University of California, Los Angeles, USA. E-mail: kaner@ chem.ucla.edu

true of inorganic semiconducting and metallic nanostructures.6 For example, ZnO nanostructures can be produced in many different sizes and shapes by changing simple reaction parameters.911 These structures can vary from nanowires to a great variety of startlingly complex structures. At the same time, interest in controlling electroactive organic materials, such as conducting polymers or small molecule organic conductors, have grown due to the potential advantages of combining an organic (semi)conductor with low dimensionality.12 But nanoscale size and shape control for these organic materials has been more difcult to achieve than with their inorganic counterparts. Many inorganic materials have well-dened thermodynamically or kinetically stable crystal structures that can be used to control their nanoscale structure; however, many organic materials (especially conducting polymers) lack a dened crystal structure that can be exploited in the same way.12 Instead, organic materials generally rely on self-assembly processes dictated by intermolecular interactions in order to form dened nanostructures. This is particularly true of polyaniline and the many different nanoscale structures observed for this polymer. In general, the processes to produce polyaniline nanostructures can be classied into two categories: those that rely on external templates or additives to direct nanostructural growth and those that do not.1 The former is a process that can involve the use of an insoluble hard-template with pre-formed nanoscale features or a soft-template, typically consisting of a surfactant that can guide the growth of the polymer in or around self-assembled micelles. Routes that do not use external templates typically exploit the predisposition of the material to self-assemble into well-dened morphologies. It is the nature of the self-assembly mechanism for this process that has been debated vigorously in recent years.
This journal is The Royal Society of Chemistry 2011

3534 | J. Mater. Chem., 2011, 21, 35343550

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

In this article, the different types of micro- and nanosized shapes commonly seen when oxidizing aniline into polyaniline will be discussed. The term polyaniline is used loosely to describe all of the products observed when oxidizing aniline into a polymer; however, it is important to note that some of the processes described in this article do not produce polyaniline in the traditional sense, i.e. a polymer consisting of purely head-totail coupled aniline monomers. This article is divided into the different types of morphologies and shapes typically seen. Various mechanisms proposed by different researchers for the production of these structures are briey discussed throughout the article. Our aim is not to provide a comprehensive review of all research activity within this eldthe amount of literature published on this topic is simply too great to discuss here. Rather, the aim is to provide the reader with a brief historical perspective on the eld and a snapshot of current research activity which includes new research done in our laboratory.

2. On the chemical structure of polyaniline


Before discussing the varying micro- and nano-structures of polyaniline, it is instructive to briey review the chemical structure of polyaniline and how its understanding evolved. The aniline monomer was isolated as early as 1826 when crystalline salts of aniline sulfuric and phosphoric acid were observed from the pyrolytic distillation of indigo.13 Although the exact date of the rst reported polyaniline is unclear, aniline was oxidized as early as 183414 and 184015 when pure aniline (observed as a colorless oil) was obtained from indigo and oxidized with chromic acid. For the next several decades, reports on the

Fig. 1 The idealized oxidation states of polyaniline (top). Leucoemeraldine (y 1) is the most reduced oxidation state, pernigraniline (y 0) is the most oxidized state, and emeraldine (y 0.5) is an intermediate oxidation state. The emeraldine base form can be doped with a strong protonic acid to generate the conducting emeraldine salt form of polyaniline (bottom).

products from the oxidation of aniline would sporadically appear in the literature and in the early part of the 20th century, two separate groups would conclude that the oxidation of aniline actually produced an aniline octamer whose color was dependent on the number of quinoimine moieties in the backbone of the octamer.16,17 The terms leucoemeraldine, emeraldine, and pernigraniline were coined during this time to describe the different oxidation states of the aniline octamer and many oxidation/reduction experiments were performed in order to correlate the above names to various oxidation states and colors. Many decades later, it was eventually concluded that the oxidation of aniline actually resulted in a polymer and reinvestigation of polyaniline as a conducting polymer would reinvigorate research into this oldest of electroactive polymers.18,19 For simplicity, the polymer is often ideally depicted as four repeat units in which the leucoemeraldine oxidation state consists of the fully reduced polymer, the emeraldine oxidation state consists of a 3 : 1 ratio of benzenoid and quinoneimine moieties linking the nitrogen atoms of the backbone, and the pernigraniline is the fully oxidized version of the polymer (Fig. 1).20,21 A more accurate and detailed representation requires eight repeat units, which is not considered here, but described in Ref. 20. Further research into the polymer established many of the conventions that are still held today. For example, that the most conductive form of polyaniline is the emeraldine oxidation state when fully protonated with an acid.21,22 The conventional chemical synthesis of polyaniline was also established during this time; mainly, an aqueous solution of ammonium peroxydisulfate (APS) is added slowly to a solution of aniline dissolved in an aqueous solution of acid at $5  C.2022 The reaction is carried out in a strongly acidic environment such as 1 M HCl because low pH promotes the head-to-tail coupling of aniline monomers. After several hours, a green precipitate forms that can be collected and puried to produce proton doped, conducting polyaniline. This material can then be dedoped with the addition of ammonium hydroxide to produce the blue, insulating emeraldine base form of the polymer. Oxidizing aniline in this fashion typically produces polyaniline with higher molecular weight than many other techniques. Furthermore, the process is believed to produce polyaniline with a structure most resembling that of the idealized emeraldine oxidation state; however, obtaining denitive evidence for this remains exceedingly complex.20 Polyaniline produced in this fashion also typically possesses a granular morphology of irregular shaped micron sized particles. Despite its long history, it was not until 1987 that the structure of polyaniline produced by the chemical oxidation of aniline was proved unambiguously by Fred Wudl and coworkers.23 The oxidation of aniline initially produces the radical cation [(PhNH2)+] which can react with another aniline molecule through the para position to produce polyaniline or through the ortho or the meta positions (although with much lower probability than para coupling). Reaction through the ortho position can lead to further oxidation which produces phenazinelike moieties along the backbone of the polymer. Furthermore, aniline reacting through tail-to-tail coupling will produce benzidine moieties or azobenzene through head-to-head coupling (Fig. 2).20,23 These possible reactions of aniline during its initial stages of oxidation are crucial to understanding many
J. Mater. Chem., 2011, 21, 35343550 | 3535

This journal is The Royal Society of Chemistry 2011

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

Fig. 2 A number of different initial products can be produced from the oxidation [O] of aniline which can include head-to-tail coupled aniline units (p-aniline dimer), o-aniline dimer which can subsequently be oxidized to form phenazine-like moieties, or tail-to-tail coupled aniline units to produce benzidine. Which pathway the oxidation of aniline takes is crucially dependent on reaction parameters such as pH, concentration of reagents, reaction temperature, type of oxidant used, and the ratio of oxidant to monomer. These initial products often determine the resulting polymers nal morphology.

of the proposed mechanisms of nanostructure formation since they form the basis for the proposed self-assembly processes (detailed explanations of the mechanism of aniline oxidation can be found in Ref. 2124). Through chemical synthesis, Wudl et al., were able to prove that the chemical oxidation of aniline by the conventional synthesis produces predominantly para-coupled aniline.23 By employing a condensation reaction between p-phenylenediamine and succinoylsuccinic acid, poly(p-phenyleneamineimine) was produced which exhibits many of the same

physical properties as polyaniline made by the direct chemical oxidation of aniline. Unfortunately, despite the fact that this synthesis produced a polymer with solely para-para coupled aniline moieties, the conductivity of the resulting doped polymer was disappointing (103 S/cm), which is at least 3 orders of magnitude lower than what polyaniline is typically reported as possessing. This may be a result of the as-synthesized polymer deviating from the idealized emeraldine oxidation state. This underscores the importance that the oxidation state of polyaniline can have on its conductivity and other physical properties. In fact, differences in reaction conditions or procedures for making polyanilinefrom large ones (such as non-oxidative routes to make polyaniline)23 to seemingly innocuous changes (such as the amount of agitation during polymerization)25can often have a profound effect on the materials oxidation state, particle size, molecular weight, and degree of aggregation. This in turn affects some of the polymers other properties such as color, conductivity, and dispersibility. Care must be taken when evaluating these processes and materials as many of the polyanilines produced from the chemical oxidation of aniline contain a signicant amount of non-para coupled aniline moietieswe believe that this actually plays a signicant role in the large

Fig. 3 A schematic summary of the synthetic pathways for the oxidation of aniline. The center red circle presents the parameters that can be changed in order to oxidize aniline into polyaniline. These variables can be optimized in order to promote either head-to-tail coupling of aniline units or ortho-coupled aniline unitseither of which can serve as the building blocks for nanostructured polyaniline. Besides the core parameters for oxidizing aniline, other reaction parameters and processes can be modied or introduced in order to tailor the type of nanostructure desired (blue). These parameters and processes include the type of template used, chemical additives introduced into the oxidation of aniline, and how the polymerization is carried out.

Fig. 4 Many synthetic parameters can be varied when oxidizing aniline. Changes in these synthetic conditions can produce many different types of nanostructures; a representative sampling is shown here. For example, either the nanober or nanotube morphology can be obtained when oxidizing aniline with varying pH values and/or by changing the addition rate of the reagents. Many other morphologies have been produced when aniline is oxidized and the protocols for obtaining these structures typically involve changing variables such as acid dopant,37,38,4046 oxidant,61 solvent,3 pH,105108 ratio of reagents,4046 and reaction temperature.3a In some cases, additives such as surfactants or aniline oligomers are introduced in order to guide the formation of polyaniline into a specic superstructure.64,68 The variety of nanostructures that can be produced offers a unique opportunity for developing an improved understanding of the evolution of nanostructures in rigid-rod, precipitation polymerization processes.

3536 | J. Mater. Chem., 2011, 21, 35343550

This journal is The Royal Society of Chemistry 2011

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

number of morphologies observed when making polyaniline by chemical oxidation. Therefore it is not surprising that Alan MacDiarmid, one of the pioneers in polyaniline research (2000 Nobel Laureate in Chemistry) stated that .there are as many different types of polyaniline(s) as there are people who make it!26 The following sections will demonstrate that even subtle changes in reaction conditions can have a profound effect on polyanilines size and shape at the nanoscale (Fig. 3). When characterizing polyaniline, not only must the materials oxidation state, molecular weight, and structural regularity be considered, but the polymers morphology must be explored as well since the as-produced morphology can have quite an impact on the materials other physical properties. Therefore, great care must be taken when reproducing results from other research groups as even slight deviations in procedure can produce a polymer with a different morphology. Its the polymers sensitivity to these slight changes in reaction conditions that give rise to polyanilines propensity for forming so many different nanoscale and microscale structures (Fig. 4).

3. Nanobers
The most commonly observed well-dened nanostructured morphology for polyaniline is the nanober (Fig. 5). This morphology was observed as far back as 1986 when a lm of emeraldine hydroperchlorate polyaniline was grown electrochemically on indium tin oxide glass at a constant current of 50 mA cm2 for 90 min.21 The lms are composed of 200 nm diameter compact brils that are reminiscent of polyacetylene nanobers observed years earlier.27 Curiously, similar experiments using hexauoroborate as the counter-anion and

Fig. 5 TEM images of polyaniline nanobers synthesized by interfacial polymerization in different acids (top). Polyaniline synthesized in HCl produces nanobers with average diameters of 30 nm, while CSA and HClO4 produce nanobers with average diameters of 50 and 120 nm. Polyaniline nanobers can form stable aqueous colloids (bottom left) that can subsequently be cast into thin lms (bottom middle). An SEM image of polyaniline nanobers synthesized by rapidly-mixing a solution of oxidant and monomer (bottom right, scale bar 100 nm). Adapted from ref. 3, 25, 55a.

a platinum electrode instead only produced lms comprised of irregularly shaped spheres. Many other electrochemical experiments have conrmed the formation of lms of polyaniline nanobers under specic growth conditions,28 although the differences in morphology observed when using different electrodes or counter-anions remains under investigation.28b Electrochemistry is a useful technique to study the formation mechanism of polyaniline nanobers because the redox potential of the system can be controlled precisely, and in some cases other morphologies can be produced as well.28c,d These studies have revealed that the dendritic degree of polyaniline nanobers is related to the dopant used during oxidation. Furthermore, the process can control the diameter of the nanobers produced by changing the sweep rate during polymerization.29 Aside from physical routes used to produce polyaniline nanobers such as electrospinning3032 or mechanical stretching,33 polyaniline nanobers are often synthesized by chemical oxidation. Soft-templates such as surfactants can self-assemble into mesophase structures that induce the formation of nanobers. The process has been referred to in the literature by a number of different names such as the template-free method or the self-assembly method.34 We should note that the term template-free refers to the generation of these structures without the use of traditional hard-templates such as anodic aluminium oxide (AAO), which was originally used to produce polyaniline nanostructures in the 1990s, and is still occasionally employed today.35,36 Much of the initial pioneering work on generating polyaniline nanostructures actually involved oxidizing aniline inside the pores of these hard templates which allowed for the production of aligned wires or nanotubes of polyaniline.35 Additionally, dissolving away the hard template allows one to obtain the pure polymer nanostructure. The use of surfactants is quite versatile and have not only been used to produce polyaniline nanobers, but other conducting polymers as well.34 Surfactants, such as cetyltrimethylammonium bromide (CTAB)37 or hexadecyltrimethylammonium (C16TMA),38 can self-assemble into transient micellular structures when their concentration reaches a critical value. One-dimensional cylindrical micelles can form by achieving a second critical micellular concentration under appropriate conditions.39 These one-dimensional nanoscale structures can subsequently serve as a scaffold for aniline to grow off of during polymerization which results in nanobers. The process is straightforward and a wide variety of softtemplates and surfactants have been used to generate polyaniline nanobers including: acids with bulky organic side groups such as camphorsulfonic acid,40 saturated fatty acids,41 naphthalenesulfonic acids and their derivatives,40,42 toluenesulfonic acid,40 azobenzenesulfonic acid,43 p-aminobenzenesulfonic acid (ABSA),36 2-acrylamido-2-methyl-propane-1-sulfonic acid (AMPSA),44 sodium dodecylbenzenesulfonate (SDBS),45 and renewable source surfactants like 4-[4-hydroxy-2((Z)-pentadec8-enyl phenylazo]-benzenesulfonic acid.46 Note that many of the surfactants and dopant acids listed above have also been reported to produce nanotubes of polyaniline, but we also list them here because a mixture of nanotubes and nanobers are occasionally observed (further discussion on nanotubes is presented in a later section). In a typical synthetic procedure to produce polyaniline nanobers using this process, aniline and
J. Mater. Chem., 2011, 21, 35343550 | 3537

This journal is The Royal Society of Chemistry 2011

View Online

Fig. 6 A schematic illustration of the different modes of nucleation and template aggregation that can lead to nanostructures for polyaniline. Nanostructures can form at a liquid/liquid interface due to suppression of secondary growth as polyaniline moves into aqueous phase or by interfacially trapped anilinium cations that can act as nuclei for nanober growth.46,55 Aniline and anilinium cations can also aggregate into micellular structures because they are amphiphiles that contain a hydrophilic and hydrophobic component.4046 Formation of polyaniline nanostructures can also be promoted by interchain pi-pi stacking64 of oligomeric segments, other intermolecular forces such as hydrogen bonding, or by aggregation of aniline oligomers or phenazine moieties into tubular micelles.105108

the surfactant or bulky dopant acid are dissolved in water in varying concentrations. This monomer/surfactant solution can be cooled and a solution of oxidant (typically APS) is added. After a period ranging from several hours to as much as one day, the crude product is collected and puried in order to obtain polyaniline nanobers. Experimental studies indicate that producing the nanobrillar morphology is dependent on several factors including monomer/oxidant concentration and ratio, temperature, solvent, and surfactant concentration and ratio.4046 The acidic surfactant is believed to play a dual role: guiding the nanobrillar morphology of the polymer and serving as the anion dopant in order to produce polyaniline in its conductive state. The proposed mechanism centers on the belief that these dopants and/or their salt complexes with the monomer/oligomer form micelles that guide the growth of polyaniline into anisotropic structures.34 By varying reaction conditions such as the dopant used or concentration of reagents, different micellular structures can form which results in a number of different polymer structures. Interestingly, changing the amount of acid and/or the strength of the acid introduced during the oxidation of aniline appears to be a key parameter in shifting from one morphology to another.4045 Besides the aggregation of the dopants or their salt complexes, aniline monomer/oligomers themselves have been suggested as molecular templates for guiding the growth of polyaniline nanobers (Fig. 6).47
3538 | J. Mater. Chem., 2011, 21, 35343550

Along with soft-templates that self-assemble into dened micellular structures, materials with pre-existing structures can also guide the growth of polyaniline into a nanobrillar morphology. Materials used in this fashion have ranged from carbon nanotubes,48 polyelectrolytes,49 cyclodextrins,50 and porphyrins,51 to other bio-inspired templates.52 Many of these processes presumably involve the absorption of aniline monomers/oligomers onto the predened nanostructure which, in turn, guides the nanobrillar growth of polyaniline. This technique opens up the possibility of producing a variety of different polyaniline nanostructures based on the template. Indeed, this has been demonstrated and is discussed in subsequent sections. One such material with a pre-existing morphology used in this fashion is V2O5 nanobers.53 In this growth process, V2O5 nanobers can essentially transfer their morphology to polyaniline when aniline is polymerized in their presence. The procedure involves the addition of a small amount of V2O5 nanobers into the oxidation of aniline. Pure polyaniline nanobers can be obtained by washing the crude product with an aqueous acid solution which removes any residual V2O5 nanobers. Studies regarding the mechanism of this process reveal that the template must be capable of oxidatively reacting with the monomer in order for brillar growth to occur. Aniline is believed to undergo a prepolymerization reaction on the surface of the V2O5 nanobers that then facilitates the evolution of a bulk nanobrillar morphology for polyaniline when an oxidant such as APS is added. The nanobrillar morphology of the V2O5 seed is crucial as reactions seeded with granular V2O5 only produce a granular, agglomerated polymer.53 Interestingly, the process is not only applicable for the production of polyaniline nanobers, but also other conducting polymer nanobers as well.54 Polyaniline nanobers can also be synthesized by interfacial polymerization (Fig. 7).55,56 The process is analogous to the synthesis of Nylon and is performed in an immiscible organic/ aqueous biphasic system. In a typical reaction, aniline is dissolved in a water immiscible solvent such as chloroform or toluene. Meanwhile, in a separate container, an oxidant such as APS is dissolved in an aqueous solution containing a strong acid such as HCl or HClO4. The two solutions are then carefully transferred to a reaction vessel and the two immiscible solvents phase segregate generating an interface between the two layers.55,56 The oxidation of aniline occurs where the two liquids meet; within several minutes polyaniline nanobers form at the interface and the polymer then migrates into the aqueous phase. Following purication, the polymers morphology consists of an entangled mat of nanobers with a tunable diameter between 30120 nm depending on the acid used during polymerization.55 An all-aqueous process can also be used to generate polyaniline nanobers by rapidly-mixing an acidic solution of aniline with an acidic solution of APS.5759 This can be carried out either at dilute concentrations59 or at standard concentrations for the oxidation of aniline.57 Polyaniline nanobers produced in this manner appear very similar to nanobers made by other methods and, like the interfacial synthesis of polyaniline nanobers, can be tuned to produce different diameter nanobers by simply varying the acid used during polymerization. The process differs from the conventional synthesis of polyaniline since the reagents are rapidly mixed instead of slowly dripping oxidant into the monomer solution. Furthermore, the reaction is generally carried
This journal is The Royal Society of Chemistry 2011

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

Fig. 7 Snapshots showing the interfacial polymerization of aniline. The reaction times are (a) 0 s, (b) 60 s, (c) 90 s, (d) 120 s, and (e) 180 s. The top layer contains an aqueous acidic solution of oxidant, while the bottom layer contains aniline dissolved in an organic solvent such as chloroform which is immiscible with water. As the reaction proceeds, polyaniline, which is hydrophilic, diffuses away from the reactive interface and into the aqueous layer. Adapted from ref. 55.

out at room temperature as opposed to 05  C.3a Other critical parameters for producing polyaniline nanobers in this fashion include the solvent polarity,3a reaction concentration,60 and type of oxidant/dopant.61 In fact, a large number of different oxidants and mixed oxidants have been explored for oxidizing aniline and the effects of these oxidants can, in some instances, be quite dramatic.5 In some cases these oxidants affect the diameter of the nanobers.61b The spontaneous formation of polyaniline nanobers under these simple conditions has elicited a number of varying theories. Our own original proposed mechanism focuses on the even distribution of aniline and APS molecules throughout the solution.3,6264 Since all the APS is rapidly consumed at the onset of the polymerization, secondary growth of polyaniline particles is suppressed which results in the production of nanobers (the

Fig. 8 A schematic illustration showing how nucleation can affect the aggregation of polyaniline nanobers during synthesis. During the initial stages of polymerization, nanobers form exclusively. As the reaction proceeds, two different pathways are possible: I) heterogeneous nucleation occurs in which pre-existing polyaniline nanobers serve as a scaffold for the growth of new polyaniline particles. This leads to particle aggregation and causes polyaniline to settle out of solution. SEM image of polyaniline produced under this pathway reveal large, micron sized agglomerates. II) homogeneous nucleation continues to occur in which new nuclei form from solution. This leads to well-dispersed polyaniline nanobers. SEM image of polyaniline produced under this pathway reveal a mat of interconnected nanobers. Adapted from ref. 63.

intrinsic morphology of polyaniline).3 This can also be explained using classical nucleation theory. Because polyaniline is a water insoluble polymer, its growth is accompanied by a precipitation event; i.e. a nucleation event in which liquid turns into solid. Based on a number of experimental observations, it has been suggested that the shape of polyaniline can be correlated to its type of nucleation: homogeneous nucleation of polyaniline results in nanobers, while heterogeneous nucleation leads to granular particulates (Fig. 8).62,63 Conditions that favor homogeneous nucleation such as rapid mixing of oxidant and monomer solution or having the reaction performed at an elevated temperature can thus lead to the production of high-quality nanobers. Interestingly, a recent study has suggested that polyaniline nanobers may form as a result of a double heterogeneous nucleation process in which nascent nanobers form on available surfaces within the reaction medium and that morphology is rapidly transcribed across multiple length scales to the bulk precipitate.65 Why nanobers appear to be the intrinsic morphology of polyaniline remains unanswered. Clearly, a self-assembly process must occurit is the nature of the self-assembly that has been debated. A variety of explanations have been proposed to explain the molecular basis for this phenomenon including the aggregation of aniline dimers/oligomers into micelles,47 the aggregation of aniline tetramers,66 the aggregation of anilinium cations into micelles,43 and the role of oxidant/oligomer interactions in guiding nanobrillar growth.67 Furthermore, the dopant-micelle model proposed for the formation of polyaniline nanobers when polymerizing aniline in the presence of bulky dopant acids is also relevant here as well.34 In trying to relate the molecular basis of nanober formation to nucleation theory, it has also been suggested that the anisotropic, rigid-rod nature of polyaniline necessitates the formation of an anisotropic cluster of oligomers due to surface energy minimization requirements when nucleating homogeneously which then leads to the nanobrillar morphology.64 Similar phenomena have been observed for other polymeric systems as well.68 This process can be promoted by the addition of oligomers into the polymerization of aniline which has been observed to not only produce high-aspect ratio nanobers of polyaniline64,69 and chiral polyaniline nanobers,71 but nanobers of other conducting polymers as well.64,70 We should
J. Mater. Chem., 2011, 21, 35343550 | 3539

This journal is The Royal Society of Chemistry 2011

View Online

Table 1 The weight average molecular weight (Mw), the number average molecular weight (Mn), and the polydispersity index (PDI Mw/Mn) of polyaniline and select substituted polyaniline nanobers measured by the procedure described by Mattes et al.73 The substituted polyaniline nanobers are all synthesized with the addition of a small amount of p-aniline dimer or p-phenylenediamine into the polymerization. In general, the molecular weights and PDI are signicantly lower than what is typically reported for non-nano forms of these polymers Polymer Polyethylaniline Nanobers Poly-N-ethylaniline Nanobers Polyanisidine Nanobers Polychloroaniline Nanobers Polyaniline Nanobers by interfacial polymerization Polyaniline Nanobers synthesized by the addition of aniline dimer Mw 7,800 5,300 14,400 7,100 26,000 8,200 Mn 4,200 2,900 6,300 2,800 9,200 5,300 PDI 1.7 1.8 2.3 2.5 2.8 1.5

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

note that the addition of oligomers in this process to promote the reaction is actually a homogeneous process because the solvent system that is selected for oxidation clearly dissolves the solid low-molecular weight oligomers. Furthermore, we have observed the formation of high quality polyaniline nanobers when the reaction is promoted using aniline oligomers that are actually incapable of forming the aniline dimer or tetramer intermediates such as aniline trimer or pentamer. Studying the detailed superstructure and chemical morphology of polyaniline nanobers also provides a useful pathway for potentially elucidating the formation mechanism of these materials. Recently, a number of papers have reported differences when comparing the molecular structure of polyaniline possessing no structural features to those possessing a nanobrillar morphology.72 While many of the differences are subtle, the most obvious difference may be the differences observed in molecular weight and polydispersity (Table 1).64 When comparing the molecular weights of polyaniline nanobers produced by interfacial polymerization or by rapidly-mixing monomer and oxidant to the molecular weights of conventional polyaniline, the nanober form generally possesses both a much lower molecular weight and a lower polydispersity. We should note that care must be taken when comparing the molecular weight of polyaniline from different studies as the results obtained are quite sensitive to how the experiment is carried out. In fact, much of the literature compares the molecular weight of polyaniline measured under different experimental conditions. We do not believe valid comparisons can be made with this data so as a basis for standardization, we have chosen to measure the molecular weights of all our materials by the process described by Mattes et al.73 NMR studies of polyaniline nanobers and of conventional polyaniline (13C and 15N) indicates spectra with slight differences in structure.72a,b In general, 15N NMR reveals that there is additional chemistry that occurs in the polymerization process for generating polyaniline nanobers which may include some cross-linking between polymer chains.72b 13C NMR of conventional and nanobrillar polyaniline are usually similar although the presence of some as yet undetermined peaks has been observed in a recent study.72d Neutron diffraction, wide angle X-ray diffraction (WAXS), and small angle X-ray scattering (SAXS) reveal differences
3540 | J. Mater. Chem., 2011, 21, 35343550

between conventional and nanobrillar polyaniline in solid-state packing, molecular conformation, and supramolecular packing in the emeraldine base form. Under conditions that produce nanobers, some crystallinity is generally observed for polyaniline, while under conventional conditions the as-synthesized polymer is essentially amorphous. Specically, a low angle peak using SAXS shows a large scattering intensity at Q 0.036 nm1 for the nanobrillar material implying an intercrystal spacing of d 1.3 nm that has been assigned to the 001 crystal plane. This peak is thought to be a reection of periodic alteration of the semi-crystalline lamellae and amorphous regions within the nanobers.72a A number of processes have been developed for making polyaniline nanobers without the use of non-native chemical additives, but with the assistance of an external energy input. These include radiolytic,74 ultrasonic and sonochemical,75 and microwave assisted76 syntheses of polyaniline nanobers. For example, by subjecting the polymerization bath of a conventional polyaniline synthesis to ultrasonic radiation, nanobers of polyaniline are formed instead of the usually observed micronsized agglomerates.75 The process may accelerate the rate of polymerization of aniline which has been shown to be benecial for the formation of polyaniline nanobers synthesized by other means.75b Besides solution based methods for producing polyaniline nanobers, a process starting from solid precursors without any solvents was recently discovered for generating these structures.77 In this process, solid aniline salts such as aniline hydrochloride

Fig. 9 A reaction scheme showing the oxidation of a substituted aniline. When substituted anilines are polymerized under conditions that typically produce polyaniline nanobers such as rapid-mixing,57 interfacial,55 or dilute polymerization,59 only irregular shaped agglomerates are typically observed. However, when a small amount of an additive such as p-phenylenediamine or p-aniline dimer is predissolved and added into a similar reaction process, the resulting polymers morphology changes from agglomerates to nanobers.78,79 Only a small amount of additive is necessary in order to induce this change. Nanobers produced in this fashion include alkyl, alkoxy, halogenated, alkyl thiol, and sulfonated/ self-doped substituted polyanilines.

This journal is The Royal Society of Chemistry 2011

View Online

are ground together for up to several hours with an oxidant such as FeCl3. The low melting point and hydroscopic nature of FeCl3 causes the formation of a slurry after several minutes of grinding the oxidant with the aniline salts. Nanobers produced in this manner have a similar nanostructure to those made using traditional solution based methods.77 While the chemical oxidation of aniline appears to produce a polymer that readily forms the nanobrillar morphology, nanobers of substituted polyanilines are much less commonly observed. In most cases, when the techniques used to make polyaniline nanobers are applied to the synthesis of substituted polyanilines, only agglomerated or large spherical structures are observed.3a This is especially true of techniques to produce polyaniline nanobers that do not rely on surfactants or other non-native additives. Recently, it was reported that nanobers of virtually any substituted polyaniline can be produced by introducing amino substituted monomeric or oligomeric additives into the oxidation of substituted anilines.78,79 For example, when ethylaniline is dissolved in an aqueous solution of acid and rapidly mixed with an aqueous acidic solution of APS, only agglomerated, micron-sized particles are observed. However, when the same reaction is performed in the presence of 12 mol% (with respect to the monomer concentration) of predissolved p-aniline dimer or p-phenylenediamine, nanobers of polyethylaniline are observed.79 Only a very small amount of oligomer or monomeric additive is typically required in order to change the morphology of the entire product from agglomerates to nanobers. A wide variety of substituted polyaniline nanobers have been reported using this general procedure including alkylated, sulfonated, halogenated, and alkoxylated polyanilines (Fig. 9).79 In addition to the oligomer assisted synthesis of substituted polyaniline nanobers, these polymer nanobers can also be produced using traditional soft templates,80 gold oxidants81 or by polymerizing certain substituted anilines in aliphatic alcohols.82 For example, self-doped polyaniline nanobers can be synthesized by copolymerizing aniline and m-aminobenzenesulfonic acid using hexadecyl-trimethylammonium bromide (CTAB) as a structural directing agent and APS as an oxidant.80 Nanobers of 6070 nm in diameter can be produced using this method and the diameter of the nanobers can be adjusted by varying the concentration of CTAB. While substituted polyanilines do not typically possess an electrical conductivity as high as the parent polymer, they often have other advantages relative to polyaniline such as increased disperability and improved stability with certain types of substituents.83

are more like molecular wires when compared to other nanostructures presented in this report which can be described generically as bundles of polymer chains entangled together to form a superstructure. In this process, the aniline monomers are believed to absorb and accumulate onto DNA strands and polymerize off of their surfaces. Reports on true molecular wires of polyaniline are, in fact, quite rare52,85 and this process remains one of the few techniques capable of generating these structures. In addition to nucleating from the surfaces of pre-existing nanostructures, discrete nanowires and related one-dimensional morphologies of polyaniline have been observed under a number of different synthetic conditions. For example, polymerizing aniline in the presence of a hydrogelator such as cholic acidpolyethylene glycol (CA-PEG400), results in discrete, nonentangled nanosticks which are 300800 nm long and range from 5070 nm in diameter.86 In this case, the framework of the hydrogel is believed to guide the growth of polyaniline into isolated nanostructures. Besides these template approaches, isolated nanowires have also been observed when aniline is polymerized in the presence of UV light illumination at depressed temperatures87 and when aniline is polymerized in the presence of a small amount of p-aniline dimer.64 In both cases, it is believed that these synthetic conditions promote homogeneous nucleation which minimizes the number of nucleation sites giving the 1-D nanostructures the opportunity to grow longer. In addition to these bulk processes for generating nonentangled 1-D nanostructures, a number of methods have been developed in order to produce single nanowires. This is often an advantage because it allows for the study and testing of a single wire as opposed to a collection of nanowires, thereby enabling the true dimensional diffusion proles of these structures to be observed and utilized which, in turn, provides unprecedented temporal and spatial resolution.88 For example, single nanowires of polyaniline can be generated by electrospinning33 or by direct drawing from solvated polymers.89 A variety of single wire devices have been produced from these materials which often exhibit exceptional properties such as parts per trillion detection of chemical gases.90

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

5. Aligned/oriented 1-D nanostructures


Aligned/oriented arrays of 1-D polyaniline nanostructures have recently received attention because the properties of nanomaterials are not only dependent on their size and shape, but also on their arrangement over macroscopic areas.1 Despite their potential importance, relatively few examples currently exist of producing oriented/aligned arrays of polyaniline nanowires. Besides the use of hard-templates,35 oriented arrays of polyaniline nanobers can be produced without the aid of templates on solid surfaces by a number of different techniques. For example, by carefully controlling the nucleation and growth of polyaniline on a solid surface, oriented arrays of nanostructured polyaniline can be produced (Fig. 10).91,92 Polyaniline typically nucleates heterogeneously on solid surfaces before nucleating in the bulk solution because of the surface energy minimization of nucleating off of foreign surfaces. After a short induction period in which polyaniline grows on solid surfaces, polyaniline nanobers begin forming in the bulk solution, thereby consuming reactive intermediates necessary for
J. Mater. Chem., 2011, 21, 35343550 | 3541

4. Discrete one-dimensional nanostructures


Polyaniline nanowires and other discrete one-dimensional (1-D) nanostructures can be produced by a number of different methods. Here we distinguish the terms nanowire from nanober: nanowires possess a non-entangled, 1-D nanostructured morphology, while nanobers refer to an entangled 1-D morphology. The synthesis of polyaniline nanowires has been reported by a number of different templated methods such as using peptide self-assembly84 or DNA.52 DNA templates can serve as scaffolds for the in situ polymerization of aniline into nanowire structures. In this case, the nanostructures produced
This journal is The Royal Society of Chemistry 2011

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

Fig. 10 An illustration representing various strategies for creating aligned 1-D polyaniline nanostructures. These strategies can include the dilute polymerization of aniline onto a substrate that results in a dense array of aligned polyaniline nanobers,92 physical techniques such as mechanical stretching,33 electroless deposition on patterned platinum substrates,94 alignment using the dewetting technique,150 or nanoimprint lithography.98

continued polyaniline growth on the solid substrate. It is believed that the suppression of polyaniline growth on the solid substrate allows polyaniline to grow vertically from the solid substrate into the bulk solution which results in the production of oriented arrays of 1-D nanostructured polyaniline.92 In addition to this nucleation controlled process, aligned 1-D nanostructures of polyaniline can also be fabricated off of a passivated gold surface modied with 4-aminothiophenol93 or by exploiting the ability of Pt to spontaneously oxidize aniline in the presence of oxygen to produce polyaniline, known as electroless polymerization.94

Fig. 11 SEM images of aligned bundles of polyanisidine nanowires (left) produced by oxidizing anisidine in the presence of an additive such as p-aniline dimer for an extended reaction period and polychloroaniline nanotubes produced by oxidizing 2-chloroaniline with APS in the presence of a small amount of o-phenylenediamine (right).

Aligned nanowires of polyaniline95 and of polymethoxyaniline,96 can also be produced as free-standing structures without the need for nucleating off of a surface. For polyaniline, by oxidizing aniline at elevated temperature and subsequently controlling crystal growth at 0  C, large polyaniline arrays can be produced (Fig. 11).95 This process presumably occurs because at depressed temperatures unreacted salts precipitate from the polymerization solution thereby acting as templates to guide the growth of polyaniline arrays. For polyanisidine, controlling the intermolecular interactions between polymer chains (through careful addition of organic additives that are subsequently incorporated into the polymer themselves) can result in large, free-standing arrays of polyanisidine nanowires.96 For example, by rapidly-mixing a solution of anisidine and o-phenylenediamine with an oxidant such as APS and allowing the reaction to proceed for several days, aligned bundles of polyanisidine nanowires are produced. Through a variety of spectroscopy and control experiments with different substituent groups, the ability of the side groups to hydrogen bond was determined to be a key parameter in generating aligned bundles. This technique relies on the ability of polyanisidine to self-assembly. Other self-assembly techniques that leverage the inherent nature of the polymer and soft-templates to generate aligned polyaniline nanowires, have also been reported.97 A variety of lithography techniques to produce aligned/ oriented arrays of polyaniline nanowires/nanostructures have been explored. A recent example of producing oriented arrays of
This journal is The Royal Society of Chemistry 2011

3542 | J. Mater. Chem., 2011, 21, 35343550

View Online

polyaniline nanowires made using lithography involves a combination lithography/copolymerization/lift-off strategy.98 In this process, a polyaniline copolymer was developed with silane groups in order to improve the adhesion of the material to the underlying surface. Using standard nanoimprint lithography techniques to create a structured pattern, the adhesive-enhanced polyaniline is deposited onto the structured surface. In the grooves of the pattern, the silane groups of the polymer reacts with the bare substrate and adhere, whereas the polymer deposited on the resist layer is removed during the lift-off process leaving behind the desired pattern on the substrate.98
Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

6. Nanotubes
Nanotubes of many different types of materials can be used in a variety of applications.1 Much of the original work on synthesizing polyaniline nanotubes involved oxidizing aniline inside a hard template such as porous alumina.99 However, in recent years, much of the synthetic efforts devoted to making polyaniline nanotubes have instead focused on using soft templates or not using any templates at all. For example, soft templates utilizing functional dopants such as b-naphthalenesulfonic acid readily generates polyaniline nanotubes.100102 In this self-assembly process, a mechanism involving the aggregation of the aniline/ dopant into mesostructures that dictate the growth of polyaniline into nanotubes has been proposed. The dopant is believed to serve the dual role of dopant and surfactant. The morphology of the resultant polyaniline is highly dependent on the ratio of dopant/ monomer, e.g. when the ratio is high ($2/1) akes and nanotubes with diameters of up to $650 nm are often observed.42b,102 As the ratio is decreased to 1/2, the average diameter of the tubes decreases to $90 nm while at 1/4, solid nanobers are produced with an average diameter of 7080 nm. A variety of naphthalene sulfonic acids have been employed in this process to form nanotubes. In addition, similar morphologies are produced by acids such as HCl, H2SO4, HBF4, and H3PO4 when polymerized in the presence of sodium dodecylbenzenesulfonate (SDBS).102c As with naphthalenesulfonic acids, the morphology is dependent on the ratio of monomer/dopant during the polymerization. For example, the average diameter of polyaniline-H3PO4 decreases slightly from 180 nm to 140 nm when the ratio of aniline/H3PO4 is changed from 1 : 0.5 to 1 : 0.01. Interestingly, a common characteristic of these processes is that the pH of the initial solution is typically higher than what is commonly employed for the oxidation of aniline. Besides pure polyaniline nanotubes, these same structures can form decorated with metal nanoparticles by using an analogous approach in which HAuCl4 is included in the polymerization.103 Nanotubes of polyaniline can also be produced using more traditional soft-templates such as n-dodecylbenzenesulfonic acid in an emulsion polymerization.104 Nanotubes are a morphology for polyaniline commonly observed when aniline is oxidized under conditions of high or constant pH.105 In the traditional synthesis of polyaniline, as well as many syntheses for making nanostructured polyaniline, aniline is oxidized in a strongly acidic aqueous environment in order to promote the preferential head-to-tail coupling of aniline. This results in a polymer with less defects and typically a higher electrical conductivity. However, when aniline is oxidized under
This journal is The Royal Society of Chemistry 2011

aqueous conditions of higher pH, ortho-coupled aniline is initially formed, which subsequently oxidizes into phenazine-like moieties. The phenazine units are believed to aggregate into a template-like structure that promotes nanotube growth.106 In this process the phenazinium units act as discotics that guide the stacking of aniline oligomers as they are produced during the oxidation of aniline. This provides the necessary organization and order that predetermines the growth of nanotubes. As the reaction proceeds, the pH of the solution slowly drops due to sulfuric acid evolution from the reaction of APS. The polymerization of aniline can thus proceed as normal once the pH reaches a sufciently low level. The nal nanotube morphology results from growth extending outward from the self-organized phenazine units (Fig. 11).105108 The presence of phenazine moieties in polyaniline nanotubes has been observed using various spectroscopic techniques and appear to be a common structural feature of polyaniline produced by this procedure as a number of different research groups have made similar observations.109 Surprisingly, recent NMR studies of the product from the initial stages of aniline oxidation in water do not reveal the presence of phenazine moieties, but rather oxygen containing quinoneimine structures.110 Similar results have been observed when oxidizing aniline in higher pH conditions as reported by MacDiarmid, et al.111 The failure to detect phenazine moieties may be caused by local immobilization of these structural units (thus causing extreme broadening of the corresponding signals), but the presence of oxygen containing quinoneimine structures adds a further potential layer of complexity towards understanding the formation mechanism of these structures. Interestingly, halogenated phenazine groups themselves can self-assemble into nanotube-like structures which suggest that the proposed mechanism centered on phenazine units is highly plausible.112 Polyaniline nanotubes produced by the falling pH method have outer diameters of 100200 nm with wall thicknesses of 50100 nm and lengths extending up to several microns.105108 The room temperature conductivity of polyaniline synthesized in this fashion is modest compared to traditional polyaniline due to the presence of non-conjugated oligomeric segments. In addition to making pure polyaniline nanotubes, these structures can also be generated in the presence of silica113 and zeolites114 thereby creating composite structures. Polyaniline nanotube/metal composites have also been produced by oxidizing aniline in a high pH solution with the subsequent addition of AgNO3.115 Our own studies have shown that nanotubes/wires of substituted polyanilines, such as polybromoaniline or polychloroaniline, can be produced by intentionally introducing a small amount of additives, such as o-phenylenediamine, into the oxidation of bromoaniline or chloroaniline. In a typical procedure, a halogenated aniline and 12 mol% of o-phenylenediamine is dissolved in an aqueous HCl solution containing a small amount of methanol to predissolve the diamine. In a separate solution, APS is dissolved in an HCl solution and the solution containing monomer and oxidant are rapidly mixed. After a one day reaction time, the puried product is composed of polychloroaniline nanotubes (Fig. 11). Although the mechanism is still under investigation, we believe it is analogous to the mechanism proposed by Stejskal et al.105108 and the assembly of halogenated phenazine moieties.112 In our process,
J. Mater. Chem., 2011, 21, 35343550 | 3543

View Online

ortho-derivatized substituted anilines (such as o-phenylenediamine) are intentionally introduced into the polymerization of the halogenated aniline in order to induce nanotube formation; however, unlike the falling pH method, the procedure is initiated in a strongly acidic environment. Addition of small amounts of o-phenylenediamine induces the formation of phenazine-like moieties that can aggregate into structures that dictate nanotube formationthis subsequently serves as the platform from which halogenated aniline can grow into polyhalogenated nanotubes. Note that we have observed that adding ortho-coupled additives induces the formation of nanotubes in this case; however, the addition of para-coupled additives induces the formation of nanobers. This process may thus be an easy way to convert between the nanober and nanotube morphology for polyaniline and substituted polyanilines. Nanotubes of polyaniline can also be produced by polymerizing from the surfaces of pre-existing templates such as electrospun poly(L-lactide) nanobers which can subsequently be removed by heating,116 thin glass tube templates,117 methyl orange brils,118 halloysite nanotubes,119 silica nanotubes,120 or templating off of MnO2 nanowires.121 In the case of using MnO2 nanowires, these materials serve as both the oxidant and as a template providing the scaffold from which aniline grows off of. The oxidation potential of MnO2 is sufcient to initiate the polymerization of aniline and as the polymerization proceeds, lms of polyaniline form on the surface of the MnO2 nanowires. The resulting polyaniline nanotubes have an external size and shape that are similar in dimension to that of the underlying MnO2 nanowires. By simply changing the morphology of the MnO2 template, different nanosizes and shapes of polyaniline should be possible. This is a simple, one-step procedure because the reactive MnO2 template is converted to soluble Mn+2 ions during the polymerization process. As a result, no special purication steps are required in order to obtain the pure polymer.121 Biological templates such as the tobacco mosaic virus (TMV) have also been used as a pre-existing scaffold in order to grow polyaniline nanotubes.122 Native TMV particles possess a tubelike structure that measures 300 nm in length and 18 nm in diameter with a 4 nm cylindrical cavity along the central core. This well-dened, discrete structure can serve as a scaffold to grow polyaniline nanostructures. A procedure has recently been developed to produce polyaniline nanotubes/wires with relatively narrow size distributions that are highly processable using TMV. Negatively charged surface residues of TMV can bind monomers with an amino group such as aniline through electrostatic interactions or hydrogen-bonding. When the monomer is mixed with TMV, these monomers accumulate on the surface of the virus. In situ polymerization of the absorbed monomers produces a thin layer of polyaniline on the surface of the TMV. Polyanilines produced in this fashion are core-sheath structures consisting of a TMV core and a polyaniline sheath. While the material produced is not pure polyaniline, it nonetheless possesses sufcient electrical conductivity for a number of applications.123 Besides the methods described above, the formation of polyaniline nanotubes has also been observed under ultrasonic irradiation.109 For example, polyaniline nanotubes can be produced by oxidizing aniline in the presence of a great number of dopant acids and APS while in an ultrasonic bath. Polyaniline nanotubes
3544 | J. Mater. Chem., 2011, 21, 35343550

containing Fe3O4 nanoparticles have also been produced by synthesizing aniline dimer capped Fe3O4 nanoparticles and subsequently polymerizing aniline in the presence of these nanoparticles under ultrasonic irradiation.123 A micelle model derived from the aggregation of dopant/monomer salts has been proposed as a mechanism for this process, while ultrasonic irradiation has been suggested as a means to limit aggregation. Loadings as high as 20 wt% has been achieved with this process.123

7. Spheres, hollow spheres, and core shells


The hollow sphere morphology is a potentially useful structure for a variety of applications such as drug delivery and encapsulation or protection of sensitive chemical or biological species.124 These structures are typically generated through the use of templates with preexisting nano- or micro-structured features; however, polyaniline hollow spheres have often been observed when oxidizing aniline in emulsion type polymerizations and/or under conditions of pH $34.106,111,125 These hollow spheres are typically microns in diameter, but the walls of the hollow spheres are on the order of tens of nanometers. In this process, aniline is oxidized with APS in the presence of a dopant acid such as naphthalene sulfonic acid at low temperatures. The resulting morphology is a mixture of nanotubes and hollow spheres when the mole ratio of dopant:monomer:oxidant is 1 : 4 : 4.124 It is believed that the anilinium salts form two kinds of particles in the emulsionmicelles and monomer droplets. The aniline droplets in the emulsion are believed to be composed of aniline monomer and aniline salts that aggregate into micelles; this, in turn, acts as the core and the shell of the polymer, respectively. Other acids have also been used to generate hollow spheres such as salicylic acid andlike naphthalene sulfonic acidthe ratio of dopant to monomer is crucial in forming hollow spheres.100 FTIR and X-ray diffraction measurements of polyaniline-salicyclic acid hollow spheres suggest that hydrogen bonding between dopant and polymer may provide the driving force for assembly of the polymer into the hollow sphere structure. Lowering the dopant/ monomer ratio from 1.0 to 0.1 changes the morphology of the resulting polymer from hollow spheres to nanotubes.100 In a study by MacDiarmid et al., nano/micro self-assembled hollow spheres were obtained by oxidizing aniline at dilute concentrations of pH $34.111 Extensive characterization of the polyaniline produced in this fashion revealed that the rstformed product of the reaction, a yellow/brown precipitate, is believed to be a member of the class of azanes. As the reaction proceeds, APS degrades into sulfuric acid and the pH of the solution falls, which allows the reaction to proceed as normal. Eventually green polyaniline is formed. The oxidation of aniline can be performed in a buffer solution of pH $3 which also produces a material containing azanes with a hollow sphere structure. Although the mechanism is not fully understood, the hollow sphere morphology should be highly sensitive to the nature and distribution of surface charges along the precipitate which is intimately related to the structure of the polymer and how aniline is initially oxidized.111 We have observed that this phenomenon appears to be quite general for many different types of weak acids, which suggest that the pH of the solution plays a crucial role in the formation of
This journal is The Royal Society of Chemistry 2011

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

View Online

Fig. 12 An SEM image of polyaniline produced by the falling pH method of oxidizing aniline.111 This process, and others analogous to it, can often lead to the production of polyaniline possessing a hollow sphere morphology.

hollow spheres. In fact, oxidizing methoxy-aniline under acid free conditions can readily produce polyanisidine hollow spheres similar to those observed in Fig. 12. In this process, anisidine is dissolved in water with small amounts of o-phenylenediamine and then a solution of APS in water is rapidly added into the solution containing the monomer. After a one day reaction time, the resulting puried polymer possesses a morphology consisting of polydisperse hollow spheres. The hollow spheres can range in size from 50 nm to 5 mm in diameter with a shell thickness on the order of 10100 nm. Interestingly, this process appears analogous to the falling pH method for generating polyaniline hollow spheres or nanotubes except that small amounts of o-phenylenediamine are added into the polymerization. However, when polymerizing anisidine, only hollow spheres are observed, and not nanotubes, unlike the parent polymer which produces a mixture. The differences in morphology observed between aniline and anisidine when oxidizing these monomers suggests that the basicity and solubility of the monomers has an effect on which hollow structures are observed: spheres or tubes. It also points to the importance of introducing additives such as o-phenylenediamine in generating these structures for substituted polyanilines and of the possible role of phenazine-like moieties or azane linkages in guiding the growth of these structures. Similar polyanisidine structures and other polyaniline derivatives126a have also been observed when o-anisidine is polymerized in the presence of a small amount of an initiator with cupric acetate or basic cupric bromide126b or in dilute p-TSA.126c Besides these approaches, polyaniline hollow spheres and spheres can also be produced by a number of different techniques such as hydrothermal synthesis127 or grown from grafted polymer substrates.128 Besides the spontaneous formation of polyaniline hollow spheres under the conditions mentioned above, hollow structures can also be generated by using soft templates such as CTAB129 or existing structures with a predened morphology.130132 For example, polyaniline has been grown off of a variety of polymer spheres in order to produce core shell structures of polyaniline on
This journal is The Royal Society of Chemistry 2011

top of the template polymer.130 Polystyrene has been the most commonly employed polymer in this process and the procedure typically involves polymerizing aniline monomers that have been absorbed onto the polystyrene surface. Layer-by-layer (LBL) deposition techniques can also be used to electrostatically deposit a layer of polyaniline onto charged colloidal particles such as negatively charged or sulfonated polystyrene particles.133 Besides polystyrene, a number of other spherically shaped polymer particles can be used as a template material.134 In many cases, the template can be dissolved away in order to liberate the pure conducting polymer. Furthermore, a variety of inorganic spherical particles have also been used as templates in order to deposit a layer of polyaniline onto their surfaces such as metal oxides.135 Silica spheres are most commonly used and by LBL assembly, a thin layer of polyaniline can be deposited onto the spheres surface yielding a core shell silica-polyaniline structure.136 Templating onto MnO2 hollow hierarchical nanostructures can also be used to produce polyaniline spherical and/or cubic shells with relatively narrow size distributions.137 Three-dimensional MnO2 hollow shells can be created using known literature methods. These structures can simultaneously oxidize aniline and guide the shape of the as-growing polymer. The reactive, oxidative template is converted to soluble Mn+2 as aniline is oxidized. An analogous process was also discovered in which CuO2 serves as the template in order to generate polyaniline microstructures.138 CuO2 can be oxidized by APS to form soluble Cu+2 ions which suggests possible utility as a template for directing polyaniline growth. Indeed, when aniline is oxidized with APS in the presence of octahedral CuO2 crystals and a surfactant such as dodecylbenzenesulfonate, the resulting polyaniline adopts the size and shape of the reactive CuO2 template. These techniques appear to be quite general and perhaps represent a viable technique to generate a wide range of different polyaniline sizes and shapes. In addition to these inorganic based templates, certain inorganic based oxidants such as auric acid are capable of oxidizing aniline and producing polyaniline with a spherical morphology that is also decorated with gold nanoparticles.139

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

8. Complex hierarchical structures


The oxidation of aniline can produce precipitates possessing many other types of complex micro- and nano-structured morphologies such as structures within structures (Fig. 4).A few examples of the elegant possibilities should sufce. For example, hollow rambutan-like spheres of a polymer produced by the oxidation of aniline is observed when aniline is oxidized with FeCl3 in the presence of peruorooctane sulfonic acid.140 The polymers morphology resembles the shell of the tropical fruit rambutan. Mechanistic studies indicate that the morphology of the polymer is dependent on the polymerization timeshorter reaction times (<60 min) generally favors half-formed hollow structures while longer reaction times ($6 h) favors fully formed hollow rambutan-like structures with diameters ranging from 600940 nm. The polymer synthesized in this process is conductive and hydrophobic because of the uorinated dopant acid.
J. Mater. Chem., 2011, 21, 35343550 | 3545

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

Polyaniline hierarchical structures composed of aligned nanobers/wires that aggregate into leaf-like structures have been observed when aniline is oxidized in the presence of an amphiphilic triblock copolymer.141 Poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide) (PEO-PPO-PEO), known as the amphiphilic triblock copolymer F127 (represented as EO100PO70EO100) is a surfactant that, when introduced into the polymerization of aniline, guides the shape of the growing polymer into a leaf-like structure in which the leaves are composed of nanobers that are 25 nm in diameter (Fig. 13). The leaves themselves are 3.3 mm in length, 1.4 mm in width with a thickness of 150 nm. By changing the concentration of F127, the leaf-like superstructures remain, but the sizes and thickness of the leaves can be tuned. Interestingly, when the concentration of aniline is increased, the substructure of the leaves can be tuned from nanobers to nanotubes. A mechanism involving hydrogen bonding between aniline and the EO segments and formation of aniline droplets at higher aniline concentrations has been proposed.141 Many other hierarchical structures have been observed when oxidizing aniline under various conditions. These processes simply rely on the polymers propensity for self-assembling into complex structures. For example, three-dimensional rose-like structures consisting of 2-D sheets have been observed when oxidizing aniline with a 1 : 1 molar ratio of monomer to oxidant in water when the humidity of the environment is high.142 1-D

nanostructures that assemble into a brain-like mat when aniline is oxidized with citric acid as a hard template associated with a solid/gas reaction have also been observed.143 In addition, a variety of ower-like superstructures have been reported besides the rose-like structures already mentioned.144 These ower-like hierarchical structures typically consist of nanobers/ wires or 2-D nanosheets that cluster into superstructures that resemble owers and have been observed under a variety of conditions for the oxidation of aniline (Fig. 13). Other unusual and hierarchical structures have also been seen including, towershaped superstructures composed of polymer discs,145 hollow rectangular nanostructures,146 sea urchin-like morphologies,147 and even centipede-like structures which typically consist of dendrites growing off of a ber.148 There are several processes for oxidizing aniline that can produce many structures aside from the above mentioned structures, though the formation mechanism of many of these processes remains unclear.149 Our own studies, which have recently focused on substituted polyanilines, have also revealed many uncommon micro- and nano-morphologies. For example, oxidizing 2-uoroaniline in the presence of a small amount of o-phenylenediamine produces poly-2-uoroaniline possessing a helical morphology (Fig. 13C). Changing the additive introduced into the polymerization of uoroaniline provides a convenient way to tune the polymers morphology. For instance, introducing 2,40 -diaminodiphenylamine into the oxidation of 2-uoroaniline produces a polymer

Fig. 13 SEM images of a variety of hierarchical polyaniline nano- and micro-structures. a) polyaniline leaves that are composed of nanobers synthesized through the use of an amphiphilic triblock copolymer,141 b) polyaniline nanotowers consisting of stacked disks, c) helical polyuoroaniline structures synthesized by oxidizing 2-uoroaniline in the presence of a small amount of aniline oligomeric additive and d) ower-like structures of poly(methylthio)aniline synthesized by oxidizing methylthioaniline with a small amount of 1,2,4,5 tetraaminobenzene added into the reaction. Scale bars in bd: 1 mm.

3546 | J. Mater. Chem., 2011, 21, 35343550

This journal is The Royal Society of Chemistry 2011

View Online

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

with a hierarchical tower-like superstructure consisting of individual nanosized discs as well as a mixture of other morphologies such as clusters of spheres that aggregate into a wire superstructure. Tuning other reaction parameters such as the dopant acid used during oxidation can produce structures such as twisted nanobers of 80 nm diameters or even large micron-sized spheres that are composed of nanobers. The oxidation of other substituted anilines can also produce a wide range of morphologies. For example, the oxidation of 2-chloroaniline with APS in the presence of o-phenylenediamine produces poly-2-chloroaniline with a hollow rectangle morphology. A large number of reaction parameters can be changed which all seem to produce different polymer morphologies, although the mechanism for many of these processes remains unclear.

9. Conclusion
The oxidation of aniline and substituted anilines can produce an almost limitless number of micro- and nano-structures of different sizes and shapes including nanobers, nanowires, nanotubes, hollow spheres, helices, owers, and many more. Modifying a wide variety of synthetic conditions such as reagent concentration, pH, homogeneous or heterogeneous polymerization, addition rate of reagents, or even the amount of sheer during polymerization can have a profound impact on

the resulting polymers nanoscale morphology. For this reason, polyaniline has the potential to substantially contribute to the eld of nano/macro science as few other materials possess such morphological exibility. This allows one to intimately study the structure/property relationships of these materials at the nanoscale and potentially develop applications tailored for different types of morphologies depending on the properties needed. Yet an understanding of this relationship for polyaniline nanostructures is quite limited despite the many different types of morphologies produced. This is in large part due to the fact that many of the different types of polyaniline nanostructures have different structural variations at the polymeric chain level; i.e. are the different optical and transport properties observed in different types of polyaniline nanostructures a result of their different morphologies or are they simply a result of having different molecular weights, oxidation states, cross-linking degree, or perhaps a combination of all of these? How does one decouple the differences between nanoscale morphology and the structural variations inherent in polymerizing aniline under different synthetic conditions? Furthermore, can a general mechanism be proposed for all these observed nanostructures or are the structural variations of each system the key to understanding why polyaniline displays such a knack for forming different well dened nano/micro structures?

Fig. 14 SEM images of emeraldine base aniline tetramer dissolved and then reprecipitated into a solution of acid. The morphology of the aniline tetramer changes based on the type of acid used to dope and reprecipitate the tetramer. A) Rectangular shaped aniline tetramer produced by reprecipitating from an aqueous solution of 0.5 M HClO4, B) a cross-sectional view of the nano rectangles, C) ower shaped aniline tetramer using 1.0 M H2SO4 to dope and reprecipitate the tetramer and D) aniline tetramer nanobers produced by doping and reprecipitating from 1.0 M HCl solution. Scale bar: 100 nm for (B), 1 mm for all the others. Adapted from ref. 150.

This journal is The Royal Society of Chemistry 2011

J. Mater. Chem., 2011, 21, 35343550 | 3547

View Online

Our group has sought to address these questions and recently, we have begun investigating aniline oligomers as a model system in order to eliminate the variations in chain structure that make interpreting the effects of nanoscale morphologies on physical properties difcult.148 Specically, when the emeraldine base form of aniline tetramer is dissolved and reprecipitated into a solution containing a dopant acid, a variety of nanoscale morphologies can be produced depending on the dopant acid used to reprecipitate the oligomer. The morphologies mimic those observed in the polyaniline systems; however, the oligomer nanostructures all possess identical molecular structures (Fig. 14). This allows for the study of how physical properties depend on the intermolecular interactions within nanostructured aniline-based systems. Furthermore, it also allows us to examine the nature of the self-assembly process without the complications of accounting for differences in molecular structure that are impossible to avoid in the polymeric system. Indeed, the electrical conductivity of the tetraaniline nanostructures is dependent on its nanoscale morphology, which is a direct result of the differences in the solid-state packing of the oligomer. In fact, the conductivities of crystalline tetraaniline nanobers is actually several orders of magnitude higher than previously reported for bulk tetraaniline and now rivals that of bulk polyaniline.150 This underscores the importance of solid-state chain packing in determining the morphology and physical properties of these materials. Given all the different micro- and nano-scale structures produced thus far, it is clear that Alan MacDiarmid anticipated the complexities involved when he stated there are as many different types of polyaniline as there are people who make it! Indeed there are. One can also say that there are also as many different morphologies and shapes of polyaniline as there are people who make them. The urry of research activity recently directed towards understanding this process has certainly been a boon to the eld. While many formation mechanisms have been proposed, they are not necessarily incompatible. In fact, many of them are complementary and the true answer may be a combination of different thoughts produced by the many excellent researchers in this eld. In the coming years, we expect this debate to continue as another generation of scientists discovers the richness and diversity of polyaniline.

Acknowledgements
Support for this research has been provided by the UCLA based Focused Center Research Program Functional Engineered NanoArchetectonics center (R.B.K.), the National Science Foundation Small Business Technology Transfer program under Award Number 1010540 (H.D.T and R.B.K.), a National Science Foundation Integrative Graduate Education and Research Training: Clean Energy for Green Industry Fellowship DGE-0903720 (Y.W.), and a National Science Foundation NanoCer summer research fellowship 0649323 (P.J.B.).

Notes and references


1 H. D. Tran, D. Li and R. B. Kaner, Adv. Mater., 2009, 21, 1487. 2 (a) M. X. Wan, Adv. Mater., 2008, 20, 2926; (b) M. X. Wan, Macromol. Rapid Commun., 2009, 30, 963.

3 (a) J. X. Huang and R. B. Kaner, Chem. Commun., 2006, 367; (b) D. Li, J. X. Huang and R. B. Kaner, Acc. Chem. Res., 2009, 42(1), 135. 4 (a) J. Jang, Adv. Polym. Sci., 2006, 199, 189259; (b) J. Stejskal, I. Sapurina, and M. Trchova, Prog. Polym. Sci., DOI: 10.1016/ j.progpolymsci.2010.07.006. 5 D. Zhang and Y. Wang, Mater. Sci. Eng., B, 2006, 134, 9. 6 Y. N. Xia, P. D. Yang, Y. G. Sun, Y. Y. Wu, B. Mayers, B. Gates, Y. D. Yin, F. Kim and Y. Q. Yan, Adv. Mater., 2003, 15, 353. 7 Y. Yin and A. P. Alivisatos, Nature, 2005, 437, 664. 8 S. Virji, J. X. Huang, R. B. Kaner and B. H. Weiller, Nano Lett., 2004, 4, 491. 9 Z. R. Tian, J. A. Voigt, J. Liu, B. Mckenzie, M. J. Mcdermott, M. A. Rodriguez, H. Konishi and H. Xu, Nat. Mater., 2003, 2, 821. 10 Z. L. Wang, J. Phys.: Condens. Matter, 2004, 16, R829. 11 M. J. Bierman, Y. K. A. Lau, A. V. Kvit, A. L. Schmitt and S. Jin, Science, 2008, 320, 1060. 12 (a) A. L. Briseno, S. C. B. Mannsfeld, S. A. Jenekhe, Z. Bao and Y. Xia, Mater. Today, 2008, 11, 38; (b) J. A. Lim, F. Liu, S. Ferdous, M. Muthukumar and A. L. Briseno, Mater. Today, 2010, 13, 14. 13 Unverdorben, Ann. Phys. Chem., 1826, 8, 397. 14 F. F. Runge, Ann. Phys. Chem., 1834, 31, 513. 15 J. Frtzche, J. fr Prakt. Chem., 1840, 20, 454. u 16 (a) A. G. Green and A. E. Woodhead, J. Chem. Soc., 1910, 97, 2388; (b) A. G. Green and A. E. Woodhead, J. Chem. Soc., 1912, 101, 1117. 17 R. Willsttter and C. W. Moore, Ber. Dtsch. Chem. Ges., 1907, 40, a 2665. 18 A. G. MacDiarmid, J.-C. Chiang, M. Halpern, W.-S. Huang, S.-L. Mu, M. L. D. Somasiri, W. Wu and S. I. Yaniger, Mol. Cryst. Liq. Cryst., 1985, 121, 173. 19 (a) R. de Surville, M. Josefowicz, L. T. Yu, J. Perichon and R. Buvet, Electrochim. Acta, 1968, 13, 1451; (b) F. Cristoni, R. de Surville, M. Josefowicz, L. T. Yu and R. Buvet, Seances Acad. Sci., Ser. C, 1969, 268, 1346; (c) M. Josefowicz, L. T. Yu, J. Perichon and R. Buvet, J. Polym. Sci. Part C, 1969, 22, 1187. 20 F. Lux, Polymer, 1994, 35, 2915. 21 W.-S. Huang, B. D. Humphrey and A. G. MacDiarmid, J. Chem. Soc., Faraday Trans. 1, 1986, 82, 2385. 22 J. Stejskal and I. Sapurina, Pure Appl. Chem., 2005, 77, 815. 23 (a) F. Wudl, R. O. AngusJr, F. L. Lu, P. M. Allemand, D. J. Vachon, M. Nowak, Z. X. Liu and A. J. Heeger, J. Am. Chem. Soc., 1987, 109, 3677; (b) D. Vachon, R. O. AngusJr., F. L. Lu, M. Nowak, Z. X. Liu, H. Schaffer, F. Wudl and A. J. Heeger, Synth. Met., 1987, 18, 297. 24 I. Sapurina and J. Stejskal, Polym. Int., 2008, 57, 1295. 25 D. Li and R. B. Kaner, Chem. Commun., 2005, 3286. 26 A. G. MacDiarmid, W. E. Jones Jr., I. D. Norris, J. Gao, A. T. Johnson Jr., N. J. Pinto, J. Hone, B. Han, F. K. Ko, H. Okuzaki and M. Llaguno, Synth. Met., 2001, 119, 27. 27 C. R. Fincher Jr., D. Moses, A. J. Heeger and A. G. MacDiarmid, Synth. Met., 1983, 6, 243. 28 (a) S. H. Weng, Z. H. Lin, L. X. Chen and J. Z. Zhou, Electrochim. Acta, 2010, 55, 2727; (b) Y. Guo and Y. Zhou, Eur. Polym. J., 2007, 43, 2292; (c) G.-R. Li, Z.-P. Feng, J.-H. Zhong, Z.-L. Wang and Y.-X. Tong, Macromolecules, 2010, 43, 2178; (d) D. S. Dhawale, R. R. Salunkhe, V. S. Jamadade, T. P. Gujar and C. D. Lokhande, Appl. Surf. Sci., 2009, 255, 8213. 29 R. E. Anderson, A. D. Ostrowski, D. E. Gran, J. D. Fowler, A. R. Hopkins and R. M. Villahermosa, Polym. Bull., 2008, 61, 563. 30 N. J. Pinto, A. T. Johnson, A. G. MacDiarmid, C. H. Mueller, N. Theylaktos, D. C. Robinson and F. A. Miranda, Appl. Phys. Lett., 2003, 83, 4244. 31 A. G. MacDiarmid, W. E. Jones, I. D. Norris, J. Gao, A. T. Johnson, N. J. Pinto, J. Hone, B. Han, F. K. Ko, H. Okuzaki and M. Llaguno, Synth. Met., 2001, 119, 27. 32 N. J. Pinto, I. Ramos, R. Rojas, P.-C. Wang and A. T. Johnson Jr., Sens. Actuators, B, 2008, 129, 621. 33 H. X. He, C. Z. Li and N. J. Tao, Appl. Phys. Lett., 2001, 78, 811. 34 M. X. Wan, Adv. Mater., 2008, 20, 2926. 35 (a) C. R. Martin, Chem. Mater., 1996, 8, 1739; (b) R. V. Parthasarathy and C. R. Martin, Chem. Mater., 1994, 6, 1627. 36 (a) Yu, Y. Li and K. Kalantar-zadeh, Sens. Actuators, B, 2009, 136, 1; (b) H. Qiu, J. Zhai, S. Li, L. Jiang and M. X. Wan, Adv. Funct.

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

3548 | J. Mater. Chem., 2011, 21, 35343550

This journal is The Royal Society of Chemistry 2011

View Online

37 38 39 40 41 42

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

43 44 45 46

47 48 49 50 51 52

53 54 55 56

57 58 59 60 61

62 63 64 65 66 67

Mater., 2003, 13, 925; (c) Y. Cao and T. E. Mallouk, Chem. Mater., 2008, 20, 5260. J. Li, Q. Jia, J. Zhu and M. Zheng, Polym. Int., 2008, 57, 337. G. C. Li and Z. K. Zhang, Macromolecules, 2004, 37, 2683. B. Lindman, H. Wennerstrom, Micelles: Amphiphile aggregation in aqueous solution. Berlin; New York: Springer-Verlag, 1980. L. Zhang, M. X. Wan and Y. Wei, Macromol. Rapid Commun., 2006, 27, 366. L. Zhang, L. Zhang, M. X. Wan and Y. Wei, Synth. Met., 2006, 156, 454. (a) Z. Zhang, Z. Wei, L. Zhang and M. X. Wan, Acta Mater., 2005, 53, 1373; (b) Z. Wei, Z. Zhang and M. X. Wan, Langmuir, 2002, 18, 917; (c) K. Huang, X.-H. Meng and M. X. Wan, J. Appl. Polym. Sci., 2006, 100, 3050. K. Huang and M. X. Wan, Chem. Mater., 2002, 14, 3486. Z. Zhang and S. K. Manohar, Chem. Commun., 2004, 2360. Z. Zhang, Z. Wei and M. X. Wan, Macromolecules, 2002, 35, 5937. (a) P. Anilkumar and M. Jayakannan, Macromolecules, 2007, 40, 7311; (b) M. J. Antony and M. Jayakannan, J. Phys. Chem. B, 2010, 114, 1314; (c) J. Wang, J. Wang, X. Zhang and Z. Wang, Macromol. Rapid Commun., 2007, 28, 84; (d) P. Anilkumar and M. Jayakannan, Macromolecules, 2008, 41, 7706. N.-R. Chiou, L. J. Lee and A. J. Epstein, Chem. Mater., 2007, 19, 3589. Z. X. Wei, L. J. Zhang, M. Yu, Y. S. Yang and M. X. Wan, Adv. Mater., 2003, 15, 1382. J. M. Liu and S. C. Yang, J. Chem. Soc., Chem. Commun., 1991, 1529. S. J. Choi and S. M. Park, Adv. Mater., 2000, 12, 1547. T. Hatano, M. Takeuchi, A. Ikeda and S. Shinkai, Chem. Lett., 2003, 32, 314. (a) P. Nickels, W. U. Dittmer, S. Beyer, J. P. Kotthaus and F. C. Simmel, Nanotechnology, 2004, 15, 1524; (b) Y. F. Ma, J. M. Zhang, G. J. Zhang and H. X. He, J. Am. Chem. Soc., 2004, 126, 7097; (c) S. Pruneanu, S. A. F. Al-Said, L. Q. Dong, T. A. Hollis, M. A. Galindo, N. G. Wright, A. Houston and B. R. Horrocks, Adv. Funct. Mater., 2008, 18, 2444; (d) B. Datta and G. B. Schuster, J. Am. Chem. Soc., 2008, 130, 2965; (e) L. Q. Dong, T. Hollis, S. Fishwick, B. A. Connolly, N. G. Wright, B. R. Horrocks and A. Houlton, Chem.Eur. J., 2007, 13, 822; (f) B. Datta, G. B. Schuster, A. McCook, S. C. Harvey and K. Zakrzewska, J. Am. Chem. Soc., 2006, 128, 14428. X. Y. Zhang, W. J. Goux and S. K. Manohar, J. Am. Chem. Soc., 2004, 126, 4502. (a) X. Y. Zhang and S. K. Manohar, J. Am. Chem. Soc., 2004, 126, 12714; (b) X. Y. Zhang and S. K. Manohar, J. Am. Chem. Soc., 2005, 127, 14156. (a) J. X. Huang and R. B. Kaner, J. Am. Chem. Soc., 2004, 126, 851; (b) J. X. Huang, S. Virji, B. H. Weiller and R. B. Kaner, J. Am. Chem. Soc., 2003, 125, 314. (a) X. Zhang, R. Chan-Yu-King, A. Jose and S. K. Manohar, Synth. Met., 2004, 145, 23; (b) S. Xing, H. Zheng and G. Zhao, Synth. Met., 2008, 158, 59; (c) S. Ameen, M. S. Akhtar, Y. S. Kim, O. B. Yang and H. S. Shin, J. Phys. Chem. C, 2010, 114, 4760. J. X. Huang and R. B. Kaner, Angew. Chem., Int. Ed., 2004, 43, 5817. J. Qiang, Z. Yu, H. Wu and D. Yun, Synth. Met., 2008, 158, 544. (a) N.-R. Chiou and A. J. Epstein, Adv. Mater., 2005, 17, 1679; (b) N.-R. Chiou and A. J. Epstein, Synth. Met., 2005, 153, 69. Y. Wang and X. Jing, J. Phys. Chem. B, 2008, 112, 1157. (a) G. Li, L. Jiang and H. Peng, Macromolecules, 2007, 40, 7890; (b) H. Ding, M. Wan and Y. Wei, Adv. Mater., 2007, 19, 465; (c) Z. Zhang, J. Deng and M. Wan, Mater. Chem. Phys., 2009, 115, 275; (d) G. Li, C. Zhang, Y. Li, H. Peng and K. Chen, Polymer, 2010, 51, 1934; (e) N. T. Tung, H. Lee, Y. Song, N. D. Nghia and D. Sohn, Synth. Met., 2010, 160, 1303. D. Li and R. B. Kaner, J. Am. Chem. Soc., 2006, 128, 968. D. Li and R. B. Kaner, J. Mater. Chem., 2007, 17, 2279. H. D. Tran, Y. Wang, J. M. DArcy and R. B. Kaner, ACS Nano, 2008, 2, 1841. S. P. Surwade, N. Manohar and S. K. Manohar, Macromolecules, 2009, 42, 1792. S. P. Surwade, S. R. Agnihotra, V. Dua, N. Manohar, S. Jain, S. Ammu and S. K. Manohar, J. Am. Chem. Soc., 2009, 131, 12528. X. Y. Zhang, H. S. Kolla, X. H. Wang, K. Raja and S. K. Manohar, Adv. Funct. Mater., 2006, 16, 1145.

68 B. Wunderlich, Crystal structure, morphology, defects. Macromol. Phys., Academic Press: New York, 1973. 69 Y. Wang, H. D. Tran and R. B. Kaner, Macromol. Rapid Commun., 2010, DOI: 10.1002/marc.201000280, ASAP. 70 H. D. Tran, K. Shin, W. G. Hong, J. M. DArcy, R. W. Kojima, B. H. Weiller and R. B. Kaner, Macromol. Rapid Commun., 2007, 28, 2289. 71 (a) W. G. Li and H.-L. Wang, J. Am. Chem. Soc., 2004, 126, 2278; (b) W. G. Li, J. A. Bailey and H.-L. Wang, Polymer, 2006, 47, 3112. 72 (a) Z. D. Zujovic, G. A. Bowmaker, H. D. Tran and R. B. Kaner, Synth. Met., 2009, 159, 710; (b) A. R. Hopkins, R. A. Lipeles and S.-J. Hwang, Synth. Met., 2008, 158, 594; (c) G. M. do Nascimento, P. Y. G. Kobata and M. L. A. Temperini, J. Phys. Chem. B, 2008, 112, 11551; (d) Z. D. Zujovic, L. Zhang, G. A. Bowmaker, P. A. Kilmartin and J. Travas-Sejdic, Macromolecules, 2008, 41, 3125. 73 D. Yang, P. N. Adams, R. Goering and B. R. Mattes, Synth. Met., 2003, 135136, 293. 74 S. K. Pillalamarri, F. D. Blum, A. T. Tokuhiro, J. G. Story and M. F. Bertino, Chem. Mater., 2005, 17, 227. 75 (a) X. L. Jing, Y. Y. Wang, D. Wu and J. P. Qiang, Ultrason. Sonochem., 2007, 14, 75; (b) Y. Li, Y. Wang, D. Wu and X. Jing, J. Appl. Polym. Sci., 2009, 113, 868; (c) Y. Wang, X. Jing and J. Kong, Synth. Met., 2007, 157, 269. 76 M. R. Gizdavic-Nikolaidis, D. R. Stanisavljev, A. J. Easteal and Z. D. Zujovic, Macromol. Rapid Commun., 2010, 31, 657. 77 (a) X.-S. Du, C.-F. Zhou, G.-T. Wang and Y.-W. Mai, Chem. Mater., 2008, 20, 3806; (b) C.-F. Zhou, X.-S. Du, Z. Liu, S. P. Ringer and Y.-W. Mai, Synth. Met., 2009, 159, 1302. 78 H. D. Tran and R. B. Kaner, Chem. Commun., 2006, 3915. 79 H. D. Tran, I. Norris, J. M. DArcy, H. Tsang, Y. Wang, B. R. Mattes and R. B. Kaner, Macromolecules, 2008, 41, 7405. 80 C. Zhang, G. Li and H. Peng, Mater. Lett., 2009, 63, 592. 81 J. Han, Y. Liu and R. Guo, J. Polym. Sci., Part A: Polym. Chem., 2008, 46, 740746. 82 B. A. Deore, I. Yu, J. Woodmass and M. S. Freund, Macromol. Chem. Phys., 2008, 209, 1094. 83 B. A. Deore, I. S. Yu, P. M. Aguiar, C. Recksiedler, S. Kroeker and M. S. Freund, Chem. Mater., 2005, 17, 3803. 84 J. Ryu and C. B. Park, Angew. Chem., Int. Ed., 2009, 48, 4820. 85 C.-G. Wu and T. Bein, Science, 1994, 264, 1757. 86 L. Meng, Y. Lu, X. Wang, J. Zhang, Y. Duan and C. Li, Macromolecules, 2007, 40, 2981. 87 J. Li, H. Tang, A. Zhang, X. Shen and L. Zhu, Macromol. Rapid Commun., 2007, 28, 740. 88 C. M. Hangarter, M. Bangar, A. Mulchandani and N. V. Myung, J. Mater. Chem., 2010, 20, 3131. 89 F. Gu, L. Zhang, X. Yin and L. Tong, Nano Lett., 2008, 8, 2757. 90 M. D. Shirsat, M. A. Bangar, M. A. Deshusses, N. V. Myung and A. Mulchandani, Appl. Phys. Lett., 2009, 94, 083502. 91 J. Liu, Y. Lin, L. Liang, J. A. Voigt, D. L. Huber, Z. R. Tian, E. Coker, B. Mckenzie and M. J. Mcdermott, Chem. Eur. J., 2003, 9, 605. 92 N. R. Chiou, C. M. Lui, J. J. Guan, L. J. Lee and A. J. Epstein, Nat. Nanotechnol., 2007, 2, 354. 93 Q. Sun, W. Bi, T. F. Fuller, Y. Ding and Y. Deng, Macromol. Rapid Commun., 2009, 30, 1027. 94 A. Vlad, S. Yunus, A. Attout, D. A. Serban, L. Gence, S. Faniel, J.-F. Gohy, P. Bertrand and S. Melinte, Small, 2010, 6, 627. 95 (a) Q. Tang, J. Wu, X. Sun, Q. Li, J. Lin and M. Huang, Chem. Commun., 2009, 2166; (b) Q. Tang, J. Wu, X. Sun, Q. Li and J. Lin, Langmuir, 2009, 25, 5253. 96 Y. Wang, H. D. Tran and R. B. Kaner, J. Phys. Chem. C, 2009, 113, 10346. 97 B. Nandan, J.-Y. Hsu, A. Chiba, H.-L. Chen, C.-S. Liao, S.-A. Chen and H. Hasegawa, Macromolecules, 2007, 40, 395. 98 B. Dong, N. Lu, M. Zelsmann, N. Kehagias, H. Fuchs, C. M. S. Torres and L. Chi, Adv. Funct. Mater., 2006, 16, 1937. 99 C. R. Martin, Acc. Chem. Res., 1995, 28, 61. 100 (a) L. Zhang and M. Wan, Adv. Funct. Mater., 2003, 13, 815; (b) L. Zhang, Y. Long, Z. Chen and M. X. Wan, Adv. Funct. Mater., 2004, 14, 693. 101 Z. Zhang, M. Wan and Y. Wei, Adv. Funct. Mater., 2006, 16, 1100. 102 (a) Z. Wei, L. Zhang, M. Yu, Y. Yang and M. X. Wan, Adv. Mater., 2003, 15, 1382; (b) H. Qiu and M. X. Wan, Macromolecules, 2001,

This journal is The Royal Society of Chemistry 2011

J. Mater. Chem., 2011, 21, 35343550 | 3549

View Online

103 104 105

106

Downloaded by Bhabha Atomic Research Centre on 26 July 2011 Published on 25 November 2010 on http://pubs.rsc.org | doi:10.1039/C0JM02699A

107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122

123 124 125 126

127

34, 675; (c) Z. Zhang, Z. Wei and M. Wan, Macromolecules, 2002, 35, 5937. X. Feng, G. Yang, Q. Xu, W. Hou and J.-J. Zhu, Macromol. Rapid Commun., 2006, 27, 31. B.-Z. Hsieh, H.-Y. Chuang, L. Chao, Y.-J. Li, Y.-J. Huang, P.-H. Tseng, T.-H. Hsieh and K.-S. Ho, Polymer, 2008, 49, 4218. (a) M. Trchova, I. Sedenkova, E. N. Konyushenko, J. Stejskal, P. Holler and G. Ciric-Marjanovic, J. Phys. Chem. B, 2006, 110, 9461; (b) Z. D. Zujovic, C. Laslau, G. A. Bowmaker, P. A. Kilmartin, A. L. Webber, S. P. Brown and J. Travas-Sejdic, Macromolecules, 2010, 43, 662; (c) C. Laslau, Z. D. Zujovic, L. J. Zhang, G. A. Bowmaker and J. Travas-Sejdic, Chem. Mater., 2009, 21, 954. J. Stejskal, I. Sapurina, M. Trchova and E. N. Konyushenko, Macromolecules, 2008, 41, 3530. J. Stejskal, I. Sapurina, M. Trchova, E. N. Konyushenko and P. Holler, Polymer, 2006, 47, 8253. E. N. Konyushenko, J. Stejskal, I. Sedenkova, M. Trchova, I. Sapurina, M. Cieslar and J. Prokes, Polym. Int., 2006, 55, 31. X. Lu, H. Mao, D. Chao, W. Zhang and Y. Wei, Macromol. Chem. Phys., 2006, 207, 2142. J. Kriz, L. Starovoytova, M. Trchova, E. N. Konyushenko and J. Stejskal, J. Phys. Chem. B, 2009, 113, 6666. E. C. Venancio, P.-C. Wang and A. G. MacDiarmid, Synth. Met., 2006, 156, 357. D.-C. Lee, B. Cao, K. Jang and P. M. Forster, J. Mater. Chem., 2010, 20, 867. G. Ciric-Marjanovic, L. Dragicevic, M. Milojevic, M. Mojovic, S. Mentus, B. Dojcinovic, B. Marjanovic and J. Stejskal, J. Phys. Chem. B, 2009, 113, 7116. G. Ciric-Marjanovic, V. Dondur, M. Milojevic, M. Mojovic, S. Mentus, A. Radulovic, Z. Vukovic and J. Stejskal, Langmuir, 2009, 25, 3122. Y. Gao, D. Shan, F. Cao, J. Gong, X. Li, H.-Y. Ma, Z.-M. Su and L.-Y. Qu, J. Phys. Chem. C, 2009, 113, 15175. H. Dong, S. Prasad, V. Nyame and W. E. Jones Jr, Chem. Mater., 2004, 16, 371. Y. Gao, S. Yao, J. Gong and L. Y. Qu, Macromol. Rapid Commun., 2007, 28, 286. T. Y. Dai and Y. Lu, Macromol. Rapid Commun., 2007, 28, 629. (a) L. Zhang and P. Liu, Nanoscale Res. Lett., 2008, 3, 299; (b) L. Zhang, T. M. Wang and P. Liu, Appl. Surf. Sci., 2008, 255, 2091. T. H. Kim, Y. Kim, S. J. Lee, W. S. Han and J. H. Jung, Chem. Lett., 2008, 37, 598. L. J. Pan, L. Pu, Y. Shi, S. Y. Song, Z. Xu, R. Zhang and Y. D. Zheng, Adv. Mater., 2007, 19, 461. (a) Z. Niu, J. Liu, L. A. Lee, M. A. Bruckman, D. Zhao, G. Koley and Q. Wang, Nano Lett., 2007, 7, 3729; (b) Z. W. Niu, M. A. Bruckman, S. Q. Li, L. A. Lee, B. Lee, S. V. Pingali, P. Thiyagarajan and Q. Wang, Langmuir, 2007, 23, 6719; (c) Z. W. Niu, M. Bruckman, V. S. Kotakadi, J. B. He, T. Emrick, T. P. Russell, L. Yang and Q. Wang, Chem. Commun., 2006, 3019. X. Lu, H. Mao, D. Chao, W. Zhang and Y. Wei, J. Solid State Chem., 2006, 179, 2609. P. Liu and L. Zhang, Crit. Rev. Solid State Mater. Sci., 2009, 34(1), 7587. Y. Tan, F. Bai, D. Wang, Q. Peng, X. Wang and Y. Li, Chem. Mater., 2007, 19, 5773. (a) J. Han, G. Song and R. Guo, Adv. Mater., 2006, 18, 3140; (b) Z. Wei and M. X. Wan, Adv. Mater., 2002, 14, 1314; (c) L. Zhang, H. Peng, J. Sui, C. Soeller, P. A. Kilmartin and J. Travas-Sejdic, J. Phys. Chem. C, 2009, 113, 9128. Y.-S. Zhang, W.-H. Xu, W.-T. Yao and S.-H. Yu, J. Phys. Chem. C, 2009, 113, 8588.

128 W. Zhong, Y. Wang, Y. Yan, Y. Sun, J. Deng and W. Yang, J. Phys. Chem. B, 2007, 111, 3918. 129 J. Li, Q. Jia, J. Zhu and M. Zheng, Polym. Int., 2008, 57, 337. 130 (a) Q. Wu, Z. Wang and G. Xue, Adv. Funct. Mater., 2007, 17, 1784; (b) C. Barthet, S. P. Armes, M. M. Chehimi, C. Bilem and M. Omastova, Langmuir, 1998, 14, 5032; (c) N. Kohult-Svelko, S. Reynaud, R. Dedryvere, H. Martinez, D. Gonbeau and J. Francois, Langmuir, 2005, 21, 1575; (d) L. Y. Wang, Y. J. Liu and W. Y. Chiu, Synth. Met., 2001, 119, 155; (e) T. Lei and K. Aoki, J. Electroanal. Chem., 2000, 482, 149; (f) Y.-M. Bai, Y.J. Cheng, S. A. Wickline and Y. Xia, Small, 2009, 5, 1747. 131 E. Donath, G. B. Sukhorukov, F. Caruso, S. A. Davis and H. Mohwald, Angew. Chem., Int. Ed., 1998, 37, 2201. 132 F. Caruso, Chem.Eur. J., 2000, 6, 413. 133 (a) M.-K. Park, K. Onishi, J. Locklin, F. Caruso and R. C. Advincula, Langmuir, 2003, 19, 8550; (b) X. Feng, C. Mao, G. Yang, W. Hou and J.-J. Zhu, Langmuir, 2006, 22, 4384; (c) X. Wang, J. Liu, X. Feng, M. Guo and D. Sun, Mater. Chem. Phys., 2008, 112, 319; (d) S.-J. Ding, C.-L. Zhang, M. Yang, Z.-Z. Qu, Y.-F. Lu and Z.-Z. Yang, Polymer, 2006, 47, 8360. 134 X. Y. Shi, A. L. Briseno, R. J. Sanedrin and F. M. Zhou, Macromolecules, 2003, 36, 4093. 135 (a) C. L. Zhu, S. W. Chou, S. F. He, W. N. Liao and C. C. Chen, Nanotechnology, 2007, 18, 275604; (b) D. P. Wang and H. C. Zeng, Chem. Mater., 2009, 21, 4811; (c) D. P. Wang and H. C. Zeng, J. Phys. Chem. C, 2009, 113, 8097. 136 (a) G. D. Fu, J. P. Zhao, Y. M. Sun, E. T. Kang and K. G. Keoh, Macromolecules, 2007, 40, 2271; (b) Z. Niu, Z. Yang, Z. Hu, Y. Lu and C. C. Han, Adv. Funct. Mater., 2003, 13, 949; (c) J. Jang, J. Ha and B. Lim, Chem. Commun., 2006, 1622. 137 J. Fei, Y. Cui, X. Yan, Y. Yang, K. Wang and J. Li, ACS Nano, 2009, 3, 3714. 138 Z. Zhang, J. Sui, L. Zhang, M. X. Wan, Y. Wei and L. Yu, Adv. Mater., 2005, 17, 2854. 139 K. Mallick, M. J. Witcomb, A. Dinsmore and M. S. Scurrell, Macromol. Rapid Commun., 2005, 26, 232. 140 Y. Zhu, D. Hu, M. Wan and Y. Wei, Adv. Mater., 2007, 19, 2092. 141 J. Han, G. Song and R. Guo, Adv. Mater., 2007, 19, 2993. 142 Y. Zhu, H. He, M. Wan and L. Jiang, Macromol. Rapid Commun., 2008, 29, 1705. 143 Y. Zhu, J. Li, M. Wan, L. Jiang and Y. Wei, Macromol. Rapid Commun., 2007, 28, 1339. 144 (a) T. Wang, W. Zhong, X. Ning, Y. Wang and W. Yang, J. Colloid Interface Sci., 2009, 334, 108; (b) G. Li, C. Zhang and H. Peng, Macromol. Rapid Commun., 2008, 29, 63; (c) C. Zhou, J. Han and R. Guo, Macromolecules, 2008, 41, 6473. 145 J. Fei, Y. Cui, X. Yan, Y. Yang, Y. Su and J. Li, J. Mater. Chem., 2009, 19, 3263. 146 (a) C. Zhou, J. Han and R. Guo, Macromolecules, 2009, 42, 1252; (b) C. Zhou, J. Han and R. Guo, J. Phys. Chem. B, 2008, 112, 5014. 147 Y. Li, B. Wang and W. Feng, Synth. Met., 2009, 159, 1597. 148 J. Molina, A. I. del Rio, J. Bonastre and F. Cases, Synth. Met., 2010, 160, 99. 149 (a) H.-Y. Ma, Y.-W. Li, S.-X. Yang, F. Cao, J. Gong and Y.L. Deng, J. Phys. Chem. C, 2010, 114, 9264; (b) B. Y. Liu, L. Liu, N. L. Shi, J. Gong and C. Sun, J. Mater. Sci. Technol., 2010, 26, 39; (c) E. Jin, X. Lu, X. Bian, L. Kong, W. Zhang and C. Wang, J. Mater. Chem., 2010, 20, 3079; (d) H.-Y. Ma, Y. Gao, Y.-H. Li, J. Gong, X. Li, B. Fan and Y.-L. Deng, J. Phys. Chem. C, 2009, 113, 9047; (e) X.-S. Du, C.-F. Zhou and Y.-W. Mai, J. Phys. Chem. C, 2008, 112, 19836; (f) P. R. Sajanlal, T. S. Sreeprasad, A. S. Nair and T. Pradeep, Langmuir, 2008, 24, 4607. 150 Y. Wang, H. D. Tran, L. Liao, X. Duan and R. B. Kaner, J. Am. Chem. Soc., 2010, 132, 10365.

3550 | J. Mater. Chem., 2011, 21, 35343550

This journal is The Royal Society of Chemistry 2011

S-ar putea să vă placă și