Sie sind auf Seite 1von 16

PERGAMON

Electrochimica Acta 44 (1999) 17331748

Electropolymerization of pyrrole and electrochemical study of polypyrrole: 1. Evidence for structural diversity of polypyrrole
Ming Zhou, Ju rgen Heinze *
Institut fur Physikalische Chemie, Albert-Ludwigs-Universitat Freiburg, Albertstrasse 21, D-79104 Freiburg, Germany Received 6 April 1998; received in revised form 8 July 1998

Abstract Conventional electropolymerization of pyrrole produces polypyrrole characterized voltammetrically by a single oxidation wave (ca. 0.01 V vs. Ag/AgCl) followed by a broad plateau. This was considered to be an inherent property of electrosynthesized polypyrrole, and intensive research based on such knowledge was carried out in the past. The present studies of electrochemical properties of polypyrrole prepared at extremely low formation potentials or currents reveal that polypyrrole may exhibit additional oxidation waves at potentials more negative than those normally observed. A number of cyclic voltammograms manifest a great structural diversity in polypyrrole doped by the same anion. The fact that the trace-crossing may not appear in some cases and, on the other hand, may remain after the electrode has become polymer-coated in other cases leads to the conclusion that the nucleation step is not the only factor responsible for the trace-crossing. # 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Pyrrole; Polypyrrole; Electropolymerization; Redox; Nucleation loop

1. Introduction Despite 20 years of studies on the basic aspects of conducting polymers (CPs) [126], many issues, such as the lm formation process, the charging/discharging mechanism, and the interplay between structure, properties and synthesis conditions still remain controversial. This is, of course, due to the complexity of the electropolymerization mechanism and the lack of eective methods for the characterization of insoluble polymers. Furthermore, the neglect of the experimental conditions during the preparation of CPs is another reason for many unresolved points in the debate. This is particularly true for property studies and structural characterization of polypyrrole (PPy).

* Corresponding author. Tel.: +49-761-203-622; Fax: +49761-203-6222; E-mail: heinze@sun2.ruf.uni-freiburg.de

One method of gaining more insight into CPs is to extrapolate the properties from structurally welldened oligomers. This has proved to be largely successful. Results obtained in recent years have greatly contributed to a better understanding of charging/ discharging [22, 27], dimerization [21, 23, 28], chain branching and cross-linking [25, 29] and conductivity [30], etc. However, the knowledge of the factors governing the electrochemical reactions and the formation of polymer structures is still far from complete. Almost no certain conclusions have been drawn so far on how a single parameter inuences the polymer's properties, or how interdependent the experimental variables are. In previous work, attention was paid to the role of applied potentials in electropolymerization, and it was found that higher potentials result in oligomeric radical cations with a higher charge level, leading to longer chains in the polymer structure [27].

0013-4686/99/$ - see front matter # 1999 Elsevier Science Ltd. All rights reserved. PII: S 0 0 1 3 - 4 6 8 6 ( 9 8 ) 0 0 2 9 3 - X

1734

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

Fig. 1. Potential dependence of potentiodynamic growth of PPy lms. Nondegassed acetonitrile solution, 0.1 M pyrrole, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mV s 1. (a) 1.121.13 V, 15 scans; (b) 1.121.03 V, 12 scans; (c) 1.120.93 V, 20 scans; (d) 1.120.88 V, 28 scans; (e) 1.120.83 V, 40 scans, from 6th, every second scan recorded; (f) 1.12 0.83 V, 30 scans, partially recorded; scan range changed to 1.120.78 V for another 28 scans, every third scan recorded.

Cyclic voltammetry is an essential method for the characterization of the electrochemical properties of CPs. Although the shape of cyclic voltammograms

(CVs) has long been a subject of discussion [11, 13, 18, 22, 27, 31, 32], it is now accepted that the shape and position of CVs can reect proper-

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1735

ties of polymer structure. In the case of pyrrole, typical voltammograms of PPy electrosynthesized under mild conditions (current density at the level of mA cm 2 in the case of galvanostatic synthesis) show an oxidation wave with a steep onset, followed by a broad plateau and, on the reverse scan, a negatively shifted, weaker reduction wave [13]. The shapes may vary, depending on supporting electrolytes, solvents and other conditions. Seeking to establish the relation between applied potentials and voltammetric properties, we conducted electrochemical polymerization experiments with pyrrole in acetonitrile (0.1 M TBAPF6 and 1 wt% H2O) at potentials and/or currents, down to levels that barely sustained lm formation. The study not only results in the ndings of electrochemically well-dened PPy, but also unequivocally demonstrates that a small change in a single experimental variable may inuence the properties of the resulting polymer enormously, and that results obtained under widely varying conditions cannot be uncritically used to draw universal conclusions. 2. Experimental Pyrrole (Aldrich) was previously distilled and stored in the solid state in a freezer. Commercial HPLC grade acetonitrile (Fisons Scientic Equipment) was used as received. Added water (1.0 wt%) was deionized. The supporting electrolytes, used always at 0.1 M in both pyrrole-containing and pyrrole-free solutions, tetrabutylammonium hexauorophosphate (TBAPF6, Fluka, electrochemical grade), tetrabutylammonium triuoromethanesulfonate (TBACF3SO3, Fluka, electrochemical grade) and lithium perchlorate (LiClO4, Aldrich, anhydrous, 99.98%), were vacuum dried in an air-tight electrochemical cell immediately before use. Unless otherwise indicated, TBAPF6 was used as standard supporting electrolyte in the study. For the pyrroleand water-containing solutions, both nondegassing and degassing protocols were adopted. In the degassing protocol, pyrrole (twice distilled in argon), acetonitrile and water were purged with argon for more than half an hour and immediately introduced into the cell under the protection of argon. Pyrrole- and water-free solutions were passed through an alumina-lled column (ICN alumina B-Super I, vacuum dried at 3008C for 2 h immediately before use), till a good background in the required potential range was obtained. All electrochemical experiments were performed at 208C and under argon protection. The prepared solutions were normally used for a series of experiments, and were also stored under argon in an air-tight cell when not in use. A Pt disk (diameter 1 mm, area 0.785 mm2) sealed in a soft glass rod was used as working electrode; it

was polished with diamond (0.25 mm) polishing paste and then rinsed thoroughly with ethanol and acetone. Pt and Ag wires were used as counter and quasireference electrodes, respectively. Potentials versus the Ag quasireference electrode were then rescaled by Ag/ AgCl, calibrated with the ferrocene/ferrocenium redox couple (0.35 V vs. Ag/AgCl). An EG&G Potentiostat/ Galvanostat Model 273 and a Kipp and Zonen Delft BV BD 92 recorder were used for electrochemical control and data recording. During electrochemical process charges were read directly from the potentiostat/ galvanostat. 3. Results and discussions 3.1. The additional oxidation wave Our initial ndings are illustrated in Fig. 1. The potentiodynamic formation of PPy was studied in the potential range between 1.12 V and dierent positive switching potentials. As can be seen, when the switching potential shifted from 1.13 to 0.93 V, the cyclic voltammograms change their shape from almost symmetrical anodic and cathodic waves (Fig. 1a) to waves with a strong hysteresis between the anodic and the cathodic scans (Fig. 1c). In addition to the normal wave at E pH0.01 V, a new sharp oxidation wave emerged at potentials about 0.23 V as the switching potential was further lowered to 0.83 V (Fig. 1e). An even stronger current response of the additional wave could be observed when the scan was reversed at an even lower potential in the course of potentiodynamic polymerization (Fig. 1f). Contrary to the well-resolved oxidation waves, the reverse reduction half-cycle exhibited no additional wave. However, there was a progressive negative shift of the peak-like reduction wave when switching potentials were lowered [E pc = 0.25 and 0.33 V in Fig. 1(b) and Fig. 1(f), respectively]. Results displayed in Fig. 1 were obtained with solutions containing 0.1 M pyrrole at a temperature of 208C, but the same phenomena were observed over a wide concentration range of pyrrole (Fig. 2), a temperature range from 40208C and also with other supporting electrolytes, such as TBACF3SO3 and LiClO4 (Fig. 3). In Fig. 2, using solutions containing 0.001, 0.01, 0.1 and 0.5 M pyrrole, the potential sweeps were made with switching potentials adjusted as low as possible to sustain a minimal rate of lm growth. It was found that the new oxidation wave occurred over a concentration range from 0.010.5 M pyrrole. Even when the concentration was as low as 1 10 3 M, the trace of the new wave could be distinguished from the wide, symmetrical and positively shifted redox wave (Fig. 2a).

1736

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

Fig. 3. New oxidation wave of potentiodynamically grown PPy doped by (a) CF3SO3 , 1.020.88 V, 18 scans; and (b) ClO4 , 0.920.83 V, 38 scans, from the 5th, every third scan recorded. Nondegassed acetonitrile solution, 0.1 M pyrrole, 0.1 M supporting electrolyte, 1 wt% water, temperature 208C, scan rate 100 mVs 1.

It should be pointed out that the switching potentials required for the appearance of the additional wave changed with the pyrrole concentration. The lower the concentration of pyrrole, the higher the switching potential required. This represents the interchangeability between concentration and applied potential. When TBACF3SO3 and LiClO4 are employed as supporting electrolytes, the separation of the two anodic waves is
Fig. 2. Potentiodynamic growth of PPy lms at extremely low formation potentials in solutions of dierent pyrrole concentration. Degassed acetonitrile solutions, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mVs 1. (a) 0.001 M pyrrole, scan range 1.031.02 V, 130 scans, from the 10th scan, every 10th scan recorded; (b) 0.01 M pyrrole,

scan range 1.030.87 V, 145 scans, from the 25th scan, every 10th scan recorded; (c) 0.1 M pyrrole, scan range 0.960.82 V, 75 scans, from the 5th scan, every 5th scan recorded; (d) 0.5 M pyrrole, scan range 1.010.74 V, 80 scans, from the 5th scan, every 5th scan recorded.

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1737

smaller than in the case of TBAPF6 as supporting electrolyte, but the ne structure is apparently distinguishable (Fig. 3). The usually observed trace-crossing, the so-called `nucleation loop' [6, 3335] (visible in Fig. 1(a and b)), did not occur in all cases of electropolymerization at lower potentials (Fig. 1(cf), Figs. 2 and 3). In a later section of the paper, this type of trace-crossing will be discussed in more detail. The thickness of lms as measured approximately by the peak current of the redox waves was dierent: in Fig. 1, from (a) to (e) the lms got gradually thinner. This raises the question of whether lm thickness played a crucial role. Both our subsequent results and the quantitative comparison among the CVs in Fig. 1, or better, between Fig. 1(c) and Fig. 4(b), indicate that, though this new peak became indistinct when the cyclic potential scan continued and the lm turned thicker, the formation of the new peak was not the consequence of a dierence in lm thickness, but instead, depended on the switching potential. However, when the switching potential is set extremely low, the mechanisms of polymerization and lm growth perhaps change during the course of potential scan. Fig. 4 reveals that the change in CVs at low switching potential depends on the stage of the lm growth. While the charging current of the polymer increased with the number of scans, the position of the new peak shifted gradually to a more positive potential (Fig. 4(a and b)), leading eventually to an apparently wide wave without observable ne structure (Fig. 4c). It is also interesting to note that the position of the switching potential at the end of the cathodic scan also inuences the additional wave of the growing polymer lm. As shown in Fig. 4(a), each time the reverse scan was switched at a more negative potential (initially at 0.64 V, then successively, 0.94, 1.34 and 1.74 V), the next oxidation scan gave rise to an immediate current jump in the new peak. But no such eect was found in the normal wave at 0.01 V. It is thus suggested that the corresponding reduction of the polymer with the additional oxidation peak must be completed at a more negative potential. 3.2. Properties of galvanostatically synthesized lms Thicker PPy lms with only one peak-like oxidation wave at a potential less positive than normal can be synthesized galvanostatically at extremely low current densities, for example ca. 6 mA cm 2 (50 nA on our electrode), which is about a few hundreds times lower than under conventional conditions [3640]. In the case of repetitive charging and discharging during a potentiodynamic multisweep in a monomer-free solution, the initially synthesized PPy lm may be structurally modied, as evidenced by the evolution of the

Fig. 4. Potentiodynamic growth of PPy lm: inuence of discharging extent and thickness. Nondegassed acetonitrile solution, 0.1 M pyrrole, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mVs 1, up-switching potential 0.86 V. (a) Discharging scan down to dierent potentials in the rst 46 scans; (b) 5080th scans, every second scan recorded; (c) 250th and 335th scans. Figures indicate the scan number, dashed lines indicate the initial position of oxidation waves.

CV (Fig. 5a). Upon repetitive cyclic potentiodynamic scans between 0.75 and 2.1 V, the peak at 0.23 V gradually grew and shifted positively little by little. Eventually, the structurally evolved polymer lm showed completely reversible redox CVs, with a sharp oxidation wave at 0.09 V, followed by a very low pla-

1738

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

Fig. 5. Evolution of voltammograms of PPy lms in monomer-free acetonitrile solution. 0.1 M TBAPF6, temperature 208C, scan rate 100 mV s 1, scan range 0.75 2.1 V. Film prepared in solution same as that in Fig. 1, galvanostatically at 50 nA, with charge 720.8 mC. Figures indicate the scan number.

teau, and very stable electrochemical properties (Fig. 5b), with no more voltammetric changes during subsequent hours of repetitive potential cycles. Note also that the oxidation peak current was about four times as high as that of the rst cycle. This implies that the as-prepared lm was much less charged than it could be. Such a dramatic evolution of the charging/ discharging properties reected by the nal CV (Fig. 5b), which reveals some structure-related reactions in the solid state during the repetitive redox process, has never been reported before. The studies on the evolution of CVs from lms prepared at dierent currents clearly demonstrated a continuous change of voltammetric properties akin to the situation in Fig. 1. These are illustrated in Figs. 6 and 7. The current used to produce the lm in Fig. 5 was 50 nA, i.e. ca. 6 mA cm 2. Increasing the current to 0.1 mA (i.e. 12.7 mA cm 2), the resulting polymer exhibited an obviously composite oxidation wave, and the more negative component became less dominant at even higher currents (Fig. 6). In Fig. 6(b and c), the

composite feature of the oxidation wave is not so clear as in Fig. 6a. However, upon a couple of cyclic potential scans in pyrrole free solution, the composite feature was disclosed by the decomposition of the oxidation wave due to the dierent evolution behaviors of the two waves in monomer free solution. The results in Figs. 5 and 6 convince us that the apparent oxidation wave of PPy normally prepared at low current densities, or low potentials (in controlled potential mode) is a combination of oxidation processes of at least two dierent PPy variants. The evolution behaviors of these two variants are dierent. In cyclic potential scan the more negative wave tended to grow, while the more positive wave tended to diminish, as the arrows and the scan number in the gure indicate. Fig. 6 shows the transition from the state displayed in Fig. 5 to the normal voltammetric properties. As the applied current further increased, another transition, namely from peak-plateau form to symmetric form, appeared. This is demonstrated in Fig. 7, where the

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1739

current was increased from 50 to 100 mA [Fig. 7(ad), current density from ca. 6.4 to ca. 12.7 mA cm 1]. Since a higher current level corresponds to a higher anodic potential, the tendency reected in Figs. 57 is in good agreement with that in the potentiodynamic cases shown in Fig. 1(af). In addition to the change in basic features of the CVs, another two aspects should be mentioned. First, the positions of the wave onset and the current maximum shift negatively with decreasing syn-

thesis current. The shift can be clearly observed by means of the dashed lines in Figs. 6 and 7. Second, in Fig. 7, to keep peak current at the same level, less charge was needed for the synthesis at a lower current. At a higher current, more charge was needed to elevate the plateau. Our investigation of the structural evolution of PPy in a monomer-free solution showed that electropolymerization is a very subtle process. A very small change in conditions (e.g. 10% change of current

Fig. 6. Evolution of voltammograms of PPy lms in monomer-free acetonitrile solution. 0.1 M TBAPF6, temperature 208C, scan rate 100 mVs 1, scan range 0.75 2.1 V. Films prepared in solution same as that in Fig. 1, with current/charge (a) 0.1 mA/716.9 mC; (b) 0.2 mA/721.2 mC; (c) 0.4 mA/721.4 mC. Figures indicate the scan number.

1740

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

details of CVs are not always reproducible when extremely low currents are employed. On the other hand, this just reects the susceptibility of the electropolymerization process to the conditions and the diversity of the structures of the resulting polymer. The change and evolution of CVs in monomer-free solution might be the result of some solid-state reactions and ion movement during charging and discharging. The relaxation eect [41, 42], as applied to polythiophene for explaining CV change, is probably too simple for our case. At present, we are unable to interpret all the phenomena reported above. In the following section we shall try to explain only some of the observations. 3.3. The real factor From the shape and separation of the two oxidation waves in Fig. 1 and other features in Figs. 47, we believe that the phenomenon related, should be attributed not to the stepwise electron transfer process, but rather to the dierent structural entities coexisting in the polymer matrix. Lei and Martin studied the reaction of pristine PPy (i.e. in its neutral state) with oxygen [43]. The CVs they obtained showed a negative shift of both the charging and discharging waves from O-doped PPy. In our ndings, the additional wave also lies negative to the normal charging wave (i.e. 0.23 and 0.01 V, respectively). Could the additional wave come from a species of some state of O-doped PPy that might have been formed in solution containing O2 when PPy was constantly switched between the charging and discharging states? Our experiments with solutions, well purged with argon, indicated that this was not the case. The additional wave can also appear in completely oxygenfree solutions once the switching potentials have been adjusted to very low levels in the potentiodynamic experiments. As a matter of fact, the CVs in Fig. 2 were all obtained from the solutions whose components (acetonitrile, TBAPF6, pyrrole, and water) were intensively degassed immediately before use. The comparison between Fig. 1(e) and Fig. 2(c) gives no support to the assumption of a possible incorporation of oxygen. The inuence of applied potential in the electropolymerization of conducting polymers is usually estimated from the viewpoint of charging levels (i.e. monocation, dication, etc.) and reactivity. The studies on oligomers of thiophene, pyrrole and phenylene have already demonstrated that the reactivity of oligomer cations increases with a higher charging level and decreases with increasing chain length [21, 22, 27, 44, 45]. This was proved by the solid-state electropolymerization experiments [27]. Another well-proved fact is that the oxidation potentials of oligomers vary linearly with the inverse of the number of repeating units. That is to

Fig. 7. Voltammograms of PPy lms prepared at relatively higher currents. 0.1 M TBAPF6, temperature 208C, scan rate 100 mV s 1, scan range 0.75 1.1 V. Films prepared in solution same as that in Fig. 1, with current/charge (a) 50 mA/777.2 mC; (b) 70 mA/1.116 mC; (c) 80 mA/1.493 mC; (d) 100 mA/2.277 mC. Figures indicate the scan number.

intensity alone, as in Fig. 7, or a slight dierence of solution aging) could give rise to an observable discrepancy between CVs. Thus, the analysis and characterization of structure and properties of conducting polymers, based on limited samples, may not be universally correct. At even lower current intensity, waves appear at even more negative potentials (one pre-wave in Fig. 5a, already present in the rst scans, subsequently merges into the main peak-like wave), and all these even more negative waves displayed instability when the lms underwent repetitive potential scans in monomer-free solution. It is worth noting that the

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1741

say, with a slight increase of the formation potential (or current) in potentiodynamic (or galvanostatic) electropolymerization experiments, a negative shift of the polymer oxidation wave should be expected, due to a longer chain length. Surprisingly, our results from both potentiodynamic and galvanostatic experiments showed the opposite. At extremely low formation potentials, the additional peak-like wave emerges at 0.23 V, which is more negative than those of `polypyrrole' obtained electrochemically from pentapyrrole and heptapyrrole [19], and those extrapolated values from Zotti et al. [19] (0.59 V vs. Ag/0.1 M Ag + or 0.205 V vs. Ag/AgCl) and from Andrieux et al. [45] (0.21 V vs. SCE, or 0.165 V vs. Ag/AgCl) for the `innite' oligomer pyrrole. It must be pointed out that both values were extrapolated from measurements of pyrrole oligomers in solution. In the case of solid state experiments, a strong potential hysteresis eect was normally observed for many conducting polymers, i.e. the charging process always requires more energy than is expected from extrapolation of solution data, whereas the discharging process needs less energy than extrapolation indicates. Since such a hysteresis eect probably holds for polypyrrole, we should expect that the oxidation potential for the innite oligomer pyrrole (polypyrrole) to be more positive than 0.205 or 0.165 V vs. Ag/AgCl. The deviation from our observation of 0.23 V thus is even greater. If we attribute the sharp wave at 0.23 V to the structural entity with longer chains, the inapplicability of the potential/chain length relationship to this case can be explained as follows. Because the potentials required to charge the oligomer to dierent levels are discrete, changing potential between two neighboring values cannot result in the more reactive species. In our experiments, the change of potential (or current) was probably not big enough to generate change in the charging level of oligomer radical cations. Therefore, the potential alone may not be a direct reason for the appearance of the sharp wave at E pa = 0.23 V. Additional factors must be taken into account for this process. One should bear in mind that decreasing the switching potential in a potentiodynamic sweep, or current in a galvanostatic experiment, can substantially decrease the generation rate of monomer cations and hence eventually lower the polymerization rate and lm growth rate. As reported above, we focused on the potential range from the oxidation onset to the peak potential. A very important fact is that a slight shift in the switching potential can cause a tremendous change in the reaction rate. Because monomer concentration, temperature and potential can be used interchangeably to reach a dened reaction rate, the potentials required to sustain the minimal formation of polymer are not xed; higher

potentials are needed for a lower monomer concentration (see Fig. 2). Likewise, in experiments with variable temperatures, a higher potential is needed for the sustainable formation of polymer at lower temperatures. With respect to the appearance of the additional peak-like wave, it appears that the crucial factor was none of these, but instead, the generation rate of radical cations, which was represented by the current level of monomer oxidation. Being a second-order reaction, the dimerization coupling rate of radical cations depends strongly on the concentration of the reactants, which are not the neutral monomers, but rather, the radical cations generated in the rst electron transfer step. In this context, it is apparent that the oxidation current, the reaction rate and the generation rate of all charged species have the same implications for our discussion of polymerization rate. According to the normally accepted polymerization mechanism, the dimerization coupling is followed by the elimination of two protons, which may change the acidity within the diusion layer as a function of the reaction rate. If it is so high that proton generation is faster than proton diusion from the electrode to the bulk solution, the reaction zone becomes rich in protons. If the reaction rate is extremely low, the relatively faster diusion of protons leads only to a minor proton gradient between the reaction zone and bulk solution. Thus, one of the signicant dierences, caused by varying the reaction rate, is the acidity of the actual reaction environment. The inuences of pH on the pyrrole electropolymerization process in aqueous solutions and on the properties of the resulting polymer have been discussed in the literature [4648]. In addition to protons, other intermediates, originating from many possible reactions and changing with the reaction rate, may also contribute to the reaction environment. As far as solid lms are concerned, the specic nature of the solid state, such as the electrical resistivity, capacitive eect, lattice relaxation, swelling properties, and the related permeability to ions and solvent, etc., must be taken into consideration in the interpretation of the voltammetrical behavior of conducting polymers [18]. These properties depend not only on the polymeric chain structure (chain-length, chain-branching, cross-linking), but also, to a great extent, on the tertiary structure, namely, the stacking mode of chains. The properties of polymers formed in dierent reaction environments may dier from one another. Such complications associated with the solid state may constitute one of the reasons for the shift and shape change of CVs. From the viewpoint of deposition and crystallization, a slow formation process favors a better crystallinity, resulting in polymer deposition dierent from

1742

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

that at higher rates. Thus, our observation of the additional wave (Fig. 1f) agrees with the fact that wellcrystallized electroactive substances show better-shaped and sharp waves. One of the referees of this paper proposes that the wave at 0.23 V may be due to a PPy with a higher conjugation length, while the wave at 0.0 V should be connected with the charging of a PPy matrix with a lot of `defects', such as interchain cross linking. At this level of discussion, such an explanation seems to be very reliable. Anyway, the above hypotheses remain to be proved and related phenomena deserve further investigation. Very recent studies have shown that upon electropolymerization of pyrrole in the dependence on experimental conditions dierent types of PPy can be generated. We will discuss in forthcoming papers [49] the mechanistic pathways leading to these dierent materials. It should be noted that the negative shift of PPy was also observed under other circumstances. Using 2,2 H bipyrrole as starting substance, Zotti et al. [50] found that anodic coupling in acetonitrile was inuenced by the electron-donor properties of the anion of the supporting electrolyte. Compared to electron-poor donor anions (ClO4 , PF6 , BF4 and CF3SO3 ), electron-rich donor anions (tosylate, benzenesulfonate, methanesulfonate, camphorsulfonate, chloride, and nitrate) favor polymerization, and lead to the resulting polymer showing a more negative redox pair. For example, the PPy doped by tosylate has a reversible oxidation process at E0 = 0.70 V (vs. Ag/0.1 M Ag + ), in comparison with 0.45 V (vs. Ag/0.1 M Ag + ) for the PPy doped by perchlorate. Furthermore, the UV-VIS spectrum and conductivity data indicate that tosylate favors the deposition of a long-chain conjugated polymer. This result supports the view that in some cases in addition to potential, other factors may play a crucial role in the formation of chain structure. What is interesting is that the potential dierence in the above example (0.25 V) is, coincidentally, almost the same as the separation of the two waves in our case (0.24 V). 3.4. The `nucleation loop' In the voltammograms of conducting polymers, a frequently mentioned feature of the rst potential cycle is the `nucleation loop' or `trace-crossing', which appears when a freshly polished electrode is used and the scan is reversed at a potential close to the peak potential. Such behavior is similar in respect to the electrochemical deposition of metal on a foreign substrate, in which an overpotential is required for nucleation and then further growth of the metallic layer occurs at the characteristic redox potential of the metal, leading to a trace-crossing on the reverse sweep. It is proposed that in the deposition of conducting polymers the

Fig. 8. Trace-crossing occurring on polymer-coated electrode. Degassed acetonitrile solution, 0.1 M pyrrole, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mV s 1. Scan range (a) 0.960.99 V; (b) 0.961.09 V; (c) 0.961.19 V. Figures indicate the scan number.

nucleation process also needs an overpotential, resulting in a higher current on the reverse scan in a certain potential range. In most of our cases, in which the switching potentials were set lower than the peak potential, no tracecrossing was observed [see Fig. 1(cf), Figs. 24]. However, the growth of polymer lms did occur. Although there is no doubt that the heterogeneous deposition is initiated by the nucleation, the nucleation process does not appear to be necessarily followed by the trace-crossing. On the other hand, if the trace-crossing in the rst cycle is attributed to the nucleation process, it should not be observed in the second and the following cycles. However, when switching potentials were set close to the peak potential, we observed the so-called `nucleation loops' even after several cycles, when the electrode was totally covered with polymer. As we can see in Fig. 1(a), the loop is still present in at least the second cycle. Besides, further experiments revealed that the tracecrossing could return after disappearing. In Fig. 8 are the results from the potential scans, carried out initially between 0.960.99 V (Fig. 8a). The trace-cross-

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1743

Fig. 9. Trace-crossing in dierent cases (rst two scans). Degassed acetonitrile solutions, 0.1 M TBAPF6, 1 wt% water, temperature 208C. (a) 0.001 M pyrrole, 100 mVs 1, 0.531.37 V; (b) 0.01 M pyrrole, 100 mV s 1, 0.531.17 V; (c) 0.01 M pyrrole, 20 mVs 1, 0.531.17 V; (d) 0.1 M pyrrole, 100 mVs 1, 0.531.07 V; (e) 0.1 M pyrrole, 20 mVs 1, 0.531.07 V; (f) 0.5 M pyrrole, 100 mV s 1, 0.531.07 V; (g) 1.0 M pyrrole, 100 mV s 1, 0.531.17 V.

ing occurred in the rst cycle, as it usually does. Although the situation in the second cycle was equivocal in this case the forward trace and the backward trace overlapped in the potential range of interest from the third cycle on, the trace-crossing was denitely absent. After the sixth cycle, as the switching potential was raised to 1.09 V in the seventh cycle, the trace-crossing reappeared, as can be seen in Fig. 8(b). Moreover, when the switching potential was further increased to 1.19 V in the eighth cycle, the trace-crossing occurred once again (Fig. 8c). It is of great interest that the electrode was already completely coated with a polymer lm after six potential cycles in the region of 0.960.99 V.

The facts that polymer lm can grow without the initial `nucleation loop', and that the `nucleation loop' can emerge even after the electrode becomes completely polymer-coated, are inconsistent with the interpretation based on the nucleation-and-growth mechanism [3335] of conducting polymers. Thus, the nature of the trace-crossing is a matter that still needs to be explained. Further experiments indicated that the occurrence of trace-crossing depends on several factors. The CVs in Fig. 9 show the rst two cycles in solutions of dierent monomer concentrations. At an extremely low pyrrole concentration (0.001 M), neither the rst nor the

1744

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

Fig. 10. Repetitive sharp oxidation wave on the backward scans. Degassed acetonitrile solution, 0.5 M pyrrole, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mV s 1, 1.011.09 V.

follow-up cycle gave rise to the trace-crossing (Fig. 9a). What is more, the current levels of the second and the follow-up scans were, at the switching potential, less than that of the rst scan. At a higher pyrrole concentration (0.01 M), we observed the normal behavior, namely, the trace-crossing appeared in the rst forward-reverse cycle, but disappeared in the second cycle, when the scan rate was set at 100 mV s 1 (Fig. 9b). However, if the scan rate was changed from 100 to 20 mV s 1, the trace-crossing appeared in both the rst and second cycles (Fig. 9c). There is no doubt that more nucleation centers and deposit were formed in the case of slower scan rate (Fig. 9c). When the pyrrole concentration was even higher (0.1 M), tracecrossing took place in the rst and second cycles at both scan rates of 100 and 20 mVs 1 (Fig. 9(d and e)); and the lower scan rate resulted in the more obvious crossing of traces (Fig. 9e). Our experiments with a monomer concentration of 0.5 M enabled us to gain more insight into the nature of the trace-crossing. In Fig. 9(f), the CV is conspicuous by a strange peaklike wave (indicated as P1) on the second reverse scan at about 0.96 V. It remained on the subsequent reverse scans (Fig. 10). The nature of this peak is yet to be explained. Nevertheless, it leads us to believe that such a strong oxidation reaction on the reverse scan is at

Fig. 11. Studies of the oxidation wave on the backward scans from solution or from newly formed polymer? Degassed acetonitrile solution, 0.1 M TBAPF6, temperature 208C, scan rate 100 mV s 1. (a) and (c) 0.5 M pyrrole, 1 wt% water, 0.511.09 V. The scan stopped at the switching potential 1.09 V on the third forward scan. (b) After (a), the same electrolytic solution was shaken, 1.09 0.51 V. (c) After (c), in pyrrole-free solution, 1.09 H 0.51 V.

least one, if not the only, reason for the higher current on the reverse scan. Although the P1 can be more easily observed in solutions with a higher concentration

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1745

of monomers, increasing the monomer concentration did not always make the phenomenon more obvious, as shown in Fig. 9(g), where monomer concentration was as high as 1.0 M.

Due to simultaneous polymer deposition on the electrode, one may legitimately ask whether the P1 on the reverse scan [Fig. 9(f) and Fig. 10] comes from the oxidation of any species in the solution or from the newly deposited polymer (or solid oligomer) on the electrode. The answer to this question was obtained from the experiments in which the potential sweep was stopped at the switching potential 1.09 V (Fig. 11(a and c)] on the third forward scan, and then further investigations were made in two protocols. In one protocol, the cell was shaken for a while so that the reaction zone around the electrode became the same as the bulk solution. Then, the potential scan was restarted from 1.09 V to the negative side. The P1 peak disappeared in this case (Fig. 11b). This is solid support for assigning the peak to the oxidation of species in the solution. In another protocol, after the third potential scan was stopped at 1.09 V (Fig. 11c), the electrolytic solution in the cell was poured out. The cell was washed several times with pure acetonitrile and 0.1 M TBAPF6 acetonitrile solutions and then relled with the pyrrole-free solution. In all the operation steps, the polymer-coated electrode stayed in the Ar-protected cell without any possible contact with atmosphere. The new potential scan was then made in the pyrrole-free solution from 1.09 to 0.51 V. The result shown in Fig. 11(d) indicates again unequivocally that the strange peak in question is not from the existing polymer.

Fig. 12. Studies of the oxidation wave on the backward scans: potential scan starting from the positive side. Degassed acetonitrile solution, 0.5 M pyrrole, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mV s 1. (a) 1.39 0.51 1.09 V; (b) 1.29 0.511.09 V; (c) 1.19 0.511.09 V; (d) 1.09 0.51 V.

Fig. 13. Studies of the oxidation wave on the backward scans: the independence of the peak on the scan direction. Degassed acetonitrile solution, 0.5 M pyrrole, 0.1 M TBAPF6, 1 wt% water, temperature 208C, scan rate 100 mV s 1. First two scans: 0.511.09 V; third scan: 0.511.24 V.

1746

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

More information pertaining to the P1 peak, was found in the results shown in Fig. 12, in which the potential sweep was initiated at potentials higher than 1.0 V. When the starting potentials were set at 1.39, 1.29 and 1.19 V, a small peak (P2) and a broad wave were always observed after the initial quick drop of current on the rst scan. Then, after the scan was switched at 1.09 V, the P1 peak appeared as it usually does. There is no doubt that the initial quick current drop is due to the depletion of monomer in the reaction zone. P2 and the accompanying broad wave are the consequence of follow-up reactions. Since the P2 peak appears after a short reaction period and has no denitive peak potential value, we believe that P1 is identical with P2 in nature. At a dened scan rate, once a certain amount of monomer is oxidized, the follow-up electrochemical reactions will take place, showing the P2 (or P1) peaks and the broad wave as well. In this context, such a peak should also be observed on the forward scan from lower potential to higher potential, if the scan is not switched back too early. This is proved by Fig. 13, in which the peak (indicated as P3) on the third forward scan is actually the P1 on the second backward scan. The concentration of 0.5 M pyrrole provides a better observation of the anodic current from the follow-up reaction. Under other conditions, the follow-up reactions in pyrrole polymerization process is also certain, but whether a sharp peak, like P1, P2 and P3, and the trace-crossing emerge depends on the conditions discussed in Fig. 9. In conclusion, the electrochemical follow-up reactions give rise to an additional anodic current. If the scan is reversed in the time scale in which the followup reactions occur, the current on the backward scan will be possibly higher than the current on the former immediate forward scan in certain potential range. This is one of the reasons for the trace-crossing in potentiodynamic electropolymerization, at least in the case of pyrrole. Early in 1970s, Feldberg [51] demonstrated theoretically the possibility of trace-crossing in certain ECEtype reactions. Some experimental examples of tracecrossing in the voltammograms of organic and organometallic reactants can be found in the literature [52 58]. In his review, Evans [59] also discussed in detail trace-crossing in an ECE reaction scheme involving electron-transfer reactions in solution. In all these examples the trace-crossing has nothing to do with any deposit on the electrode. Very recently, it was shown that during the electropolymerization of methylthiothiophenes trace crossing results from the slow formation of a neutral tetramer in the diusion layer at the electrode and its subsequent oxidation to its mono- and dication. Although the oxidation of the monomeric starting species stops in the reverse scans,

the tetramer is still formed on account of the slow kinetics of the dimer-dimer coupling in front of the electrode and, therefore, an anodic current, generated by the mono- and dication formation of the tetramer, can be observed at potentials negative to that of the `monomer' oxidation [60]. While discussing trace-crossing, we should point out another feature of CVs that many researchers of conducting polymers tend to ignore. In the potentiodynamic growth of conducting polymers, the current at a certain oxidation potential increases from cycle to cycle; in other words, the potential required to reach a certain oxidation current decreases with each successive scan. From general observations and experiments using electrodes coated with dierent thickness of the same doped PPy lm, we established that the oxidation current at a dened potential was dependent on the amount of the existing polymer lm on the electrode. The thicker the existing polymer deposition is, the larger the current of monomer oxidation. One subject of speculation is the increase of the electrode surface available for electron transfer. At the early stage of deposition, the initial nuclei and the roughness of the polymer formed, actually increase the electrode-solution interface available for electron transfer, and hence contribute to the enhancement of monomer oxidation. However, the nuclei and surface roughness is signicant only in the rst few cycles. They do not play any role once the electrode is completely covered with polymer. Whatever deposition mechanism is considered, it is very dicult to imagine the electrode area always on the increase. Taking the already deposited polymer into consideration, we propose a mechanism in which oxidation takes place not only on the solution-electrode interface, but also in the polymer matrix. There is evidence for the swelling ability of PPy in acetonitrile [61] and the swollen polypyrrole matrix may contain pyrrole molecules. Thus, the oxidation of pyrrole molecules in polymer matrix is likely, especially on the newly formed surface layer. In the potential scan experiments, the contribution to the anodic current from interface oxidation is almost constant after the electrode is entirely polymer-coated and surface state no longer changes, whereas the contribution to the anodic current from matrix oxidation increases with deposition. The thicker the polymer layere, the greater the monomer amount in the polymer matrix, and the stronger the total oxidation current, will be. Accordingly, the current in the range of monomer oxidation always increases with the number of scans. However, due to the limits imposed by the swelling prole and the increasing resistance of deposit, the increment in the current at the switching potential, after subtraction of the ever-growing current plateau, becomes smaller and smaller.

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748

1747

The above two cases of monomer oxidation can also be used to elucidate trace-crossing. As explained above, monomer molecules diuse into the polymer matrix due to the swollen structure of the polymer lm produced on the forward scan. This enhances the monomer oxidation on the reverse scan, producing a current higher than that on the forward scan in the same cycle. Because the relative increase in the amount of polymer on the initial scans is bigger than that on later scans, trace-crossing occurs in the rst cycles and disappears in the later cycles. The monomer oxidation in the polymer matrix can explain why there is no trace-crossing when the switching potential or monomer concentration is very low, and why, after disappearing, trace-crossing returns at higher switching potentials (Fig. 8). If the relative change in the amount of deposited polymer after each scan is too small (as in the cases of low switching potential and low monomer concentration, or when the existing polymer lm is already thick), the monomer oxidation in the matrix cannot make a substantial contribution to the increase in total oxidation. Raising the switching potential can generate a bigger relative change in the amount of polymer and increase the contribution of monomer oxidation in the polymer matrix, leading to a higher current in the immediate reverse scan. The overall electrochemical polymerization of conducting polymers comprises complex homogeneous and heterogeneous reactions and processes, and a net electrochemical response may be the result of many factors. We do not rule out the nucleation-and-growth mechanism as an explanation for trace-crossing. However, it cannot explain all the facts described above. Though we need further study, we feel that the above suggestions, i.e. the possible oxidation of oligomers to higher charging levels, the increase in the electrode-solution interface at the initial stage of deposition and the swollen structure of polymer lm, can improve our understanding of trace-crossing. 4. Conclusion Both the potentiodynamic and galvanostatic syntheses of polypyrrole at extremely low reaction rates enable us to nd an additional oxidation wave at ca. 0.23 V, which is more negative than the normal oxidation wave of polypyrrole. The well-dened polymer showing only the sharp oxidation peak at 0.23 V can be galvanostatically synthesized at extremely low current levels. Although the potential determines the charging levels of oligomer radical cations and, accordingly, inuences the chain-length of the resulting polymer, it does not necessarily follow that changing the potential in a certain range can always bring about

the expected change in polymer structure. In some cases, the change of chemical environment due to dierent reaction rates may have a more profound impact on the properties of resulting polymers. We have displayed many dierent shapes of CVs of PPy doped PF6 . Hence, one point our experiments unambiguously established is the structural diversity of PPy. It must be emphasized that in the studies of conducting polymers the results obtained under widely varying conditions cannot be uncritically used to draw universal conclusions, for a small change in a single experimental variable may greatly inuence the properties of the resulting polymer. We hope the opinions stated here will shed some light on some controversial issues. The nucleation-growth mechanism can explain tracecrossing in the rst potential cycle, but not all tracecrossing phenomena. The follow-up reactions and the swelling property of existing polymer should also be taken into account in interpreting trace-crossing.

Acknowledgements MZ is grateful to the Alexander von Humboldt Foundation for an Alexander von Humboldt Research Fellowship. The nancial support of the Fonds der Chemischen Industrie is gratefully acknowledged.

References
[1] A.F. Diaz, K.K. Kanazawa, J.P. Gardini, J. Chem. Soc. Chem. Commun. 635 (1979. [2] A.F. Diaz, Chem. Scr. 17 (1981) 142. [3] G. Tourillon, F. Garnier, J. Electroanal. Chem. 135 (1982) 173. das, R.R. Chance, R. Silbery, Phys. Rev. B26 [4] J.L. Bre (1982) 5843. [5] E.M. Genies, G. Bidan, A.F. Diaz, J. Electroanal. Chem. 149 (1983) 113. [6] S. Asavapiriyanont, G.K. Chandler, G.A. Gunawardena, D. Pletcher, J. Electroanal. Chem. 177 (1984) 229. [7] P. Puger, G.B. Street, J. Chem. Phys. 80 (1984) 544. das, G.B. Street, Acc. Chem. Res. 18 (1985) 308. [8] J.L. Bre [9] J. Roncali, F. Garnier, M. Lemaire, R. Garreau, Synth. Met. 15 (1986) 323. [10] T.A. Skotheim (Ed.), Handbook of Conducting Polymers, Marcel Dekker, New York, 1986. [11] J. Heinze, M. Sto rzbach, J. Mortensen, Ber. Bunsenges. Phys. Chem. 91 (1987) 960. [12] F. Devreux, F. Genoud, M. Nechtschein, B. Villeret, Synth. Met. 18 (1987) 89. [13] J. Heinze, Top. Curr. Chem. 152 (1990) 1. [14] J. Roncali, Chem. Rev. 92 (1992) 711. [15] P. Ba uerle, Adv. Mater. 4 (1992) 102. [16] M.G. Hill, K.R. Mann, L.L. Miller, J.-F. Penneau, J. Am. Chem. Soc. 114 (1992) 2728.

1748

M. Zhou, J. Heinze / Electrochimica Acta 44 (1999) 17331748 [39] B. Wehrle, H.-H. Limbach, J. Mortensen, J. Heinze, Synth. Met. 38 (1990) 293. [40] R.G. Davidson, L.C. Hammond, T.G. Turner, A.R. Wilson, Synth. Met. 81 (1996) 1. [41] C. Odin, M. Nechtschein, Synth. Met. 44 (1991) 177. [42] C. Odin, M. Nechtschein, P. Hapiot, Synth. Met. 47 (1992) 329. [43] J. Lei, C.R. Martin, Chem. Mater. 7 (1995) 578. [44] Z. Xu, D. Fichou, G. Horowitz, F. Garnier, J. Electroanal. Chem. 267 (1989) 339. [45] C.P. Andrieux, P. Hapiot, P. Audebert, L. Guyard, M. Nguyen Dinh An, L. Groenendaal, E.W. Meijer, Chem. Mater. 9 (1997) 723. [46] W. Wernet, M. Monkenbusch, G. Werner, Mol. Cryst. Liq. Cryst. 118 (1985) 193. [47] R. Qian, Q. Pei, Z. Huang, Makromol. Chem. 192 (1991) 1263. [48] Q. Pei, R. Qian, J. Electroanal. Chem. 322 (1992) 153. [49] M. Zhou, J. Heinze, in prepperation. [50] G. Zotti, G. Schiavon, S. Zecchin, Chem. Mater. 7 (1995) 1464. [51] S.W. Feldberg, J. Phys. Chem. 75 (1971) 2377. [52] M.A. Fox, R. Akaba, J. Am. Chem. Soc. 105 (1983) 3460. [53] J.G. Gaudiello, T.C. Wright, R.A. Jones, A.J. Bard, J. Am. Chem. Soc. 107 (1985) 888. [54] M. Dietrich, J. Heinze, H. Fischer, F.A. Neugebauer, Angew. Chem. 98 (1986) 999. [55] M. Dietrich, J. Heinze, H. Fischer, F.A. Neugebauer, Angew. Chem. Int. Ed. Engl. 25 (1986) 1021. [56] R.D. Moulton, D.J. Chandler, A.M. Arif, R.A. Jones, A.J. Bard, J. Am. Chem. Soc. 110 (1988) 5714. [57] D.J. Kuchynka, J.K. Kochi, Inorg. Chem. 27 (1988) 2574. [58] J. Heinze, M. Dietrich, K. Hinkelmann, K. Meerholz, F. Rashwan, DECHEMA-Monogr. 112 (1987) 61. [59] D.H. Evans, Chem. Rev. 90 (1990) 739. [60] A. Smie, A. Synowczyk, J. Heinze, R. Alle, P. Tschuncky, G. Go tz, P. Ba uerle, J. Electroanal. Chem., 452 (1998) 87. [61] M. Zhou, M. Persin, J. Sarrazin, J. Membrane Sci. 117 (1996) 303.

[17] M.G. Hill, J.-F. Penneau, B. Zinger, K.R. Mann, L.L. Miller, Chem. Mater. 4 (1992) 1106. [18] J. Guay, P. Kasai, A. Diaz, R. Wu, J.M. Tour, L.H. Dao, Chem. Mater. 4 (1992) 1097. [19] G. Zotti, S. Martina, G. Wegner, A.D. Schlu ter, Adv. Mater. 4 (1992) 798. [20] P. Ba uerle, U. Segelbacher, A. Maier, M. Mehring, J. Am. Chem. Soc. 115 (1993) 10217. [21] G. Zotti, G. Schiavon, A. Berlin, G. Pagani, Chem. Mater. 5 (1993) 430. [22] K. Meerholz, H. Gregorius, K. Mu llen, J. Heinze, Adv. Mater. 6 (1994) 671. [23] P. Audebert, J.-M. Catel, G.L. Coustumer, V. Duchenet, P. Hapiot, J. Phys. Chem. 99 (1995) 11923. [24] J.A.E.H. van Harre, L. Groenendaal, E.E. Havinga, R.A.J. Janssen, E.W. Meijer, Angew. Chem. 108 (1996) 696. [25] A. Smie, J. Heinze, Angew. Chem. Int. Ed. Engl. 36 (1997) 363. [26] G. Schopf, G. Komehl, Adv. Polym. Sci. 129, 1997, Springer-Verlag, Berlin, 1997. [27] K. Meerholz, J. Heinze, Electrochim. Acta 41 (1996) 1839. [28] G. Zotti, G. Schiavon, A. Berlin, G. Pagani, Chem. Mater. 5 (1993) 620. [29] G. Zotti, R. Salmaso, M.C. Gallazzi, R.A. Marin, Chem. Mater. 9 (1997) 791. [30] K. Meerholz, J. Heinze, Synth. Met. 55-57 (1993) 5040. [31] S.W. Feldberg, J. Am. Chem. Soc. 106 (1984) 4671. [32] J.F. Oudard, R.D. Allendoerfer, R.A. Osteryoung, J. Electroanal. Chem. 241 (1988) 231. [33] A.J. Downard, D. Pletcher, J. Electroanal. Chem. 206 (1986) 139. [34] A.J. Downard, D. Pletcher, J. Electroanal. Chem. 206 (1986) 147. [35] R.E. Noftle, D. Pletcher, J. Electroanal. Chem. 227 (1987) 229. [36] A.F. Diaz, J.I. Castillo, J.A. Logan, W.-Y. Lee, J. Electroanal. Chem. 129 (1981) 115. [37] L.F. Warren, J.A. Walker, D.P. Anderson, C.G. Rhodes, J. Electrochem. Soc. 136 (1989) 2286. [38] R. Yang, D.F. Evans, L. Christensen, W.A. Hendrickson, J. Phys. Chem. 94 (1990) 6117.

Das könnte Ihnen auch gefallen