Sunteți pe pagina 1din 7

ELSEVIER

SyntheticMetals 87 (1997) 53-59

Stability of n-type doped conducting polymers and consequences for polymeric microelectronic devices
D.M. de Leeuw *, M.M.J. Simenon, A.R. Brown, R.E.F. Einerhand
Philips Research Laboratories, Prof: Holstlaan 4, 5656 AA Eindhoven, Netherlands

Received11 October 1996;accepted12November1996

Abstract Present polymeric microelectronic devices are typically unipolar devices, based on p-type semiconducting polymers. Bipolar devices stable under ambient conditions are desirable, but have not yet been reported due to a lack of stable n-type doped conducting polymers. Starting from the standard redox potentials of, especially, water and oxygen, stability requirements on electrode potentials of n-type doped conducting polymers are derived. The predictions are then compared with experimental data on stability of conducting polymers. A good agreement is obtained. An electrode potential of about 0 to + 0.5 V (SCE) is required for stable n-type doped polymers, similar to the requirement on the electrode potential for stable undoped p-type polymers. Consequences for bipolar devices are analysed. Huge overpotentials for the redox reaction with wet oxygen are required in order to realize thermodynamically stable bipolar devices from known doped p-type and n-type conducting polymers. Finally, possible solutions, accepting thermodynamic instability, are discussed.
Keywords:

Stability;Doping; Microelectronicdevices

1. Introduction In the last decade, application of conjugated polymers as the active element in microelectronic devices has been investigated. Several types of devices have been studied, e.g. Schottky diodes [ l-71, metal-insulator-semiconductor field-effect transistors (MISFETs) [2,8-121, light-emitting diodes [ 131, light-emitting electrochemical cells [ 141, photodiodes [ 15,161 and grid triodes [ 171. These polymeric microelectronic devices are typically unipolar devices, based on p-type semiconducting polymers. The availability of stable n-type doped semiconducting polymers is desirable to widen the scope of polymeric electronics. It would allow for the fabrication of stable light-emitting p(i)n diodes, bipolar transistors, and the polymeric analogue of the standard silicon MOSFETs, namely, npn and pnp MISFETs. However, these devices have not yet been reported due to the lack of environmentally stable n-type doped conducting polymers. Conjugated polymers, in general, can be doped to both ptype and n-type semiconductors. Some heavily p-type doped polymers are very stable under ambient conditions, e.g. polypyrrole and polyaniline. Conjugated materials that are stable as an undoped or very slightly doped p-type doped material * Correspondingauthor. 0379-6779/97/$17,000 1997Elsevier Science%A. All rights reserved
PIISO379-6779(96)04785-6

are also known, e.g. polythiophene. Until now no n-type doped conjugated polymers having similar stability are known. The difficulty to arrive at such a polymer is related to the well-known instability of organic anions, especially of carbanions. Typically, these ions are easily oxidized in contact with air or water. Starting from the standard redox potentials of, especially, water and oxygen, we will derive the requirements on standard electrode potentials of p- and n-type conducting polymers in order to be electrochemically stable under ambient conditions. Because standard electrode potentials of conducting polymers are not yet available, their values have been approximated by the middle potentials of the anodic and cathodic waves in cyclic voltammograms, El12. We note that the peak-to-peak separation of chemically reversible redox reactions of conducting polymers varies typically from 75 to 110 mV. This indicates that the redox processes are not fully electrochemically reversible. However, for the discussion of the stdbility of n-type doped conducting polymers, this uncertainty in electrode potential can be disregarded. Subsequently, literature data on the stability of n-type doped conducting polymers will be compared with their electrode potentials. Finally, consequences for the feasibility of bipolar polymeric microelectronic devices are discussed.

54

D.M.

de Leeuw et al. /Synthetic

Metals 87 (1997) 53-59

2. Redox reactions involved Possible reactions and their standard electrode potentials, I?, for water and oxygen, which are the main active components in air, are taken from [ 181 and listed in Table 1. Standard redox potentials for the more exotic components, ozone and hydrogen peroxide, are included for completeness as well. Actually, the 12 reactions listed in Ref. [ 181 can be reduced to five. All these reactions are strongly pH dependent. Taking all activity coefficients equal to unity, theredoxpoten-

tials under non-standard conditions, e.g. at different pH, can


be calculated by the Nemst equation [ 191: E=I?+$lnK=l?+0.059

It

log [Red]

[OXI

chemical window for the electrochemical stability of n-type doped conducting polymers assuming values of the overpotential between 0 and 1 V. A second complication arises from the fact that conducting polymers are processed using, typically, nonaqueous solvents, while the redox potentials of Table 1 hold for aqueous solutions. Finally, we note that the reduction o.f oxygen to the superoxide ion (O,-) has been investigated in carefully dried aprotic media [ 2 1] ,The formal reduction potential for this reduction in the solvents DMSO, DMF and acetonitrile using graphite, glassy carbon or platinum as an electrode amounted to about - 0.9 V (SCE) [ 211. As the aim of the present discussion is the stability of n-type doped conducting polymers under ambient conditions, this redox reaction can be disregarded.

e.g. for reaction 1: [o,] [H,O]

CO,1 W+l*

3. Stability Stability of conducting polymers can be divided into electrochemical stability and stability of the polymer due to chemical reactions. Polyacetylene may serve as an example. Undoped polyacetylene is argued below to be electrochemitally stable under ambient conditions. However, poIyacetylene is highly susceptible to triplet oxygen insertion, forming a reactive hydroperoxy free radical [ 221. Here attention will be focused on electrochemical stability because it is the minimum requirement which has to be fulfilled. For the sake of the discussion we note that ap-type doped polymer is capable of accepting electrons and hence is an oxidator, while an ntype doped polymer is capable of donating electrons and hence is a reductor. As an example we discuss the electrochemical stability of an n-type doped conducting polymer towards reduction of water (Table 1, reaction 5). Take an n-type doped polymer with an electrode potential of - 1 V (SCE) . The redox potential for water reduction, at a pH of 7 and an overpotential of 0 V, amounts to -0.658 V (SCE). Combining these halfreactions then yields: 2H20 + 2e 2PolO+ 2e 2Pol- + 2Hz0
1.412

where n is the number of electrons, F is the Faraday constant (96 500 A s mol-) and I? is the standard potential of the half-reaction. The redox potentials for those reactions as a function of pH are given in Table 1. Here, the potential of the standard calomel reference electrode (SCE) is taken as 0.244 V versus NHE. The values presented for the redox potentials are equilibrium values. However, a free energy of activation is often necessary for the reaction to proceed and therefore an excess applied voltage, or overpotential, serves to provide this free energy. Overpotentials required for reactions involving especially gaseous hydrogen or oxygen on metal electrodes may amount to about a volt [ 201, strongly dependent on the type of metal. As will be argued below, the stability of n-type doped conducting polymers strongly depends on the value of the overpotential for the reactions of Table 1 at these polymer electrodes. Unfortunately, theseoverpotentids have not yet been reported. Hence, we will derive an electroTable 1 Possible reactions involving oxygen and water, as well as of ozone and hydrogen peroxide, with their redox potentials as a function of pH No. Reaction 03+2H++2e-@O,+H,O O,fH,Of2eF?02+20H!?=1.826-0.05912XpH Hz02+2Hf+2e-*2H,O H,0Z+2e-F?20HHO,-+H,O+2e-G30H@=1.532-0.05912XpH 0,+4H+ +4e-F?2H20 02+2Hz0+4e+40H,!?=0.985-0.05912XpH 0a+2H++2eW F?HzOz 02+2H,0+2e*HOz+OH02+2H,0+2e@H202+20HEO=0.438-0.05912xpH 2H+ +2e- +H, 2H,O+2e-*HH2+20HEo= -0.244-0.05912xpH EpH-, (V vs. SCE)

= e +

H,+20H2po12Pol+H,+20H-nF(Eox-ERcd)

E= -0.658 V E= - 1.0 V AE= iO.343 V

The free (energy of this redox reaction follows from: A GA = -nFAE=

1.118

= - nF[ - 0.658 - ( - l.O)] = - 63.4 kJ/mol This redox reaction will proceed due to the free energy being negative. This implies that an n-type doped conducting polymer, which can be oxidized below -0.658 V versus SCE, is not stable towards reduction of water. This is graphically represented in the diagram of Fig. 1. A redox reaction will occur in the direction of the reduction half-reaction with higher electrode potential and, similarly, in the direction of the oxidation half-reaction with lower electrode potential. Hence, in the specific example of Fig. 1,

0.571

0.024

- 0.658

D.M.

de Leeuw

et al. /Synthetic

Metals

87 (1997)

53-59

55

Stability

of n-types

towards

H,O reduction

DOPED

N-TYPES UNSTABLE Pal (Polo -->

(or NEUTRAL TOWARDS -za Polo Pal + es + e)

P-TYPES) Hz0

2H20+2em+H,+20H Potential (vs SCE)

Fig. 1. Stability diagram of conducting polymers towards reduction of water, at pH of 7 with negligible overpotentials, TJ= 0.

n-type doped conducting polymers which can be oxidized at a potential lower than the potential of water reduction are unstable in aqueous solution with a pH of 7, assuminga negligible overpotential, v = 0. Because undoped p-type polymers behave similarly to n-type doped polymers, they are included in the stability diagram aswell. A complete stability diagram can be derived when taking all reactions of Table 1 into account. In practice, however, reactions involving hydrogen peroxide and ozone can be disregarded. Under ambient conditions the concentrations are negligibly small and the overpotentials for their in situ formation areprohibitively large. The instability of n-typedoped conducting polymers under ambient conditions is basically due to the redox reaction with wet oxygen. The stability of both doped and undoped n-type as well asp-type conducting
Electrochemical
Reactions pH=7 q=o El involving

polymers towards only oxygen and water is presented in Fig. 2. This diagram as calculated for a pH of 7 and by neglecting overpotentials. The requirementsfor the electrode potentials at different pH and including ozone and hydrogen peroxide can easily be adaptedby using Table 1. Fig. 2 showsthat the stability of n-type doped conducting polymer strongly depends on the electrode potential. The potential at which the n-type dopedpolymer can be oxidized shouldbe higher than - 0.658 V in order to be stabletowards water and higher than + 0.57 1V in order to be stabletowards water and oxygen. Similarly, for p-type doped conducting polymers the oxidation potential should be lower than + 1.412 V in order to be stabletowards oxidation of oxygen to ozone, lower than + I.118 V in order to be stabletowards oxidation of water into hydrogen peroxide, lower than
of conducting
hydrogenperoxide

stability
ozone and

polymers
discarded

2H,O--+0,+4H*+4e

NEUTRAL Pol v, AL 1 DOPED Pol unstable towards -4 Hz0 -0.5 N-TYPES Polo (or NEUTRAL UNSTABLE (Polo unstable --> towards 0 P-TYPES) Pal + e-) 0.5

N-TYPES (or DOPED UNSTABLE Pal , 1 , ( (Pal , 15

P-TYPES) Polo)

+ e -> , ,

+ e ->

+ e

Hz0 and O2

2 HZ0 + 2e ->H,t2OH I Potential

0,+4H++4e--->2Hz0 (KS SCE)

Fig. 2. Stability diagram of conducting polymers under ambient conditions taking only reactions of water and oxygen into account as calculated for pH of 7 with negligible overpotentials.

56

D.M.

de Leeuw

et al. /Synthetic

Metals

87 (1997)

53-59

+ 0.57 1 V in order to be stable towards oxidation of water into oxygen, and lower than + 0.157 V in order to be stable towards oxidation of hydrogen peroxide into oxygen, Under equilibrium conditions n-type doped polymers are stable when the electrode potential is higher than + 0.57 1 V. In practice, however, significant overpotentials may be expected. Taking an overpotential of 0.5 or 1.0 V implies that the requirement for electrochemical stability of n-type doped conducting polymers under ambient conditions is an electrode potential more positive than about 0 or - 0.5 V respectively. The stability of p-type doped conducting polymers is dominated by the oxidation of water. Assuming equilibrium conditions, p-type doped conducting polymers are stable under ambient conditions when the electrode potential is lower than + 0.571 V. Taking into account an overpotential of 0.5 V this then yields an electrode potential lower than about + 1 V. Finally, we reiterate that the requirements on electrode potentials as derived above only describe electrochemical stability with respect to the redox reactions of water and oxygen. Superimposed are requirements on stability due to chemical reactions with components present under ambient conditions. Let us take polyacetylene again as an example. The oxidation potential is about + 0.3 V (SCE) [23]. According to the diagram of Fig. 2, p-type doped polyacetylene should be stable under ambient conditions. However. polyacetylene is highly susceptible to triplet oxygen insertion, forming a reactive hydroperoxy free radical [ 221. Also, an oxygen-free environment is a prerequisite to study polyacetylene [ 22,231. The instability of conducting polymers due to the occurrence of these types of chemical reactions is not accounted for.

Fig. 3. Electrode potentials which a few n-type conducting polymers and at organic compounds can be reduced (solid circles) and electrode potentials at which two well-known p-type conducting polymers can be oxidized (solid squares). Included are only materials of which the stability under ambient conditions has been reported.

4. Comparison

with experimental

data

In order to compare the predictions for stability with experimental data we took literature data on well-characterized conducting polymers. Both the electrode potentials and conductivity in the doped state should have been reported as well as data on the stability under ambient conditions. Materials obtained are summarized in Fig. 3. Filled circles represent electrode potentials at which n-type conducting polymers can be reduced and filled squares represent electrode potentials at which p-type conducting polymers can be oxidized. Poly(pyridine-2,5-diyl) (3) [24,25], poly(quinoline-2,6diyl) (4) [26], poly( 1,5-naphthyridine-2,6-diyl) (5) [27], viologens (6) [ 281, quaternized poly( quinolinium) (8) [ 291 and the diquarternary salt of 1,5-naphtyridine (8) [ 301 can all be reduced. Optical measurements show, as expected, upon reduction a decrease of the intensity of the n-q* transition, and the appearance of a new red-shifted absorption. The conductivity of the n-type doped polymers is reported to be lo- S/cm or more. However, all these materials are extremely unstable under ambient conditions. Lower electrode potentials are required in order to arrive at n-type doped

materials which are stable under ambient conditions. An example is tetracyanoquinodimethane (9, TCNQ) . The potential at which TCNQ can be reduced amounts to + 0.127 V (SCE) [ 191. Thin films of co-evaporated tetrathiofulvalene and TCNQ have been investigated as the active component in MISFETs [ 311. The measured field-effect indicated n-type conductivity. Upon exposure to air the conductivity slightly decreased with time. Lifetime tests showed that the devices were still operational after a few months exposure to air [ 311. This indicates that electrode potentials in the order of about + 0.5 V (SCE) are required for stable n-type doped conducting polymers. This conclusion is supported by stability data of undoped p-type polymers such as polypyrrole (1) and polythiophene (2). Polypyrrole can be oxidized at an electrode potential of about 0 V (SCE) . Undoped polypyrrole is unstable under ambient conditions, and is readily oxidized. On the other hand, the electrode potential at which polythiophene can be oxidized is much higher, about 0.7 V (SCE) . Not surprisingly, undoped polythiophene is stable under ambient conditions. On the right-hand side of Fig. 3 the requirements on the electrode potential in order to arrive at stable n-type doped and stable undoped p-type materials are indicated. The requirements are derived for a pH of 7 and an overpotential of both 0 and 0.5 V. Taking into

D.M.

de Leeuw et al. /Synthetic

Metals

87 (1997)

53-59

57

account the approximations involved, a fair agreement is obtained. Also, stability under ambient conditions of n-type doped polymers requires an electrode potential higher than about 0 to f 0.5 V. This requirement on the electrode potential is not surprising. The stability of n-type doped polymers is dominated by the redox reaction with wet oxygen, similar to undoped p-type polymers. It is well known that undoped p-type polymers are stable under ambient conditions when the potential for oxidation is higher than about 0 to + 0.5 V. Also, the same stability requirement should hold for the potential at which n-type doped conducting porymers can be reduced.

5. Consequences

for polymeric

microelectronic

devices

A number of requirements have to be fulfilled in order to fabricate practical polymeric p(i) n diodes, bipolar transistors, and pnp and npn MISFETs. Firstly, the polymers used should preferably be stable under ambient conditions. Secondly, a p(i)n junction should arise from the combination of a p-type doped and an n-type doped conducting polymer. This implies a difference in Fermi energy, or work function, of at least about 0.5 V or more. The work function of undoped conducting polymers is unambiguously correlated with the electrode potential [ 32,331. For unintentionally p-type doped conducting polymers the work function, or solid state ionization potential, is equal to the electrochemical potential at which the polymer can be oxidized (versus SCE) plus 4.4 V. For unintentionally n-type doped conducting polymers the work function, or electron affinity, is equal to the potential at which the polymer can be reduced (versus SCE) plus 4.4 V. Thus, a prerequisite for formation of a p( i)n junction is a difference between the electrode potentials of the p-type doped and the n-type doped conducting polymer of at least about 0.5 V. Finally, the conductivity of unintentionally doped polymers is extremely low, typically less than 10s6 S/cm. Similar to polymeric Schottky diodes, current transport in hypothetical p(i) n junction diodes made from unintentionally doped p-type and n-type conducting polymers will be bulk resistance limited [ 41. Hence, these p( i>n diodes and devices based thereof will operate at low frequency only. In order to decrease the bulk series resistance, conducting (u 2 10e3 S/cm) and, hence, highly doped [4] p-type and n-type polymers should be applied. Interface chemistry is then required to yield a doping profire in both polymers and hence the formation of a p(i)n junction. These requirements are unlikely to be met simultaneously. As argued above (see Fig. l), n-type doped conducting polymers are expected to be stable under ambient conditions only if the electrode potential is more positive than about 0 to + 0.5 V (SCE) . On the other hand p-type doped conducting polymers are stable under ambient conditions when their electrode potential is more negative than +0.5 to - 0 V (SCE) . Equally, interface chemistry to realize a dopant profiIe in both polymers and hence a p(i) n junction requires that

the electrode potential of the n-type doped conducting polymer should be lower than that of the p-type doped polymer. Thus thermodynamically stable polymeric bipolar devices cannot be expected. We note that the same is true for inorganic microelectronic devices. However, there the redox reactions involved often are kinetically hampered, and the devices are either passivated or encapsulated to minimize exposure to wet air. The analysis on stability of doped polymers presented above may also be applied for poIymeric light-emitting diodes and light-emitting electrochemical cells. PoIymeric light-emitting diodes typically consist of a stack of thin undoped polymeric films sandwiched between a hole injecting and an electron injecting electrode. At thermal equilibrium the polymeric films are not charged and, hence, environmentally stable. Upon applying a forward bias, however, the polymeric film at the electron injecting contact gets negatively charged. In a first approximation, ignoring the role of charge compensating counter ions, the stability of negatively charged polymers will be comparable to that of n-type doped conducting polymers. For the presentIy used poly(phenylene-vinylene) -type polymers, taking into account the electrode potentials at which they are reduced, this then leads to environmentally unstable devices. If relatively stable electron-injecting metallic contacts such as aluminium are used giving long shelf lives, encapsulation is still worthwhile to improve the operational lifetime. Similarly, there is no fundamental limitation to the shelf lifetime of light-emittingelectrochemical cells. However, upon biasing the device a p(i) n junction is formed. The n-type doped polymers presently used are thermodynamically unstable. Thus, the present polymeric light-emitting electrochemical cells operate in oxygen-free environment only. In order to fabricate bipolar devices, given that thermodynamically stable devices are not to be expected, we aim at n-type doped conducting polymers with an electrodepotential of about -0.5 V. A typical electrode potential for p-type doped polymers is around 0 to +0.5 V. Hence, upon combination a p( i)n junction may be expected. Obviously, stability under ambient conditions then requires either encapsulation of the final devices or a high overpotential for the redox reaction with wet air. Methods for stabilizing thermodynamically unstable polymers, e.g. polyacetylene, have been reported. An increase of stability has been reported by incorporation of free-radical quenchers, or oxygen scavengers [ 341. Encasement in materials with low permeability for wet oxygen might be a solution as well, Reliable values for overpotentials have not yet been reported. High overpotentials, up to 2 V, have been claimed for electrochemically n-type doped polythiophenes functionalized with alkoxy side groups or pendant crown ethers [ 35,361. However, the spectral and conduction characteristics of n-type doping are not displayed. As demonstrated by Zotti et al. [ 371, the reduction peaks in the cyclic voltammograms are not due to n-type doping but due to alkali-ionassisted reduction of the polymer conjugated backbone. The

58

D.M. de L.eeuw et al. /Synthetic

Metals 87 (1997) 53-59

irreversible reduction can be attributed by the pinning of the negative charge at the double bond by the smallmetal cation with formation of a tight pair. The reaction is fasterthe smaller the cations, due to the greaterelectrophilicity, corresponding to strongerion-pair formation. The phenomenonthat the stability of n-type doped conducting polymers dependson the type of counterions used for charge compensationis quite general [ 371. Polythiophene may serve asan example.Electrochemical n-type doping of polythiophene by usingtetraalkylammonium ions leads to an n-type doped state which could be thoroughly investigated by cyclic voltammetry [ 381, UV-Vis spectroelectrochemistry [39-42] and in situ ESR [42,43]. However, electrochemical doping with alkaline cations is not possible [ 38,441. Alkaline ions are detrimental to electrochemical n-type doping, even in very low concentrations [45], and only with hexamethylphosphoramide as a solvent have lithium and potassiumsaltsproved to someextent to result in n-type doping [ 381. This anomalous effect is not peculiar to polythiophene sincepolyacetylenewhich can be n-type dopedchemically with alkalimetals [46] may be doped electrochemically only with tetraalkylammoniumsalts, whereasalkaline cations lead to its degradation [47]. Finally, we note that even design and synthesisof an ntype conducting polymer with an electrodepotential of about - 0.5 V (SCE) is not trivial. The sum of ionization energy and electron affinity for aromatic hydrocarbon crystals is constant and equal to 8.3 eV [ 33,481. Thus, for a throughconjugatedpolymeric aromatic hydrocarbon with a bandgap of 3 eV an electron affinity of 2.6 eV is expected. This corresponds with an electrode potential 2.6 -4.4 = - 1.8 V (SCE) [ 32,331, It is thusnot surprisingthat electrodepotentials of well-known n-type doped conducting polymers are typically about -2 V (SCE). A stable n-type doped conducting polymer requires an electrode potential which is higher by about 2 V (Figs. 2 and 3). As explained above, for application in bipolar polymeric microelectronic devices we aim for n-type dopedpolymers with an electrodepotential of about - 0.5 V (SCE). Realization of this shift of about I1.5 V is an enormouschallengefor synthetic chemists. Electronegative heteroaromaticpolymers, namely,polythiadiazoles, are presently being synthesizedfor this purpose.

[2] J.H. Burroughes, CA. Jones and R.H. Friend, Nature,


137. [31 [41

335

(1988)

H. Koezukaand S. Etoh, J. Appl. Phys., 54 (1983) 2511. D.M. de Leeuw and E.J. Lous, Synth. Met., 65 (1994) 45. [5] E.J.Lous, P.W.M. Blom,L.W. Mo1enkampandD.M. deLeeuw,Phys. Rev.5.51 (1995) 17 251. [6] H. Tomozowa, D. Braun, S.D. Philips, R. Worland, A.J. Heeger and H. Kroemer, Synth. Met., 28 (1989) C687. [7] A. Assadi, C. Svensson, M. Willander and 0. InganPs, J. Appl. Phys., 72 (1992) 2900. 181 H. Fuchigami, A. Tsumuraand H. Koezka, Appl. Phys. Len., 63 ( 1993) 1372. [9] F. Gamier, R. Hajaoui, A. Yassar and P. Srivastava, Science, 265 (1994) 1684. [ 101 A. Dodabalapur, L. Torsi and H. Katz, Science, 268 (1995) 270. [ 11I A.R. Brown, D.M. de Leeuw, E.E. Havinga and A. Pomp, Synrfz.Mel., 68 (1994) 65. [ 121 A.R. Brown, A. Pomp, CM. Hart and D.M. de Leeuw, Science, 270 (1995) 972. [ 131 J.H. Burroughes, D.D.C. Bradley, A.R. Brown, R.N. Marks, K. Mckay, R.H. Friend, P.L. Bum and A.B. Holmes, Nature, 347 (1990) 539. [ 141 Q. Pei,G. Yu,C.Zhang.Y. YangandA.J. Heeger,Science,269 (1995) 1086. [ 151 N.S. Saricitrci, L. Smilowitz, A.J. Heeger and F. Wudl, Science, 258 (1992) 1474. [ 161 J.J.M. Halls, C.A. Walsh, NC. Greenham, E.A. Marseglia, R.H. Friend, S.C. Moratti and A.B. Holmes, Nature, 376 (1995) 498. [ 171 Y. Yang and A.J. Heeger, Nature, 372 (1994) 344. [ 181 Handbook of Chemistry and Physics, CRC Press,Boca Raton, FL, 60th edn., 1979-1980. [ 191 A.J. Bard and L.R. Faulkner, Electrochemical Methods, Wiley, New York, 1980. [20] W.J. Hamer and R.E. Wood, in E.U. Condon and H. Odishaw (eds.), Handbouic of Physics, McGraw-Hill, New York, 1958. [21] D. Vasudevan and H. Wend& J. Electroanal. Chem., 192 (1995)
69.

3. Pochan,inT.A. Skotheim (ed.), Handbookof CondlcctingPolyrners, Vol. 2, Marcel Dekkcr, New York, 1986, pp. 1383-1405. [23] A.G. McDiarmid and R.B. Kaner, in T.A. Skotheim (ed.), Handbook ofconducting Polymers, Vol. 1, Marcel Dekker, New York, 1986, pp.
[22] 689-727.

Acknowledgements We gratefully acknowledgeour colleagues,K. Chmil, E.E. Havinga andA.J.W. Tol, at the Philips Research Laboratories for critical discussions.

T. Yamamoto, T. Maruyama, Z. Zhou, T. Ito, T. Fukuda, Y. Yoneda, F. Begum, T. Ikeda, S. Sasaki, H. Takezoe, A. Fukudaand K. Kubota, J. Am. Gem. Sot., 116 (1994) 4832. [25] T. Maryama, K. Kubota and T. Yamamoto, Macromolecules, 26 (1993) 4055. [26] S.E. Tunney, J. Suenaga and J.K. Stille, Macromolecules, 16 (1983) 1398. [27] N. Saito and T. Yamamoto, Macromolecules, 28 (1995) 4260. [28] CL. Bird and A.T. Kuhn, Chem. Sot. Rev., IO (1981) 49. [29] T. Kanbara and T. Yamamoto, Macromolecules, 26 (1993) 1975. [3OJ L.A. Summers and J.E. Dickeson, J. Chem. SOL, Chem. Commun., (1967) 1183. [31] A.R.Brown,D.M.deLeeuw,E.J.LousandE.E.Havinga,Synrlz.Met.,
[24] 66 (1994) [32] 257-261.

References
[l] J. Kanicki, in T.A. Skotheim (ed.), Handbook of Conducting Polymers, Vol. I, Marcel Dekker, New York, 1986, pp. 544-660.

J.L. BrBdas, R. Silbey, D.S. Boudreaux and R.R. Chance, J.Am. Chem. SOL, 105 (1983) 6555. [33] F. Lohmann, 2. Naturforsch., Teil A, 22 (1967) 843. [ 341 J.M. Pochan, H.W. Gibson and J. Harbour, Polymer, 23 (1982) 439. [35] Y. Miyazaki and T. Yamamoto, Chem. Left., (1994) 41. [36] Y. Miyazaki and T. Yarnamoto, Synth. Met., 69-71 (1995) 317. [37] G. Zotti, G. Schiavon and S. Zecchin, Synth. Met., 72 (1995) 275. [38] M. Mastragostino and L. Soddu, Electrochim. Acra, 35 (1990)
463. [391 31

M. Onoda, Y. Manda, S. Morita and K. Yoshino, Jpn. J. Appl. Phys., (1992) 2265.

D.M. [40]

de Leeuw et al. /Synthetic

Metals [44] [45] [46] [47] [48]

87 (1997)

53-59

59

K. Kaneto, S. Ura, K. Yoshino and Y. Inuishi, Jpn. J. Appl. Phys., 23 (1984) L109. [41] M. Sato, S. Tanaka and K. Kaeriyama, J. Chem. SOL, Chem. Commun., (1987) 1725. [42] M. Onoda, S. Morita, T. Iwasa, H. Nakayama and K. Yoshino, Jpn. J. Appl. Phys., 31 ( 1992) 1107. [43] M. Onoada, Y. Manda, S. Morita and K. Yoshino, J. Phys. Sot., 59 (1990) 213.

M. Kobayashi, N. Colaneri, M. Boysel, F. Wudl and A.J. Heeger, J. Chem. Phys., 82 (1985) 5717. R. Bojas and D.A. Buttry, Chem. Muter., 3 (1991) 872. A.C. MacDiarmid and A.J. Heeger, Synth. Met., I (1979) 101. T. J0sefiak.E.J. Ginsburg, C.B. Gorman, R.H. Grubbs andN.S.Lewis, J. Am. Chem. Sot., 115 (1993) 4705. N.S. Hush and J.A. Pople, Trans. Faraday Sot., 51 (1955) 600.

S-ar putea să vă placă și