Sie sind auf Seite 1von 6

Available online at www.sciencedirect.

com

Inorganica Chimica Acta 361 (2008) 792797 www.elsevier.com/locate/ica

Note

Unusual photoelectrochemical behaviour of nanocrystalline TiO2 lms


R. Solarska, I. Rutkowska 1, J. Augustynski
*
` ve, Switzerland partement de Chimie Mine rale, Analytique et Applique e, Universite de Gene De Received 4 February 2007; received in revised form 10 May 2007; accepted 23 May 2007 Available online 2 June 2007 Dedicated to Professor Michael Gra tzel.

Abstract Although photooxidation of water and numerous other species which are part of reversible redox couples is poorly ecient at nanocrystalline TiO2, conversely high photocurrent eciencies are observed for the oxidation of various organic molecules. This is associated with the fact that in most cases photooxidation of organic molecules does not produce species able to act as electron scavengers. The behaviour of nanocrystalline TiO2 photoelectrodes is clearly dominated by the indirect recombination or redox cycling where intermediates or products of the hole transfer act in turn as electron scavengers. These processes occur whatever the applied anodic bias showing that the actual potential in most of the nanocrystalline TiO2 lm is disconnected from that imposed to the conducting substrate. 2007 Elsevier B.V. All rights reserved.
Keywords: Nanocrystalline titanium dioxide; Semiconductor lm photoanode; Photooxidation reactions; Photocurrent eciency

1. Introduction Successful application of mesoporous anatase TiO2 lms in dye-sensitised solar cells (DSSCs) by Michael Gra tzel and co-workers [13] caused a wide-spread interest in the properties of this new category of porous nanocrystalline (NC) semiconducting electrodes. Interestingly, already rst investigations aimed at characterising such photoelectrodes under band-gap illumination [47] revealed a number of specicities and also anomalies in their behaviour. In particular, Hodes et al. [4] who studied lms formed from a few nm in diameter crystals of two other semiconductors, CdS and CdSe, in contact with solutions of reversible redox couples, observed a regular decrease of the photoresponse (particularly pronounced in the region of short waveCorresponding author. Tel.: +41 22 7402053. E-mail address: Jan.Augustynski@chiam.unige.ch (J. Augustynski). 1 Present address: Department of Chemistry, Warsaw University, Poland. 0020-1693/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.ica.2007.05.037
*

lengths) on increasing the lm thickness. Excluding the built-in of a conventional space-charge layer, like that being formed in bulk semiconductors, the authors postulated the charge separation to occur at the interface between the individual nanoparticles and the electrolyte lling the pores in the lm. The small particle size and the porous lm structure greatly facilitate the withdrawal of the photogenerated holes (for the case of an n-type semiconductor) by the electron donors in the electrolyte. Such a mechanism provided an explanation for fairly large quantum eciencies observed with thin lms and also their drop when the lm thickness became comparable or exceeded the light absorption depth. The current losses occurring in the latter case were attributed to the charge recombination at crystal boundaries and, in addition, to the so-called indirect recombination by the electron injection into electrolyte. Another consequence of increasing the thickness of the NC lms evoked by the authors [4] is the loss in photovoltage and deterioration of ll factor due to a high (dark) resistance of colloidal CdS and CdSe lms.

R. Solarska et al. / Inorganica Chimica Acta 361 (2008) 792797

793

2. Experimental TiO2 lms were prepared from a commercial P25 TiO2 powder composed of ca. 80% anatase and 20% rutile, about 30 nm in diameter particles (Degussa). A few lm-thick P25 lms were deposited on conducting glass substrates (Nihon Sheet Glass Co., 12 X/square comprising a 0.5 lm thick Fdoped SnO2 overlayer). A suspension of the P25 powder in DMF containing dissolved poly(vinylideneuoride), PVDF, was sprayed on the substrate, dried in air for 40 min at 100 C to evaporate the solvent, and then annealed for 1 h at 450 C. The lms were formed by multiple applications and annealing; the thickness of individual layers was of ca. 1 lm. Fig. 1 shows the scanning electron microscopic view of an NC TiO2 lm deposited on the conducting glass substrate. The photoaction spectra, showing the variation of incident photon-to-current conversion eciency (IPCE) vs. excitation wavelength, were determined using a 500 W xenon lamp (Ushio UXL-502HSO) and a Multispec 257 monochromator (Oriel) with a bandwidth of 4 nm. The absolute intensity of the incident light from the monochromator was measured with a Model 730A radiometer/photometer from Optronic Lab. Measurements involving variable light intensities were carried out using a Spectra Physics model 2025-04 argon-ion laser (334.0, 351.1 and 363.8 nm emission lines) equipped with a series of neutral density lters. The absolute intensity of the incident light was monitored with a laser power/energy meter (model DG, Newport Instruments). All measurements were carried out in a two-compartment Teon cell equipped with a quartz window, by irradiating the TiO2 electrode from the side of the lm/solution interface. The cell contained about 30 cm3 of carefully deaerated solution. The counter electrode consisted of a platinum sheet and Hg2Cl2/Hg/Cl was the reference electrode. All solutions were made from reagent grade chemicals and Milli-Q water and were purged with nitrogen. The measurements were

0.25

b
Photocurrent / mA cm-2
0.2

a
0.15

0.1

0.05

0 -0.3

-0.1

0.1

0.3

0.5

0.7

0.9

1.1

Potential / V vs. RHE


Fig. 2. Photocurrent vs. potential plots recorded for: (a) a 2.5 lm-thick and (b) a 50 lm-thick nanocrystalline TiO2 lm, immersed in a 0.1 M HClO4/0.1 M HCOOH solution and irradiated with a 300 nm monochromatic light (ca. 700 lW cm2). The potential is given with respect to the reversible hydrogen electrode (RHE) in the same solution.

carried out under potential-controlled conditions using an EG&G Princeton Applied Research Model 362 potentiostat. 3. Results and discussion Photocurrentvoltage measurements involving nanocrystalline TiO2 anatase2 photoelectrodes irradiated with the band-gap UV light [68], revealed a behaviour apparently in striking contrast to that of the colloidal CdS and CdSe lms described in Section 1. In fact, unusually high incident-photon-to-current eciencies (IPCEs), close to 150%, were observed for the oxidation of some small organic molecules at NC TiO2 lms of large thicknesses, an order of magnitude larger than those of the CdS and CdSe lms [6,8]. Although the fraction of IPCE exceeding 100% is obviously due to the occurrence of photocurrent doubling during photooxidation of formic acid and/or methanol, still the IPCEs reached for the NC TiO2 lms are much larger than those for the bulk TiO2 (both anatase and rutile) photoelectrodes [9]. Particularly intriguing is the marked increase in IPCEs on increasing the NC lm thickness well above the light absorption depth. This is illustrated in Fig. 2 showing photocurrent vs. potential plots recorded for two NC TiO2 lms having thicknesses of ca. 2.5 and 50 lm, illuminated with a 300 nm monochromatic light. Since the absorption depth, a1, of 300 nm light is close to 30 nm (i.e. roughly the size of one particle) the majority
2

Fig. 1. Scanning electron micrograph of a mesoporous TiO2 lm formed from P25 nanoparticles.

These and other NC TiO2 photoelectrodes described in the present account were formed from P25 (Degussa) nanoparticles composed of ca. 80% anatase and 20% rutile.

794

R. Solarska et al. / Inorganica Chimica Acta 361 (2008) 792797

charge carriers (electrons) photogenerated in the very thin outermost part of the lm have to pass through hundreds of grain boundaries before reaching the substrate. Despite the high dark resistance of NC TiO2 lms in the direction of the current ow (in the range of 1071010 X cm) [10], comparable with that of the colloidal CdS and CdSe lms (109 X cm) [4], the slopes of photocurrentvoltage plots in Fig. 2 for the thinner and the thicker lm are practically identical. Note that the 50 lm lm comprises roughly 1500 layers of individual TiO2 particles and quite as many grain boundaries. While in the case of colloidal CdS and CdSe lms the maximum IPCE decreased and shifted towards longer wavelengths on increasing the lm thickness, the shape of the photoresponse of the NC TiO2 counterparts is essentially unaected by the lm thickness with the maximum remaining at ca. 300 nm, i.e. at the high-energy side of the spectrum (cf. Fig. 3). These characteristics of the photocurrent spectra suggest that the principal loss mechanism in the NC TiO2 lms is due to charge recombination (both direct and indirect) occurring principally in the illuminated part of the lm. With regard to the small resistance losses in these initially highly resistive NC TiO2 networks, the most plausible explanation is the lm charging occurring from the beginning of the photocurrent ow, before the steady state is reached [8]. Although the charging of the NC TiO2 network is due to the photogenerated electrons, this process is to be distinguished from the classic photoconductivity since it embraces the whole lm, even if its illuminated portion does not exceed 0.1% of the total lm thickness (corresponding to the case of the 50 lm thick lm illuminated with 300 nm light). Such charging of large unilluminated, initially insulating parts of the lm appears as a unique feature of porous nanocrystalline semiconductor/electrolyte junctions oering a specic mechanism of charge compensation. Given their large porosity (ca. 50%) and due to the

small size of semiconductor particles, such lms exhibit a particularly large internal surface-to-volume ratio. Almost every particle is, at least partially, in contact with an electrolyte lling the pores in the lm. Under such conditions, a moderate adjustment in the cation concentration in the Helmholtz layer is enough to compensate for a large increase in concentration of electrons in the NC semiconductor [8]. The charging of the porous NC TiO2 lm immersed in an electrolyte actually results in its temporary doping similar to that permanently produced, for example, by an admixture to the semiconductor of Nb(V) [11]. Although charging of the TiO2 nanoparticulate network which actually leads to a large rise in concentration of free electrons may account for the small resistance losses in such lms, it does not provide any direct explanation for the observed shape of the photocurrent spectra in Fig. 3. In this regard, an insight comes from the comparison of spectral responses of the NC TiO2 lms recorded in solutions containing diering amounts of electron donors. As shown in Fig. 4, rising concentration of methanol causes a signicant increase in photocurrents extending over the whole range of potentials. This strongly suggests that the reactive species on the surface of the TiO2 particles is in adsorption equilibrium with methanol in solution. The adsorption control of the photocurrent was conrmed by experiments conducted in a series of methanol/ water mixtures using NC TiO2 lms deposited on interdigitated arrays of microelectrodes [12]. The fact that for those relatively thin (ca. 200-nm-thick) TiO2 lms, the photocurrent regularly increased on increasing the CH3OH/H2O

25

c
20

Photocurrent / mA cm-2

b
15

160 140 120

b c a

10

IPCE / %

100 80 60 40 20 0 280 300

0 -0 .6
320 340 360 380 400 420 440 460 480

-0.4

-0.2

0.2

Potential / V vs. SCE Wavelength / nm


Fig. 4. Photocurrent vs. potential curves for a 3 lm-thick NC TiO2 lm irradiated with ca. 700 mW/cm2 argon-ion laser light (334.4, 351.1 and 363.8 nm emission lines) in 0.1 M HClO4 solutions containing increasing amounts of methanol: (a) 0.1 M MeOH, (b) 1 M MeOH, (c) 3 M MeOH. The potential is given with respect to the saturated calomel electrode.

Fig. 3. Photocurrent action spectra expressed as IPCEs for: (a) 2.5 lmthick, (b) 10 lm-thick and (c) 50 lm-thick NC TiO2 lm polarised anodically at 0.7 V vs. RHE in a 0.1 M HCOOH/0.1 M HClO4 solution.

R. Solarska et al. / Inorganica Chimica Acta 361 (2008) 792797


16 14 12 10

795

IPCE / %

a b

and, consequently, low photocurrent eciency for the photooxidation of water whatever the imposed anodic bias. The observed poor eciency for oxygen generation is most probably directly connected with the formation of surface TisO radical species, considered as primary intermediates in the photo-oxidation of water at anatase [13,9] TiO2 hm ! h e Tis OH h ! Tis O

8 6 4 2 0 280

1 H aq 2

These species, which in a further step form peroxo dimers leading subsequently to oxygen evolution, appear as plausible recombination centres for the photogenerated electrons
330 380 430 480

Tis O e H aq ! Tis OH

Wavelength / nm
Fig. 5. Photocurrent action spectra for: (a) 2.5 lm-thick, (b) 10 lm-thick and (c) 50 lm-thick NC TiO2 lm polarised at 0.7 V vs. RHE in a 0.1 M HClO4 solution.

ratio, up to a nal ratio of 99/1, is a strong indication that the rate of the photooxidation process is, in fact, governed by the surface coverage of the electron donor. The latter result also conrms that methanol undergoes preferential photooxidation with respect to water. This is consistent with poor IPCEs for the photooxidation of water from an HClO4 solution shown in Fig. 5 and similar results obtained using NaOH as supporting electrolyte. Note that the sequence of photocurrent spectra in Fig. 5 diers from those for the photooxidation of formates and methanol, with the IPCEs decreasing when the lm becomes thicker (except for the vicinity of the long-wavelength cut-o). In this case, in a way similar to that observed for the colloidal CdS and CdSe lms, the extent of recombination increases when the e/h+ pairs are generated farther from the substrate. As will be discussed later, both intermediates of the photooxidation of water and its nal product oxygen may be responsible for the enhanced recombination. Note that such a behaviour is in stark contrast to that of bulk TiO2 (both anatase and rutile) photoelectrodes exhibiting a substantial activity for oxygen evolution with IPCEs reaching 60% [9]. In the case of the bulk TiO2 electrodes, the indirect recombination is, in fact, prevented by the withdrawal of the electrons from the surface region of the semiconductor by the applied anodic potential.3 Accordingly, in such a case, application of an anodic bias on the order of a few tenths of Volts renders the back reaction insignicant. In contrast, the nanocrystalline TiO2 electrodes still exhibit an important extent of recombination

The latter indirect recombination mechanism which one can also call redox cycling occurs in addition to the direct recombination of photogenerated e/h+ pairs. The indirect charge recombination following reactions 2 and 3 also explains a relatively sharp concentration dependence of IPCEs for oxidation of small organic molecules. The extent to which an aliphatic alcohol is able to compete for photogenerated holes with surface hydroxyl groups largely depends on its surface coverage. In fact, the TisO species formed may either: (i) react with alcohol molecules diusing from the solution, (ii) recombine with photogenerated electrons or (iii) dimerize to form surface peroxo species. As shown by the low IPCEs recorded in the solution of the supporting electrolyte alone (Fig. 5), the recombination dominates over the reaction pathway leading to oxygen evolution. Conversely, particularly high photocurrent eciencies already attained in moderately concentrated solutions of small organic molecules demonstrate their ability to compete eciently with direct charge recombination and to practically rule out oxygen evolution. It is interesting to note, however, that even for these ecient photooxidation reactions, the photocurrents are markedly aected by the presence of electron acceptors in
80 70 60

IPCE / %

50 40 30 20 10 0

280

330

380

430

480

For a bulk n-type semiconductor the surface electron density, Ns, decreases with the applied anodic potential, E , as: N s = N bexp [e(E Vfb)/kT] where Nb is the bulk electron density in the semiconductor, Vfb is its at-band potential, e is the elementary charge, k is Boltzmanns constant and T is the absolute temperature.

Wavelength / nm
Fig. 6. Photocurrent action spectra for a 2.5 lm thick NC TiO2 lm recorded in a 3 M MeOH/0.1 M HClO4 solution: (a) saturated with nitrogen and (b) in the presence of oxygen.

796

R. Solarska et al. / Inorganica Chimica Acta 361 (2008) 792797


100 90 80 70

the solution. Fig. 6 shows photocurrent spectra for a 2.5 lm thick NC TiO2 lm recorded in 3 M CH3OH solutions deaerated with nitrogen and/or saturated with oxygen. Despite the largely positive value of the imposed potential (exceeding the equilibrium potential of oxygen formation) the presence of oxygen in the solution causes a marked drop of IPCEs, especially at the high-energy side of the spectrum where the e/h+ pairs are generated close to the outer lm/solution interface. An important feature related to the photooxidation of most of organic species is that neither the nal products (most often CO2 and H aq ions) nor the intermediates of the reaction act as eective electron acceptors. For example, the radical species formed in the course of photooxidation of small organic molecules, due to their largely negative oxidation potentials, very rapidly undergo further oxidation by electron injection into the conduction band of the semiconductor (the process leading to the photocurrent doubling). The photooxidation of water is by no means the only photoreaction aected by the indirect recombination. Every time when the product of the photooxidation reaction may act as an eective acceptor for the conduction-band electrons the redox cycling leads to marked photocurrent losses. This is actually the case for a number of reversible redox couples such as quinone/hydroquinone, I3 =I , or Br2/Br. There are also certain redox species able to act both as electron donors and electron acceptors. Such behaviour is exemplied by methyl viologen dication, MV2+ [14]. As shown in Fig. 7B, the addition of even a small amount, 1 mmol L1, of MV2+ species to a solution containing an ecient hole scavenger-sodium formate causes a large drop of the photocurrent eciency at the NC TiO2 electrode. However, the addition of MV2+ to an acidic solution containing the same amount of formic acid produces an opposite eect, i.e., a moderate increase of the IPCEs (cf. Fig. 7A). This contrasting behaviour is explained by a changeable position of the reduction potential of the MV2+/MV+ couple (E0 = 0.446 V) with respect to the conduction band of TiO2. In fact, the potential of the conduction band edge of TiO2, Ecb = 0.1 0.059 pH (V vs. NHE), which is more positive than E(MV2+/MV+) in acidic solutions, becomes more negative than the latter in alkaline solutions where a signicant portion of photogenerated electrons are intercepted by MV2+ cations before reaching the substrate: TiO2 hm ! h e HCOO 2h ! CO2 H aq MV
2

a b

IPCE / %

60 50 40 30 20 10 0 280 330 380 430 480

Wavelength / nm
50 45 40 35

IPCE / %

30 25 20 15 10 5 0 280 330 380 430 480

Wavelength / nm
Fig. 7. Photocurrent action spectra for a 2.5 lm-thick NC TiO2 lm obtained (A) in a 0.1 M HCOOH/0.1 M HClO4 solution (pH 1.5): (a) without MV2+ and (b) after the addition of 1 mM MV2+ (B) in a 0.1 M NaClO4/0.1 M HCOONa/borax solution (pH 9.5): (a) without MV2+ and (b) after the addition of 1 mM MV2+. The cut-o at ca. 300 nm is due to the absorption of the MV2+ solution.

10 4 5

e ! MV

Another important feature is the persistence of a strongly diminished photocurrent, due to the electron scavenging process (reaction 5), at positive applied potentials, much more positive than the onset potential for the reduction of MV2+ cations from a 1 mM solution (0.35 V) [15]. This is a clear indication that the actual potential prevailing across a signicant part of the NC TiO2 lm must be sub-

stantially more negative than the potential applied to the conducting SnO2F/glass substrate. The fact that at the anodically polarised NC TiO2 electrode, the MV2+ cations start to scavenge photogenerated electrons in the solution of a pH higher than 7 is suggesting that, in a major part of the lm, the quasi-Fermi level of electrons remains close to the conduction band edge potential of TiO2 [14], far from the potential actually applied to the substrate. Although in alkaline solutions the behaviour of MV2+ cations is dominated by their role of electron scavengers, in acidic solutions, they appear as quite eective hole scavengers undergoing oxidative photodegradation at the NC TiO2 electrode. This is consistent with the increase in the IPCEs observed after the addition of MV2+ to the solution of formic acid (cf. Fig. 7A). Since in the case of porous NC lms the semiconductor/ electrolyte junction extends in general over the whole lm thickness, the charge transfer reactions occur over the entire three-dimensional region of the lm where the charge

R. Solarska et al. / Inorganica Chimica Acta 361 (2008) 792797


90 80 70 60 50 40 30 20 10 0 -0.9 -0.7 -0.5 -0.3 -0.1 0.1 0.3 0.5

797

Potential / V vs. SCE


Fig. 8. Photocurrent vs. potential curves for a 10 lm-thick NC TiO2 lm irradiated with: (a) 700 mW/cm2; (b) 70 mW/cm2 argon-ion laser light (334.4, 351.1 and 363.8 nm emission lines) in a 1.5 M HCOOH/1.5 M HCOONa solution. The potential is given with respect to the saturated calomel electrode.

charge recombination or redox cycling induced by surface TisO radical species. A similar role is played by all intermediates or products of the hole transfer reactions able to act as electron scavengers. The occurrence of the intensive redox cycling is to be related to the absence of the conventional depletion layer within the NC TiO2 lms. The persistence of the redox cycling despite a large anodic bias indicates moreover that the actual potential in most of the NC TiO2 lm is disconnected from that imposed to the conducting substrate. As discussed earlier [14], the electric potential drop in the NC TiO2 lm might occur in the rst two-three layers of nanoparticles adjacent to the FSnO2/glass substrate. Probing the photoelectrochemical behaviour of relatively thick NC TiO2 photoanodes using incident UV light of different wavelengths and intensities revealed their another feature which is the signicant electronic conductivity of large unilluminated portions of the lms in contact with the conducting glass substrate. This is highlighted by the observation of an ecient collection of electrons injected into the very outermost part of a 50 lm thick TiO2 lm, exposed to short-wavelength strongly absorbed UV light, incident from the side of the lm/solution boundary. Acknowledgement

carriers are generated (corresponding to the absorption depth of the incident light, i.e. several microns for the wavelengths close to the absorption edge of TiO2). Consequently, the amount of the photocurrent is no more controlled by the migration of the minority charge carriers across the depletion layer (as in the case of bulk semiconductors) but by the adsorption/diusion of the reacting species from the electrolyte. One of the consequences of such a situation is that particularly large photocurrents are reached at the NC TiO2 electrodes for photooxidation of large concentrations of the most eective hole scavengers under high intensity illumination as illustrated in Fig. 8. However, this also leads to photocurrent limitations associated with the transport of the reacting species within the pores of the lm, arising at high illumination levels [16]. As a counterpart, the NC electrodes expose a large surface area to the direct and indirect charge recombination. The occurrence or not of the latter process, i.e. of the redox cycling appears as the principal reason of largely diering eciencies of photooxidation reactions at NC TiO2. 4. Conclusions In contrast to the bulk-single crystal or polycrystallineTiO2 photoelectrodes, their nanocrystalline counterparts show a particularly wide range of activities dependent upon the nature of both hole and electron scavengers present in the solution. While photooxidation of water proceeds with relatively high current eciencies (IPCEs) at the bulk anatase and rutile, it is aected by a large extent of recombination at the nanocrystalline TiO2. We explain these low photocurrent eciencies by the occurrence of an indirect

Photocurrent / mA cm-2

This work was supported by a grant from the Swiss National Science Foundation. References
[1] B. ORegan, J. Moser, M. Anderson, M. Gra tzel, J. Phys. Chem. 94 (1990) 8720. [2] B. ORegan, M. Gra tzel, Nature 353 (1991) 737. [3] M.K. Nazeerudin, A. Kay, I. Rodicio, R. Humphrey-Baker, E. Mu tzel, J. Am. Chem. Soc. ller, P. Liska, N. Vlachopoulos, M. Gra 115 (1993) 6382. [4] G. Hodes, I.D.J. Howell, L.M. Peter, J. Electrochem. Soc. 139 (1992) 136. [5] A. Hagfeldt, S.E. Lindquist, M. Gra tzel, Solar Energy Mater. Solar Cells 32 (1994) 254. [6] A. Wahl, M. Ulmann, A. Carroy, J. Augustynski, J. Chem. Soc., Chem. Commun. (1994) 2277. [7] J. Augustynski, A. Carroy, A. Wahl, in: G.J. Meyer, P.C. Searson (Eds.), Nanostructured Materials in Electrochemistry, The Electrochemical Society Proceedings Series, vol. 95-8, Pennington, NJ, 1995, p. 88. [8] A. Wahl, J. Augustynski, J. Phys. Chem. B 102 (1998) 7820. [9] A. Wahl, M. Ulmann, A. Carroy, B. Jermann, M. Dolata, P. Kedzierzawski, C. Chatelain, A. Monnier, J. Augustynski, J. Electroanal. Chem. 396 (1995) 41. [10] J. Nelson, in: A.J. Bard, M. Stratmann, S. Licht (Eds.), Encyclopedia of Electrochemistry, vol. 6, Wiley-VCH, Weinheim, 2003. s, P.E. Schmid, F. Le vy, J. Appl. Phys. [11] H. Tang, K. Prasad, R. Sanjine 75 (1994) 2042. [12] R. Morand, C. Lopez, M. Koudelka-Hep, P. Kedzierzawski, J. Augustynski, J. Phys. Chem. B 106 (2002) 7218. [13] J. Augustynski, Struct. Bonding 69 (1988) 1. [14] R. Solarska, R. Morand, J. Augustynski, Collect. Czech. Chem. Commun. 68 (2003) 1596. [15] Q. Feng, W. Yue, T.M. Cotton, J. Phys. Chem. 94 (1990) 2082. [16] R. Solarska, J. Augustynski, K. Sayama, Electrochim. Acta 52 (2006) 694.

Das könnte Ihnen auch gefallen