You are on page 1of 20

753

Review

Electrocatalytic Properties and Sensor Applications of Fullerenes and Carbon Nanotubes


Bailure S. Sherigara,a,c Wlodzimierz Kutner,*b Francis DSouza*a
a

Department of Chemistry, Wichita State University, 1845 Fairmount, Wichita, Kansas 67260-0051, USA; * e-mail: Francis.DSouza@wichita.edu b Institute of Physical Chemistry, Polish Academy of Sciences, Kasprzaka 44/52, PL-01-224 Warsaw, Poland; * e-mail: wkutner@ichf.edu.pl c On leave from the Department of Industrial Chemistry, Kuvempu University, Jnana Sahyadri, Shankaraghatta 577451, Shimoga, India Received: December 10, 2002 Final version: January 19, 2003 Abstract The electrochemical behavior of fullerene and fullerene derivatives are reviewed with special reference to their catalytic and sensor applications. Recent work on carbon nanotubes, used as catalyst supports in heterogeneous catalysis and sensor development is also presented. An overview of recent progress in the area of fullerene electrochemistry is included. Several cases of electrocatalytic dehalogenation of alkyl halides, assisted by the electrode charge transfer to fullerenes, are discussed. Research work on the electrocatalysis of biomolecules, such as hemin, cytochrome c, DNA, coenzymes, glucose, ascorbic acid, dopamine, etc. have also been considered. Based on the studies of the interaction of fullerenes, fullerene derivatives, and carbon nanotubes with other molecules and biomolecules in particular, the possibilities for the preparation of electrochemical sensors and their application in electroanalytical chemistry are highlighted. Keywords: Fullerenes, Carbon nanotubes, Electrocatalysis, Chemical sensors and Electrode reactions

1. Introduction
Since the first report in 1990 on the electrochemical behavior of C60 [1], several laboratories have contributed to the research on fundamental and application aspects of the fullerene and fullerene derivative electrochemistry. Presently, fullerene science is one of the fastest growing areas of research in chemistry, physics and materials science. In this area, a large number of original research publications and many review articles [2 10] have been published. A review by Bard and others [4] covered the most important general aspects of the solution chemistry of fullerenes and fullerene derivatives. Echegoyen and co-workers [7] have given an up-to-date account of redox properties of pristine fullerenes and a large number of their derivatives, as revealed by electrochemical studies of fullerenes in solution including a section on electrosynthetic methods. A comprehensive review by Reed and Bolskar [6] serves the need for development of the fullerene chemistry with an emphasis on their charged redox states. Kadish and co-workers determined and summarized two especially important factors, viz., half-wave electroreduction potentials, E1/2, i.e., formal redox potentials, E0, [11, 12] and electron affinities which are decisive for redox properties of fullerenes. A descriptive account of fullerene chemistry with a wide ranging topics is also available [13]. However, electrocatalytic properties of fullerenes and carbon nanotubes, CNTs, in particular have not been critically accounted so far, although some sensor
Electroanalysis 2003, 15, No. 9

applications of CNTs have recently been highlighted [14 17]. The structure of the fullerene, C60, is a truncated icosahedron made out of five and six member rings of sp2 carbons [18]. This is unlike graphite, consisting of only fused hexagons of sp2 carbons, which is essentially a planar sheet. Higher fullerenes are also made of five and six member carbon rings [18]. In contrast, the CNTs have closed topology and tubular structure that are typically several nanometers in diameter and many microns in length [19 21]. The CNTs are produced as single-wall carbon nanotubes, SWNTs, and multi-wall carbon nanotubes, MWNTs. SWNTs are made out of a single graphite sheet rolled seamlessly with 1 2 nm in tube diameter in size. MWNTs are composed of coaxial tubules, each formed with a rolled graphite sheet, with diameters varying from 2 to 50 nm and distance between sheets around 0.34 nm. CNTs easily aggregate forming bundles of tens to hundreds of nanotubes in parallel and in contact with each other. These CNTs are conductive and stable in harsh chemical environment [19 21]. The objective of the present article is to review recent investigations on electrocatalysis and sensor applications of fullerenes and CNTs. Electrochemical energy conversion and analytical detection are two major areas related to electrocatalysis [22]. Electrocatalysis is one of several very efficient methods of amplification of a detection signal. Slow electrode reactions of many analytes require potential

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim 1040-0397/03/0906-0753 $ 17.50+.50/0

754
Table 1. Electron affinity, EA, and ionization potential, IP, values estimated for selected fullerenes [230]. Fullerene C60 C70 C76 C78 C82 C84 C84 EA (eV ) 2.7 2.8 3.2 3.4 3.5 3.5 3.3 IP (eV ) 7.8 7.3 6.7 6.8 6.6 7.0 7.0

B. S. Sherigara et al.

2. Electrochemistry of Fullerenes and Their Derivatives


2.1. Electrochemistry of Pristine Fullerenes in Solutions The electron accepting ability of C60 has been explored extensively. Now, it is well documented that both C60 and C70 can be electroreduced by six one-electron reversible steps in non-aqueous solvent solutions at low temperatures [24 26] or in liquid ammonia [27], in agreement with the quantum chemistry calculations [28, 29] and predicted electronic structures of the molecules [30, 31]. Formal potentials of these fullerene redox couples are almost equally separated by as much as ca. 0.4 V, depending on experimental conditions. Moreover, these potentials are largely influenced by solvation [11, 32 37] and, to some extent, by ion pairing [34, 35]. This is why these potentials can be conveniently tuned by selecting proper solvent and/or salt of a supporting electrolyte [11], as depicted in Figure 1 for six C60n/(n1) redox couples. Because of this large potential separation, electrochemically generated fullerene anions are stable in deaerated solutions [38 40]. For these reasons, these anions are attractive as reducing agents, which can be exploited in electrocatalysis. Moreover, these anions can participate in the acid-base equilibria. The fullerene anion is more basic the more negative is its charge. That is, monoanion, C60, is a very weak base while dianion, C602, is a relatively strong base undergoing stepwise protonation; the first of which is easier than the second [41]. For instance, pKa in DMSO is equal to 4.7 and 16 for the first (C60H2/C60H H) and

( D2 isomer) ( C2n isomer) ( C2 isomer) ( D2 isomer) ( D2d isomer)

greatly exceeding their formal redox potentials in order that these reactions proceed at desirably high rates. The acceleration of such kinetically-hindered electrode reactions by the catalysts permits the quantification of these analytes at less extreme potentials, because catalyzed electrode reactions usually occur near the formal redox potential of the catalyst. By applying less extreme potentials, one can improve significantly both detectability and selectivity, as compared to those obtained in the absence of the catalyst [23]. The estimated values of electron affinity, EA, and ionization potential, IP, (Table 1) of C60 and higher fullerenes [7] clearly indicate that fullerenes easily accept and, under certain conditions, donate electrons as well as reveal very rich electrochemistry. These features, required for redox couples to be selected for electrocatalytic applications, make fullerene materials attractive with that respect, on one hand. On the other, however, the combination of electronic properties with a high double-bond character of junction of five and six-member rings provide fullerenes with an ability to undergo addition reactions leading to a large number of mono- and multiple adducts. These addition reactions may compete with the electrocatalytic reactions. A detailed introduction to this review article seemed most obvious to serve as an overview of recent progress in the area of fullerene electrochemistry and related analytical applications. Needless to emphasize that electrochemistry has played a pivotal role as a tool in the study of electronic properties of fullerenes and their derivatives as well as in assessing the properties and stability of many new fullerenebased materials.

Fig. 1. Tuning of formal redox potentials of the C60n/(n1) redox couples with different solvents for 0.1 M (TBA)ClO4 or 0.1 M (TBA)PF6 at a Pt working electrode. (Based on data from [11, 12, 24, 232]).

Table 2. Formal redox potentials, E0', anodic-to-cathodic peak potential differences, DEp (in parentheses), and formal redox potential difference for fullerenes in 0.1 M ( TBA )PF6, in 1,1,2,2-tetrachloroethane, with the ferrocene couple as an internal potential reference and a glassy carbon working electrode (according to [45]). Fullerene E 0', V vs. Fc/Fc and ( DEp ), V
' E0 2=1 ' E0 1=0 ' E0 0=1 ' E0 1=2 ' E0 2=3 ' 0' E0 1=0 E 0=2

C60 C70 C76 C78[a] C78[b] C84

1.75 1.30 1.43 1.17

1.26 1.20 0.81 (0.07) 0.95 (0.07) 0.70 (0.06) 0.93

1.06 1.02 0.83 0.77 0.77 0.67

(0.08) (0.10) (0.10) (0.07)

1.12 (0.07) 1.08 (0.09) 1.08 (0.09) 0.96 (0.08)

1.27 (0.07)

2.32 2.22 1.64 1.72 1.47 1.60

[a] Major isomer, [b] minor isomer.

Electroanalysis 2003, 15, No. 9

Applications of Fullerenes and Carbon Nanotubes

755 2.2. Electrochemical Behavior of Fullerene Derivatives in Solution In spite of their rigid pseudo-spherical cage structures, many different derivatives of fullerenes could be synthesized, including endohedral and exohedral compounds, binary compounds, transition metal complexes, and organic adducts [7, 4]. Endohedral fullerenes are formed by encapsulation of one or more foreign atoms or ions inside a fullerene cage. In metallofullerenes, the fullerene either encapsulates a transition metal atom or cation inside its closed carbon cage, or coordinates it externally to form an endohedral or exohedral compound, respectively. A cyclic macromolecule of sufficiently large ring diameter, such as CD or calixarene, can non-covalently trap a fullerene molecule in its cavity to form a exohedral supramolecular complex [4]. Alternately, a metal atom can replace a carbon atom in a fullerene cage to form so called metcars. 2.2.1. Electrochemical Behavior of Fullerene Adducts in Solution Fullerene adducts have been synthesized by covalent addition across the fullerene double bond to yield binary or tertiary compounds [18, 82], transition metal adducts [82 87], alkylfullerenes [88 91], or fullerene cycloadducts [18, 41]. The ability of C60 and higher fullerenes to easily undergo addition reactions has been largely exploited [18, 41] towards attachment of many different functional groups in an effort to synthesize new materials of electronic properties different than those of the parent fullerenes. In addition, fullerenes have been widely derivatized electrochemically [41, 88, 90, 91]. A shift in the electroreduction or electrooxidation potential is indicative of interaction between C60 and a metal or inductive effect of an addend. The latter involves either electron donating or withdrawing ability of its constituent group or atom, and its orientation. With that respect, certain generalizations have already been formulated. Among them, the important one is the prediction that each sequential loss of the double bond results in a cathodic potential shift of formal redox potential of the derivatized fullerene [90, 92]. In other words, electroreduction becomes more difficult as the fullerene cage becomes increasingly derivatized [93, 94]. Moreover, alkylfullerenes appear to be more stable towards electroreduction than hydrofullerenes [7]. A number of halogenated, viz., fluoro [95], chloro [96], bromo [97] fullerene adducts, have been synthesized. Due to high electronegativity of fluorine, multiflourinated derivatives of C60, such as C60F48 and C60F46, exhibit the highest EA values reported for the fullerene derivatives [82, 100]. Electrochemical properties of C60F36 and C60F46 in 0.1 M (TBA)PF6, in CH2Cl2 have been studied by cyclic voltammetry, CV, [99] (Table 3). For C60F36 the first and second formal redox potentials are located at 0.05 and 0.48 V (vs. Ag/Ag) while for C60F46 these potentials are located at 0.51 and 0.01 V (vs. Ag/Ag) indicating high electron
Electroanalysis 2003, 15, No. 9

second (C60H/C602 H), respectively, dissociation step of dihydrofullerene [42]. For C76, C78 (C2n isomer), five reversible electroreductions were found in a mixed toluene-acetonitrile solution and for all C84 isomers in different solvent solutions [43]. However, only four electroreductions were observed for C78 (D3 isomer) [44]. Generally, the difference in formal potential of the first electrooxidation and first electroreduction, (Eox Ered), is smaller the larger the fullerene size, with the exact order reported [45] being C78(D3) < C84 < C76 < C78(C2n) < C70 < C60 (Table 2). The choice of a solvent, supporting electrolyte, and temperature has a profound impact on both the electroreduction and electrooxidation potentials [38, 46]. Moreover, solubility of the electrochemically generated fullerene ions is dependent on both the nature of solvent and supporting electrolyte selected [38]. The electronegative properties of C60 are further evidenced by formation of fullerides, MxC60 (x 1 to 6) [27], whose preparation involves either a solid-state interaction between C60 and alkali or alkali earth metals [47, 48] or, alternatively, a chemical [49], or electrochemical [50] fullerene reduction. Some of the resulting fullerides reveal superconducting and ferromagnetic properties [47]. Apparently, oxidative electrochemistry of fullerenes is not as rich as reductive electrochemistry. One reason for scarce investigations of the fullerene electrooxidation might be too low positive potential limit, available with the use of the combinations of the solvents and supporting electrolytes studied to-date. Accordingly, one irreversible electrooxidation corresponding to transfer of four electrons was found both for C60 and C70 in 0.1 M tetrabutylammonium hexafluorophosphate, (TBA)PF6, in benzonitrile [46] as well as for C60 in a solid film [51]. It was postulated that chemical reactions coupled to the electron transfer contributed to the irreversibility of the electrochemical reaction. For C60, this irreversible electrooxidation is typical for many other supporting electrolytes [4]. Reversible one-electron oxidations were reported for C60 and C70 in 1,1,2,2-tetrachloroethane, TCE, [52]. For C70, a second irreversible electrooxidation in this solvent solution was also observed. Experimental values of the formal redox potentials of fullerenes indicate that higher fullerenes are more easily oxidized than C60 or C70. Unlike electroreduction, the electrooxidation of fullerenes becomes easier after derivatization [53 57]. Investigation of electrochemical properties of fullerenes in aqueous solutions was only possible after successful water solubilization of these hydrophobic molecules. At least two general routs are available for that purpose, i.e., i) chemical derivatization with highly polar addends [58 63] or ii) transfer of genuine fullerenes to water by formation of supramolecular complexes, like those with cyclodextrins, CDs [65 69], calixarenes [66, 70 76] or by formation of colloids. The latter approach may not use [77], or use surfactants or lipids [78 80].

756
Table 3. Formal redox potentials for fluorinated fullerenes in 0.1 M ( TBA )PF6, in CH2Cl2, at a scan rate of 0.2 V s1 (according to [99]) Compound E0', V (vs. Ag/Ag )
' E0 0=1 ' E0 1=2 ' E0 2=3

B. S. Sherigara et al.

C60 C60F36 C60F46 C60F48 C70 C70F54


[a] Irreversible process.

0.88 0.05 0.51 0.53 0.81 0.76

1.32 0.48 0.01[a] 0.18 1.26 0.36

1.65 0.15[a]

affinity and, consequently, different electronic structure of C60Fx for different number of fluorine atoms attached to the C60 molecule. In case of epoxyfullerene, C60O, during the stepwise transfer of three electrons, [54, 100, 101] the first CV cathodic peak is due to the formation of C60O, which subsequently undergoes decomposition to C60. The second cathodic peak corresponds to irreversible formation of C60O2, which can initiate formation of a C60OC60O polymer. The third cathodic peak represents formation of C60O3. Electrochemical processes of the transition metal derivatives of C60 is accompanied by loss of the metal fragments [82 87]. Symmetry-lowering is responsible for this effect. That is, electroreduction potential was shifted negatively by approximately 0.35 V after sequential addition of metal phosphine groups to the fullerene core indicating a linear decrease of EA with the number of metal phosphine addends [82, 83]. Electrooxidation of monosubstituted complexes of these organometallic derivatives showed stepwise electrochemical disassembly of the molecules [85]. In addition, a change of the metal resulted in a negative shift of the electrooxidation potential, indicating ease of oxidation, which followed the order: Pt < Pd < Ni [85]. As mentioned above, electrochemical properties of a fullerene adduct depend upon the number of addends linked to its cage [53]. This is also the case for alkyl or aryl addends where the C60-centered electroreductions become increasingly difficult and irreversible as the number of addends increases from one to six. In contrast, electrooxidations are easier when going from the mono- to tris-adduct and remains the same for higher adducts. Electronegative hetero-atoms, like O and Si, attached covalently to the fullerene moiety in groups or rings, exhibit an electronwithdrawing effect [54], which increases in the order: O < C < Si, resulting in a negative shift of the electroreduction potential. Electrochemistry of C60 provided some information on properties of the fullerene core itself. Mostly, C60 pseudo-spheres do not interact with each other, but behave rather independently. However, in certain exceptional cases, like for fullerene dimers in which carbon cages are linked closely together, fullerene cores communicate electronically [102].
Electroanalysis 2003, 15, No. 9

Recently, a number of C60-donor and C60-acceptor type molecules have been synthesized and their properties studied [10]. Organofullerenes bearing electron-donor moieties are promising due to their interesting optical and electronic properties. Some of the systems studied are intermolecular C60-donor charge transfer complexes involving organic donors, like ferrocene, or C60-donor polymers, like conducting polymers derived from poly(phenylene vinylene) or oligothiophene, etc., as well as C60-donor dyads, in which C60 is covalently attached to different donor molecules, like dimethylaniline, ruthenium(II) tris(bipyridine) complexes, phthalocynine, terpyridines, porphyrins, carotenoids, calixarenes and copper(I)-complexed rotaxane. The reduction potentials reported on these C60-donor dyads appear at slightly more negative potentials than the parent C60. In the fullerene acceptor dyads, the reduction ability of the compound is improved, as expected. Some of the fullerene acceptor dyads studied include covalently linked C60-quinonoid, C60-tertcyano-1,4-quinodimethane, dicyano-1,4-quinonediimine derivatives, spiroannelated methanofullerenes, nitroaromatic acceptors, cyanodihydrofullerenes, etc. The synthesis and properties of the fullerene derivatives bearing covalently linked donor or acceptor moieties have recently been reviewed [10]. 2.2.2. Electrochemical Behavior of Endohedral Fullerenes in Solutions Smalley and co-workers first prepared a stable endohedral fullerene, viz. lanthanum-doped endohedral C82 fullerene, La@C82 [103]. Theoretical and experimental studies on endohedral metallofullerenes relating to their cage structure and metal position, electronic structure and properties, electrochemical properties and reactivities, their preparation and separation, have recently been reviewed [104, 106]. Ionization potentials and electron affinities (Table 4) are major parameters determining the electronic properties and reactivity of endohedral metallofullerenes. These values indicate that endohedral metallofullerenes are both stronger electron donors and acceptors than the parent empty fullerenes. Arguably, the electron spin resonance, ESR, spectroscopy studies have proven that metal cations rather than atoms are trapped in endohedral metallofullerene carbon cages [106]. Therefore, the cages bare negative charges, i.e., constitute anions, for charge compensation.
Table 4. Electron affinity, EA, and ionization potential, IP, values calculated for M@C82 and fullerenes (according to [229]) Fullerene Sc@C82 Y@C82 La@C82 Ce@C82 Eu@C82 Gd@C82 C60 C70 EA, eV 3.08 3.20 3.22 3.19 3.22 3.20 2.57 2.69 IP, eV 6.45 6.22 6.19 6.46 6.49 6.25 7.78 7.64

Applications of Fullerenes and Carbon Nanotubes Table 5. Formal redox potentials, E0', for endohedral metallofullerenes and empty fullerenes in nonaqueous solvent solutions. Fullerene E0', V
' E0 2=1 ' E0 1=0 ' E0 0=1 ' E0 1=2 ' E0 2=3 ' E0 3=4 ' E0 4=5 ' E0 5=6

757

References

Y@C82 La@C82 La@C82 (minor) Ce@C82 Pr@C82 Gd@C82

1.07 1.07 1.08 1.08 1.08 1.08

0.10 0.07 0.07 0.08 0.07 0.09

0.37 0.42 0.47 0.41 0.39 0.39

1.34 1.37 1.40 1.41 1.35 1.38

1.53 1.53 1.46

2.22 2.26 2.01 1.79 2.21 2.22

2.47 2.46 2.41 2.25 2.48 2.50

[110] [111] [111] [112] [109] [113]

Hence, endohedral fullerenes can also be treated as a new class of organic salts, which cannot undergo electrolytic dissociation in solution without breaking covalent bonds of cages, thus chemically decomposing the molecules. It has been demonstrated that metals of group 2 (M Ca, Sr, Ba) and 3 (M Sc, Y, La), and most lanthanide metals can be trapped inside the cages of higher fullerenes, such as C80, C82, and C84 to form soluble and relatively stable endohedral metallofullerenes, Mm@Cn [107, 108]. Electrochemical properties of endohedral monometallofullerenes are different from those of corresponding empty fullerenes [109 113], as presented in Table 5 [114]. Monometallic endohedral fullerenes are both strong electron donors and acceptors as compared with empty fullerenes [105]. For example, a CV study of Y@C82 showed one reversible electrooxidation and four reversible electroreductions of lower E0' values than those of parent empty fullerenes [111]. Moreover, the electrochemical behavior of many other endohedral metallofullerenes has been examined [114]. The observed differences in electrochemical properties of isomers of minor and major abundance were attributed to differences in their cage structures [114]. To summarize, endohedral metallofullerenes differ significantly in their electronic properties and reactivity from the parent empty fullerenes and are prospective [104, 115] as new materials for catalytic applications.

2.3. Electrochemical Behavior of Fullerenes in Modified Electrodes Chemically modified electrodes, CMEs, have attracted considerable interest during the past two decades as researchers have attempted to exert more direct control over the chemical nature of the electrode. The chemistry of electrode surface modifications and their relevance to the various areas of study of modified surfaces was discussed by Murray [116]. Certain theoretical and practical aspects, like chemisorption of aromatic molecules from aqueous solution at well defined electrode surfaces [117], modification of electrode surfaces with self-organized electroactive microstructures [118], dynamics of electron transport in polymeric assemblies of redox centers [119], catalysis at redox polymer coated electrodes [120], electrodes modified with clays,

zeolites, and related microporous solids [121], electron transfer mediation in metal complex polymer films [122], mass and charge transport in electronically conductive polymers [123], have been described. The terminology, definitions and preparation methods of CMEs have been described and classified as well as recommendations provided in an IUPAC report [124]. CMEs have found numerous important applications in, e.g., solar energy conversion, selective electroorganic synthesis, molecular electronics, electrochromic display devices, corrosion protection, and electroanalysis. In a more recent IUPAC report, analytical aspects of chemically modified electrodes were critically reviewed [23]. That is, effects of analyte and/or reagent accumulation, chemical transformation, electrocatalysis, permeability, ionic equilibria, controlled release, and change of mass, as well as combination of these aspects were evaluated and classified. Also, relevant definitions were provided and recommendations formulated for the most effective CME operations. A review of literature pertaining to fullerene-based thin film materials [125] and electrochemical behavior of the fullerene thin films have recently been reviewed [3, 4, 126]. The idea of introducing a C60 chemically modified electrode to the electrochemical research was first presented by Compton and co-workers [127]. Their CMEs were prepared by immobilizing C60 films by drop coating onto surfaces of the noble metal electrodes, which were then coated with the Nafion protecting films. That way, the amount of C60 required for performing electrochemical experiments was reduced and a current signal enhanced as compared to those using C60 dissolved in solution. Afterwards, electrochemical behavior of the C60 CMEs, initially in non-aqueous solvent solutions [126 134] and later in aqueous solutions [135, 136], has been widely investigated providing the possibility of their electroanalytical applications [135, 137]. Now, fullerene-based CMEs are prepared in several different ways. The most common, simplest and efficient procedure involves the electrode drop coating by using a fullerene solution of a volatile solvent, such as dichloromethane or chloroform [129, 130, 138]. However, the smallcrystalline films prepared that way are usually non-uniform and porous. Moreover, morphology of these films depends on the rate of solvent evaporation, concentration of the
Electroanalysis 2003, 15, No. 9

758 casting solution, nature of the solvent and roughness of the electrode surface. Most often, drop coated films suffer from residual small amounts of the solvent entrapped in the fullerene microcrystallites [139, 140], which may influence voltammograms of the first few CV cycles [129, 141]. An efficient CME preparation procedure involves vapor deposition of fullerene films [132 135] which results in films of high uniformity and purity. However, the resulting films are composed of fullerene aggregates showing rich vibrational features in the visible range of the electronic UV-vis spectra [142]. Electrochemical deposition is yet another technique of the fullerene film preparation. The electrochemically grown films are more uniform than the solution cast films. But, properties of both film types are comparable. Constantpotential bulk electrolysis of C60 in an electrolyte solution of low-polarity solvent results in anions C60n, which are then spontaneously deposited onto the electrode surface as an insoluble salt of cation of the supporting electrolyte [131, 132]. Conversely, similar electrolysis, performed on a vigorously agitated C60 suspension in a polar solvent solution, results in a soluble anion, C60n. Then, this anion is electrooxidized at sufficiently positive potential resulting in deposition of a neutral C60 film onto the electrode surface [132, 144]. Recently, it was demonstrated that C60, C70, and their derivatives form optically transparent nanoscopic clusters of diameter 100 300 nm in a mixed toluene and acetonitrile (1 : 1, v : v) solvent solution at room temperature [144 147]. These clusters could be charged under the influence of a DC electric field and get deposited onto the positive electrode surface. The thickness and morphology of the resulting film could be controlled by the deposition voltage and the fullerene concentration. The ability to assemble the fullerene clusters as three-dimensional arrays opens up new avenues to design high-surface-area electrode materials that are promising for developing chemical sensors and lightenergy converting devices [148]. Moreover, fullerene CMEs can be conveniently prepared by electropolymerization leading either to homopolymers with C60 pseudo spheres linked together via [2 2] cycloaddition [149] or heteropolymers [96, 97, 150 158] in which C60 cages are linked to side polymer chains [157, 158] or spaced with heteroatoms. The latter heteropolymers are prepared from C60 epoxides [100 101 or from genuine C60 in the presence of either traces of oxygen [157, 158], or different transition metal complexes [152 156]. Importantly, the procedure selected for the film preparation influences the fullerene electrochemical behavior, as it leads to films of different solvent and counter cation content, porosity and crystallinity [129, 159]. To-date, electroreductive rather than electrooxidative studies of the C60 film behavior were more successful [160, 161] because the electrooxidation steps are irreversible and reaction products more difficult to identify. The film electroreduction involves incorporation of cations for neutralization of electrochemically generated fullerene anions. Therefore, the nature of cation of the supporting
Electroanalysis 2003, 15, No. 9

B. S. Sherigara et al.

electrolyte determines both electrochemical properties and stability of the film with respect to dissolution [132]. Generally, properties of a new fullerene phase, such as solubility, conductivity or kinetics of phase transition, affect electrochemical properties of the film. The nature of solvent mainly affects solubility of the electroreduced film. That is, resulted fullerene anions dissolve easily from the film in many non-aqueous polar solvent solutions [162, 163]. Unlike the dissolved fullerenes, which show multiple reversible electroreduction steps, only the first two electroreductions show chemical reversibility in films [131, 132, 164]. Extension of the CV negative potential reversal beyond the first, or in some cases the second electroreduction potential in certain electrolyte solvent systems, results in the film dissolution. Prolonged CV potential cycling leads to a gradual loss of the film electroactivity [129]. The SEM study indicated that this loss is, presumably, due to formation of an electrochemically inactive, resistive crystalline C60 phase formed electroreductively [134]. In this study, even trace amounts of impurities could change the long-time CV behavior of the film, because the fullerene was used in a nanomole or subnanomole quantity for its preparation. Another important inference made was that the time-dependent phase transformations in films were responsible for different dependencies of the CV peak potentials on the potential scan rate [126, 127]. Intrinsic electrochemical behavior of the fullerene films have been investigated by using suitably selected techniques, including scanning electrochemical microscopy (SECM), simultaneous electrochemistry and piezoelectric microgravimetry with the use of an electrochemical quartz crystal microbalance (EQCM), laser desorption mass spectrometry (LDMS), surface enhanced Raman spectroscopy (SERS), X-ray photoelectron spectroscopy (XPS), ESR spectroscopy, UV-vis and FTIR spectro-electrochemical techniques, optical microscopy, conductivity measurements, scanning electron microscopy (SEM) and atomic force microscopy (AFM), as well as X-ray powder diffraction (XRD). Details of these studies are summarized elsewhere [4]. Splitting of the CV cathodic and anodic peaks was considerably decreased, as compared to that for a polycrystalline C60 film, by isolating C60 as the Langmuir-Blodget films [129]. For that purpose, a drop of a dilute C60 solution of benzene was spread onto the surface of an aqueous phase to form the Langmuir monolayer or multilayer films, depending on the surface pressure. Subsequent Langmuir-Blodgett vertical dipping of a hydrophobic electrode, such as the alkylthiol modified Au electrode, lead to transfer of these films. Another method of the solid-phase isolation of C60 films involved formation of self-assembled monolayer (SAM) films on the electrode surfaces [129, 165]. SAMs approach is useful for constructing highly ordered, two- and threedimensional structures on solid substrates. Several groups have already reported on the formation, electrochemistry, photo-electrochemistry, and surface properties of the fullerene SAMs. These SAMs have displayed interesting

Applications of Fullerenes and Carbon Nanotubes

759

physical and chemical properties, such as electrochemical reversibility, fewer solubility problems and more uniform electrode coverage than polycrystalline C60 films deposited onto electrodes [166 176]. Recently, the multiple electron transfer reactions of higher fullerene, C84, embedded in surface films of cationic amphiphiles, tetraoctylammonium bromide or didodecyldimethylammonium bromide, were described [177]. The report highlights structural aspects and electron transfer properties of other higher fullerenes and metallofullerenes in lipid films on electrodes.

Scheme 1.

3. Electrocatalytic Activity and Sensor Applications of Fullerenes and Carbon Nanotubes


3.1. Fullerenes in Electrocatalytic and Sensor Applications Because of lowering the potential of electroreduction of many different redox substrates and increasing the reaction rates, improving thus sensitivity and selectivity, fullerenes have been used both as redox catalysts and mediators in electrochemical catalysis. Here, the reaction components are dissolved in solution or, in some cases, substrates dissolved in solutions are chemically transformed at the electrode modified with a fullerene moiety. Such a CME sensor could be used for biomolecular sensing and environmental monitoring [178, 179]. Huang and Wayner first applied C60 anions for homogenous catalysis [180]. They determined the rate constants for the intermolecular electron transfer from C60 mono-, di- and trianions to electron-deficient organic dibromides, such as 1,2-dibromotetrachloroethane, and vicinal dibromides, such as a,a'-dibromo-p-xylene, ethyl-2,3-dibromopropionate, ethyl-2-bromopropionate, and ethyl-2-bromoisobutyrate (Table 6). The determined reactions rates were compared to those of other mediators of the same E0' value. Regeneration of the oxidized form of the C60 catalyst at

the electrode surface was responsible for the observed CV electrocatalytic cathodic currents, as shown in Scheme 1. Moreover, selectivity of electrocatalytic reduction of the bromides and dibromides by C60 anions has been highlighted. Electrochemically generated anions of C60 were used as bases for deprotonation of several organic acids in 0.05 M (TBA)Br, in a mixed acetonitrile-toluene (2 : 3, v : v) solvent solution under CV conditions [181]. A monoanion, C60, weak base was able to deprotonate a relatively strong CH acid, like ethyl nitroacetate. Moreover, this monoanion efficiently catalyzed reaction of ethyl nitroacetate with ethyl acrylate and acetonitrile to form corresponding doubleaddition products. However, a dianion, C602, stronger base was required for deprotonation of weaker acids, like diethyl malonate, diethyl methylmalonate, 2-nitropropane and noctanethiol. This dianion promoted reaction of diethyl malonate with acrylonitrile and nitromethane with ethyl acrylate, which lead to formation a double- and tripleaddition product, respectively. Moreover, the dianion catalyzed the addition of n-octanethiol to styrene oxide. The highly basic trianion, C603, promoted reactions of still weaker acids, like addition of pyrrole to acrylonitrile and Witting-Horner reaction of diethyl benzylphosphonate with banzaldehyde. Various organic halides were indirectly reduced on a preparative scale in nonaqueous solvent solutions by using C60 anion catalysts [182]. For that purpose, cathodically

Table 6. Values of cathodic peak potential, Epc, as well as formal redox potential, E0', and catalytic rate constant, kcat, for reduction of bromides and vicinal dibromides by C60 anions (according to [180]) Bromide[a] (Epc vs. Ag/AgCl) 1,2-Dibromoterachloroethane ( 0.84) Ethyl 2,3-dibromopropionate ( 1.41) a,a'-Dibromo-p-xylene ( 1.48) Ethyl 2-bromopropionate ( 1.84) Ethyl 2-bromoisobutyrate ( 1.99) Catalyst (oxidized form) C60 C60 C60 C60 C60 C60 C602 C60 C602 E0' [a],[b] V vs. Ag/AgCl 0.416 (59) 0.416 (59) 0.836 (59) 0.416 (59) 0.836 (59) 0.836 (59) 1.378 (61) 0.836 (59) 1.378 (61) kcat [c] M1 s1 2.2 0.1 103 nr 1.2 0.1 104 nr 22.8 0.1 nr 8.1 2 102 nr 4.5 2 104

[a] Potential scan rate: 50 mV s1; formal redox potential for the Fc/Fc couple in a toluene-acetonitrile (5.4 : 1, v : v) solution was 0.571 V. [b] The value of Epa Epc, in mV, is shown in parentheses. [c] Average rate constant from two different scan rates using three different concentrations of bromides; nr: no reaction observed up to 19 equivalents of added bromide.

Electroanalysis 2003, 15, No. 9

760 generated C60 anions were used as efficient electron carriers. Moreover, C60 was active as a catalyst for the preparativescale indirect cathodic reduction in the presence of aromatic redox mediators, such as anthracene. Reducing ability of electrochemically generated six C60 anions, C60n (n 1 to 6) is different. Therefore, reducing ability of the C60 catalyst was conveniently controlled, without any chemical modification, by electrochemical generating appropriate C60 anion. So, C60 was projected as a tunable reducing organic reagent (Fig. 1). Electrocatalytic reduction of different a,w-diiodoalkanes was investigated by using C60n generated electrochemically in 0.1 M (TBA)PF6, in benzonitrile, and in the C60 solid film coated electrode [183]. The C602/3 and C60/2 couples and not the C600/ couple were electrocalytically active. The second-order rate constant, kcat, of the electrocatalytic reduction of a,w-diiodoalkanes was detrmined from the RDE voltammetry catalytic limiting currents. The RDE limiting current plateaus were well defined for all three electroreductions. For the excess of a,w-diiodoalkanes with respect to C60, i.e., under pseudo-first-order conditions with respect to a,w-diiodoalkanes, kcat is given by [184],
2 / (nFACR)2DRCS kcat icat

B. S. Sherigara et al.

(1)

where icat is the catalytic limiting current, DR and CR are the diffusion coefficient and concentration of C60 catalyst, respectively, CS is the concentration of the substrate, a,wdiiodoalkane, F is the Faraday constant, n is the number of electrons transferred, and A is the electrode surface area. The values of icat were determined by extrapolating the curves of the dependence of limiting currents vs. the square root of the rotation rate to the zero abscissa. The determined kcat values increased in the order: C5 < C8 < C4 < C3 < C6 < < C2 for the number of carbon atoms of the alkyl chain of a,w-diiodoalkanes. After subsequent potential cycling under CV conditions over a positive potential range, a new anodic peak was found. This peak was ascribed to electrooxidation of iodide released from the substrates during the preceding electroreduction. Electrocatalytic reduction of a,w-diiodoalkanes at the C60 drop-coated film modified electrode in 0.1 M (TBA)PF6, in benzonitrile, showed different behavior than that observed for C60 in solution [183]. That is, only the first two, C600/ and C60/2, electroreductions could be examined, because the film dissolved when the negative potential reversal was extended beyond the potential of the third, C602/3, electroreduction. For the C600/ couple, electrochemical behavior was steady state under multi scan CV conditions, even after prolonged potential cycling. The cathodic and anodic peaks were largely separated and peak currents were nearly equal. But the first two cathodic peaks increased markedly in the presence of 1,2-diiodoethane, which indicated the electrocatalytic reduction of the a,w-diiodoalkane substrate. The GC-MS and HPLC analyses of the electrocatalysis products, obtained by constant-potential bulk electrolyses, revealed formation of alkanes and monoiodo alkanes, for the investigated series of a,w-diiodoalkanes, except for the
Electroanalysis 2003, 15, No. 9

reaction of C603 with either 1,3-diiodopropane or 1,5diiodopentane where C60 adduct formation was favored. Moreover, a,w-dihalogenated alkanes were electrocatalytically reduced by anions of a higher fullerene, C70 [185] (Fig. 2). That is, both di- and trianons of C70 electrocatalytically reduced a homologous series of a,w-dihaloalkanes, X(CH2)mX (X Cl, Br, I), in 0.1 M (TBA)PF6, in benzonitrile, at a Pt working electrode. The second-order rate constants for electrocatalytic dehalogination of the a,wdihaloalkanes by the C70n (n 2 or 3) catalysts were determined by using the RDE voltammetry under pseudofirst-order conditions with respect to the a,w-dihaloalkanes. The calculated kcat values increased in the order: Cl < Br < I for the investigated 1,2-dihaloethanes. Generally, the kcat values for C70n (n 2 or 3) catalysts were lower than those for the C60n (n 2 or 3) catalysts. For the C70n catalysts, log kcat monotonically decayed with the number of carbon atoms in alkyl chains of the a,w-dihaloalkanes (Fig. 3). The GC-MS and HPLC analyses of the electrocatalysis products, obtained by the constant-potential bulk electrolyses, mostly revealed mixtures of alkanes and alkenes both for the 1,2diiodoethane and 1,3-diiodopropane substrate while mostly monohalogenated alkanes were the reduction products for other a,w-dihaloalkanes [186]. No reaction between C70n (n 2 or 3) and a,w-dihaloalkanes leading to alkyl adducts of C70 was observed, neither on the CV and RDE voltammetry nor on the bulk electrolysis time scale. This result was unlike those for the C60 electrocatalyses [183, 185], where corresponding electrocatalytic reductions were accompanied by chemical reactions between C60n and certain a,wdihaloalkanes yielding alkyl adducts of C60 [187]. Electro-

Fig. 2. Cyclic voltammograms for 1) 0.1 mM C70, 2) 0.1 mM C70 and 0.4 M 1,2-dichloroethane, 3) 0.1 mM C70 and 0.4 M 1,2dibromoethane, as well as 4) 0.1 mM C70 and 1.4 mM 1,2diiodoethane in 0.1 M (TBA)PF6, in benzonitrile, at a Pt disk electrode; potential scan rate 0.1 V s1. (Adapted from [185]).

Applications of Fullerenes and Carbon Nanotubes

761 serve as better electrocatalysts than anions of C60 [185]. All di- and trianions of C60, C70, C76 and C78 as well as trianions of C84, electroreductively generated in 0.1 M (TBA)PF6, in benzonitrile, catalyzed dehalogenation of 1,2-dihaloethanes and the potential of this dehalogenation was governed by the E0' values of the fullerene used. Importantly, the higher the fullerene, the more positive was the E0' value of electrocatalytic reduction of 1,2-dihaloethanes. The kcat values for these electrocatalytic reductions, determined by the RDE voltammetry under pseudo-first-order conditions with respect to 1,2-dihaloethanes, showed a Brnsted-type dependence of log kcat versus E0' of the fullerene for the electrocatalytic reduction of 1,2-dihaloethanes (Fig. 4). For each fullerene anion, kcat increased in the order: Cl < Br < I, as predicted from the theory of dissociative reduction [189]. In addition, the kcat values increased in the order: C84 < C78 < C76 < C70 < C60 for each 1,2-dihaloethanes [188] (Table 7). The anions of higher fullerenes did not react with 1,2-dihaloethanes. So, no alkyl adducts of fullerenes were formed. Because of high stability with respect to the adduct formation and more positive E0' values of the electrocatalysis, the higher fullerenes may be more useful than C60 as catalysts, even though the corresponding catalytic rate constants are lower for these fullerenes. Noticeably, homogeneous electron transfer between C602 and alkyl bromides depended upon solution conditions [89]. That is, C60 was first reduced stepwise by 1,4-benzoquinone

Fig. 3. Semi-logarithmic dependence of the second-order catalytic rate constant, kcat, for reduction of a?w-diioalkanes on the number of carbon atoms in their alkyl chains for a) C702 and C703 as well as, b) C602 and C603 catalysts. (Adapted from [185]).

catalytic efficiency was higher the higher was the negative charge of C70 anion. The proposed mechanism of electrocatalytic reduction of a,w-dihaloalkane substrates by the C70n (n 2 or 3) catalyst is summarized in Scheme 2. Structure-reactivity emphasizing electrocatalytic study, which involved higher fullerenes, i.e., C76, C78 and C84, and 1,2-dihaloethanes as substrates [188], confirmed the hypotheses that anions of higher fullerenes are relatively more stable with respect to derivatization and, therefore, may

Scheme 2.

Fig. 4. Dependence of log kcat versus E1/2 for a) the second electroreduction, and b) the third electroreduction of the C60, C70, C76, C78, and C84 fullerenes for 1) 1,2-diiodoethane, 2) 2-dibromoethane, and 3) 1,2-dichloroethane. (Adapted from [190]).
Electroanalysis 2003, 15, No. 9

762

B. S. Sherigara et al.

Table 7. Values of the formal redox potential, E0', and second-order rate constant, kcat, for electrocatalytic reduction of 1,2-dihaloethanes by C n m ( m 60, 70, 76, 78, 84, and n 2 or 3) in 0.1 M ( TBA )PF6, in benzonitrile, determined by the RDE voltammetry. Fullerene redox couple C60 C70/2 C76/2 C78/2 C84/2 C602/3 C702/3 C762/3 C782/3 C842/3
/2

E0' [a] V vs Ag/AgCl 0.86 0.82 0.63 0.47 0.42 1.34 1.26 1.12 0.84 0.78

kcat, [b] M1 s1 Cl( CH2)2Cl 3.5 10 2.2 2.1 0 0 5.5 10 3.3 3.2 0 0 Br( CH2)2Br 1.6 10 3.3 10 2.5 10 10 0 3.2 102 1.3 102 8.3 10 8.9 1.7
2

References I( CH2)2I 2.8 105 1.1 105 3.3 104 9.8 103 0 9.4 105 5.1 105 5.1 104 1.6 104 1.9 102 [183] [185] [198] [198] [198] [183] [187] [198] [198] [198]

[a] The E0' values determined in agreement with the literature values [21]. For the Fc/Fc couple, E0' 0.45 V (vs. Ag/AgCl) under the used solution conditions. [b] The estimated error of the kcat values is 15% (standard deviation).

dianion in the acetonitrile or benzonitrile solution. Subsequently, the generated C602 was used as the starting material in the synthesis of RxC60 (where x 2 for R C6H5 CH2 and x 1 for R o-xylyl) by the reaction of C602 with benzyl bromide or a,a'-dibromo-o-xylene, as shown in Scheme 3. Figure 5 illustrates the trend of the log kcat change against the number of carbon atoms on the fullerene cage. In one study, the C60 CME was prepared as a C60 film drop coated onto the electrode surface which was overlaid with a Nafion protecting coat [190]. For this CME, four quasireversible one-electron electroreductions were observed by CV for a 2% (C2H5)4NOH water-acetonitrile (7 : 3, v : v) solution. However, only three electroreductions appeared for a 2% (C2H5)4NOH water-DMF (3 : 2, v : v) solution. Moreover, the more negative was the E0' value, the more reversible was the electrochemical process. Oxygen dissolved in a 2% (C2H5)4NOH water-DMF (3 : 2, v : v) solution was electrocatalytically reduced at this CME and, therefore, this electrode may find practical application as a sensor.

Fig. 5. Dependence of log kcat versus number of carbon atoms of the fullerene cage for the fullerene 1) dianions and 2) trianions, in the presence of 1,2-diiodoethane in 0.1 M (TBA)PF6, in benzonitrile. (Adapted from [198]).

Scheme 3.
Electroanalysis 2003, 15, No. 9

In another study, C60 CME was applied for electrocatalytic reduction of nitrobenzene in a mixed organic-aqueous solvent solution [191]. Initially, electrochemical behavior of a C60 modified carbon paste electrode (CPE) in 2% (TBA)OH, in an acetonitrile-water (3 : 2, v : v) solution was examined. Two pairs of quasi-reversible one-electron cathodic and anodic CV peaks were recorded and conditions optimized to reach the highest electrochemical stability. The effect of C60 content in this CPE and solution composition on the electrocatalytic reaction was examined. If the C60 content in the carbon paste electrode was too high (20%) or too low (1%), then the sensitivity was low due to the presence of a large background current or insufficient amount of the C60 catalyst, respectively. For a 10% content of C60, sensitivity of this CPE was the highest. Similarly, the electrocatalytic efficiency was low if the amount of the acetonitrile content was too low (40%) or too high (80%). The second cathodic peak revealed an electrocatalytic response with respect to reduction of nitrobenzene. The semi-differential catalytic current difference was linearly

Applications of Fullerenes and Carbon Nanotubes

763 bin was electrocatalytically reduced at C60-(g-CD)2/Nafion CME, similarly as at C60-(g-CD)/Nafion CME [196]. The fullerene aqueous colloids were used for preparation of electrochemically stable films, deposited onto electrode surfaces, to form CMEs suitable for electrocatalysis [78, 80, 197]. That way, hemoglobin was electrocatalytically reduced in an aqueous solution at CME, prepared by casting a film of C70 and cationic surfactant, didodecyldimethylammonium bromide, onto GCE, C70-1 CME [198]. Composition and structural integrity of the C70-1 films were examined by SEM, XPS, and UV-vis spectroscopy. In 0.4 M KCl, C70-1 CME revealed two pairs of quasi-reversible diffusion controlled cathodic and anodic CV peaks corresponding to the C700/ and C70/2 redox couples, respectively. The release of solvent, which was entrapped during film casting and the ion-pair formation between the C70 anion and the cation from solution entering the film during electroreduction of the C70 sites, was accompanied by structural rearrangements in the film. For C70-1/GCE in 0.4 M KCl in the presence of hemoglobin, the CV curve, recorded between 0 and 0.6 V, showed significant increase and decrease of the cathodic and anodic peak, respectively. This electrocatalytic behavior was ascribed to the presence of the surfactant, whose apparent role was to change microenvironment of the electrode surface and enhance electron transfer to C70. Electrodes, functionalized with the electron-mediating monolayer or thin multilayer films, offer active interfaces capable to electrically communicate redox enzymes with electrodes and to activate the respective biocatalyzed transformations [199]. Accordingly, an Au electrode, coated by a C60 monolayer, activates the biocatalyzed electrooxidation of glucose with glucose oxidase, GO, [200]. Poor solubility of low-polarity C60 derivatives in aqueous solutions prevents their applications as the diffusion charge mediators. Instead, the C60 monolayer films, immobilized onto an electrode surface, was able to mediate electron transfer between the enzymes and electrodes. For that purpose, the Au electrode was primarily modified by a cystamine monolayer film. Then, 1,2-dehydro-1,2-methanofullerene[60]-61-carboxylic acid was covalently bound to it. The observed electrocatalytic CV anodic currents, transduced by the C60 modified Au electrode in the presence of GO (2 mg mL1) and glucose of different concentrations (0 100 mM) in 0.1 M phosphate buffer, pH 7.1, indicated that glucose was biocatalytically electrooxidized. In this process, the flavin adenine dinucleotide (FAD) active site of the GO enzyme was reduced by glucose to FADH2 and, then, the C60 moiety mediated oxidation of FADH2 regenerating thus the oxidized active site. Scheme 5 shows both the assembly of the C60 derivative monolayer attached to the Au electrode and biocatalytic electrooxidation of glucose in the presence of GO. Electrochemical behavior of DNA was studied at the C60-(g-CD)/Nafion CME in aqueous solution [201]. Watersoluble C60-(g-CD) inclusion complexes were prepared by the earlier described procedures [202, 203]. The watersoluble C60-(g-CD) film, deposited onto GCE, was overlaid
Electroanalysis 2003, 15, No. 9

dependent on concentration in the range 5 105 to 6.0 103 M nitrobenzene. The lower detection limit was as low as 3.0 105 M nitrobenzene. The host-guest complexes of fullerenes with CDs and calixarenes are interesting supramolecular systems revealing electrocatalytic activity with respect to some slow electron transfer reactions. Earlier works showed that solution electrochemistry of supramolecular complexes of fullerenes with cyclodextrins and calixarenes were different from those of fullerenes themselves [58, 62 68]. In recent years, Li and co-workers [194] have performed a detailed study of electrochemistry of these supramolecular complexes of fullerenes. For the C60-(g-CD)/Nafion and C60-(gCD)2/Nafion CMEs (made of a water-soluble C60-(g-CD) complex, deposited as a film on to a glassy carbon electrode, GCE, then protected with a Nafion coat), several reversible CV one-electron electroreductions in aqueous supporting electrolytes were observed. Their behavior was typical for surface-confined redox couples [195]. The CMEs revealed appreciable stability and reproducibility under CV conditions in 0.1 M KCl, in the potential range 0.20 to 0.80 V (vs. SCE) and, moreover, electrocatalytic activity towards some biomolecules, such as hemoglobin, cytochrome c, and hemin. Cytochrome c was electrocatalytically reduced at the C60(g-CD)/Nafion CME in aqueous solutions [196]. The rate of electron transfer to cytochrome c at a bare GCE, in 0.1 M KCl, was very low. But CV curves, recorded in the presence and absence of cytochrome c, showed that C60-(g-CD)2 is able to shuttle charge between the electrode and cytochrome c. Electrocatalytic activity of the C60-(g-CD)/Nafion CME was much lower than that of the C60-(g-CD)2/Nafion CME. The proposed mechanism of elecrocatalytic reduction of cytochrome c at these CMEs is shown in Scheme 4. Hemoglobin was electrocatalytically reduced at C60(g-CD)/Nafion CME [197]. Electroreduction of the C60(g-CD)/Nafion CME in 0.1 M KCl yielded a C60-(g-CD) radical monoanion in a one-electron process. Repeated potential cycling produced no significant decrease in the cathodic and anodic peak currents indicating appreciable stability and reproducibility of this CME under multiscan CV conditions. However, its performance was influenced by the solution acidity, which was adjusted with 0.1 M acetate buffer in the range 4.5 pH 6.2. Apparently, the electrode reaction was accompanied by a proton transfer. The C60-(gCD) complex binds to the heme protein, causing its rearrangement due to week Van der Waals interactions, thereby promoting the electron transfer. The electroreduction mechanism involves transfer of one electron to hemoglobin. Under the same solution conditions hemoglo-

Scheme 4.

764

B. S. Sherigara et al.

Scheme 5.

with a thin Nafion coat in order to prevent the film dissolution. A CV curve exhibited single redox couple for the C60-(g-CD)/Nafion/GCE in 0.1 M KCl. The cathodic peak current was linearly dependent on the potential scan rate, as expected for the surface-confined redox couple revealing finite diffusion behavior. Repeated potential cycling revealed stability and reproducibility of the CME sufficient for electrocatalytic reduction of DNA. That is, both the cathodic and anodic peak increased but peak potentials changed only slightly for the C60-(g-CD)/Nafion/ GCE in 0.1 M KCl in the presence of DNA. Scheme 6 shows a two-way mechanism of electrocatalytic reaction of DNA at the C60-(g-CD)/Nafion/GCE. A film of a supramolecular complex, (C70)2-(4-tert-butylcalix[6]arene), (C70)2-2, was deposited onto GCE for electrocatalytic reduction of nitrite [204]. The (C70)2-2/ GCE featured two redox couples in 0.04 M (TBA)ClO4, in an acetonitrile-water (1 : 1, v : v) solution. The first redox couple, being fairly stable, was used for the reduction of nitrite. Importantly, direct electroreduction of nitrite requires a very negative potentials at most unmodified electrodes. But nitrite was electrocatalytically reduced at (C70)2-2/GCE in the potential range 0 to 0.75 V (vs. SCE). In order to find optimum electrocatalytic performance,

nitrite was also electrocatalytically reduced in mixed acetonitrile-water solvent solutions containing phosphate buffer, 7 pH 8. Moreover, lower potential scan rates appeared beneficial for the electrocatalysis. The electrocatalytic efficiency, represented by the ratio of catalytic current to the cathodic peak current for the (C70)2-2 film in solution not containing nitrite, decreased with an increase of the scan rate being another evidence of electrocatalysis. A proposed mechanism of the electrocatalysis is shown in Scheme 7. In order to explore possibility of analytical application, the electrocatalytic property of a complex of C60 with octasodium calix[8]aryloxy octakis-(propane-3-sulfonate), 3, immobilized in a Nafion film, (C60-3)/Nafion, was investigated towards determination of chloroacetic acid [205]. A CV behavior of the C60-3/Nafion CME was stable in 0.05 M (TBA)ClO4, in a mixed acetonitrile-water (3 : 2, v : v) solution. The first electroreduction and re-oxidation of the (C60-3)/Nafion CME was a one-electron process and the (C60 -3)0//Nafion redox system was capable to dehalogenate electrocatalytically chloroacetic acid. A mechanism of electrocatalytic reduction of this halogenated acid by the

Scheme 6.
Electroanalysis 2003, 15, No. 9

Scheme 7.

Applications of Fullerenes and Carbon Nanotubes

765 two- and four-electron transfer, respectively. Both cathodic CV signals were electrocatalytically active towards reduction of chloroacetic acid. The (C70)2-4/GCE was also applied for catalytic CV electroreduction of nicotinamide adenine dinucleotide (NAD) in an acetonitrile-water (1 : 1, v : v) solution [209]. The study was undertaken in an effort to understand the role of the co-enzyme redox couple, NAD/NADH, in biological redox processes involving transfer of proton and two electrons from a substrate to the co-enzyme. The (C70)2-4/ GCE revealed two pairs of cathodic and anodic CV peaks in 0.04 M (TBA)ClO4, in an acetonitrile-water (1 : 1, v : v) solution, as (C70)2-4 was stepwise electroreduced to [(C70)2-4]2, then to [(C70)2-4]6. These electroreductions were accompanied by the TBA cation ingress into the film for neutralization of the electrochemically generated negative charge of the complex, as each C70 molecule in the complex accepted one and two electrons at the first and second electroreduction step, respectively. In the presence of NAD, the peak currents decreased from cycle to cycle. Presumably, the reduction of NAD was electrocatalyzed at the surface of (C70)2-4/GCE. However, the TBA cations partitioning into the film for charge compensation might be simultaneously ion-exchanged in part by the NAD cations from solution. The electrocatalysis product, NADH, might stay in the film decreasing thus its electrocatalytic activity. A C60-glutathione modified Au electrode, C60-5/Au, has been characterized electrochemically and applied for electrocatalytic oxidation of NADH [210]. In this CME, C60 was anchored covalently to 5, which formed a SAM film on the Au electrode surface. Structural integrity of this CME assembly was established by the FTIR spectroscopy, CVand electrochemical impedance spectroscopy (EIS) studies. The pinholes in the C60-5 monolayer film showed an openingclosing response to the change of the solution acidity. Moreover, this response resulted from modification of the C60-5 monolayer structure in the course of the CV and EIS investigations. The NAD/NADH dependent dehydrogenases consists of a large group among redox enzymes known to date. These enzymes have aroused considerable interest for the biosensor development. However, electrooxidation of NADH requires high positive potentials, such as 1.10 V and 1.30 V (vs. SCE) at a GC and Pt electrode, respectively. Therefore, the C60-5/Au electrode was employed to electrocatalyze oxidation of NADH. That way, NADH was determined via catalytic oxidation of NADH by cation radical of the fullerene moiety, C60. Additionally, solution acidity drastically affected electrocatalytic activity of this CME toward oxidation of NADH, which might be related to structural changes in the film. The fullerene-ferrocene, C60-Fc, film CME revealed electrocatalytic behavior [211]. The C60-Fc adduct has been deposited in forms of the nanocrystalline films onto both a conducting glass optically transparent electrode and GCE by electrophoresis in a mixed acetonitrile-toluene (9 : 1, v : v) solvent solution. The resulting CMEs revealed irreversible CV electroreduction of the C60 moiety when the
Electroanalysis 2003, 15, No. 9

Scheme 8.

(C60-3) catalyst could be that presented in Scheme 8. The C60-3 complex might be applied for the electrocatalytic dehalogenation of organohalide pollutants in environmental analysis. In a more extensive study, the (C60-3)/Nafion CMEs were applied [206] for the CV electrocatalytic reduction of a series of halogenated organic acids, such as monochloroacetic acid, dichloroacetic acid, trichloroacetic acid, 2- or 3bromoacetic acid, 2-bromopropionic acid, 2-bromobutyric acid, and 4-bromobutyric acid in mixed acetonitrile-water (3 : 2, v : v) solutions. The (C60-3)/Nafion CME was able of mediating electron transfer from the electrode to the halogenated organic acids, except to 4-bromobutyric acid. Values of the rate constant for electrocatalytic dehalogenation of chloroacetic acid, dichloroacetic acid, trichloroacetic acid by (C60-3) were determined by using RDE voltammetry. Electrochemical behavior of the C60-(g-CD)2/Nafion CME in a mixed acetonitrile-water (3 : 7, v : v) solvent solution containing TBA as a cation of the supporting electrolyte was explored for electrocatalytic reduction of some 2-halogenated organic acids [207]. The C60-(g-CD)2/ Nafion CME revealed two pairs of stable cathodic and anodic CV peaks in a 2% (TBA)OH, in an acetonitrilewater (3 : 7, v : v) solution, but no electrocatalytic response in the presence of the 2-halogenated acid. In contrast, the first redox couple of this CME was stable in 0.03 M (TBA)ClO4, in the same solvent solution, and was able to dehalogenate electrocatalytically some 2-halogenated acids, indicating that C60-(g-CD)2 effectively transferred electrons. The cathodic and corresponding anodic peak was markedly increased and decreased, respectively, after addition of the chloroacetic acid to the solution. In addition, other halogenated acids, such as dichloroacetic acid, trichloroacetic acid, bromoacetic acid, 2-bromopropionic acid, and 2bromobutyric acid, were dehalogenated electrocatalytically under the same solution conditions. Moreover, 2-halogenated organic acids were electrocatalytically reduced at a (C70)2-(4-tert-butylcalix[8]arene), film coated GCE, (C70)2-4 CME [208]. This CME was prepared by evaporating solvent out of 5-mL samples of 0.1 mM (C70)24 in toluene, which were dropwise dispensed onto the GCE surface. In an acetonitrile-water (2 : 3, v ; v) solution, this CME exhibited two CV redox couples corresponding to

766 CV cathodic potential cycling was limited to the range 0 to 0.80 V (vs. SCE). However, this electroreduction became reversible after the potential window was extended positively i.e., 0.75 to 0.8 V (vs. SCE). Then, anion, C60, stabilized in the film by formation of the C60-TBA ion pair, was oxidized in the anodic CV half-cycle by the electrooxidatively generated Fc. The CME made of the C60-Fc film exhibited electrocatalytic behavior by modulating formal redox potential of the Fc/Fc redox couple. In a control experiment, no anodic peaks were observed for the C60 film without incorporated Fc. The ability of the C60-Fc CME to participate in an electrocatalytic reaction should prove useful in determination of analytes at low concentrations. This interesting property of the new C60-Fc fullerene composite material opens possibilities for developing fullerene-based electrochemical sensors. Another application of the fullerene-based CME involves catalytic electrooxidation of ascorbic acid, AA, at a reduced C60-[dimethyl-(b-CD)]2/Nafion CME [212]. This electrode was prepared by casting a film from the C60-[dimethyl-(bCD)]2 aqueous solution onto an Au electrode, then, overlaying it with a protective Nafion film. A C60 site of this CME was electroreduced irreversibly in 1 M KOH. Both neutral and reduced CMEs were examined by XPS and SEM and it appeared that the films became smoother after electroreduction of their C60 sites. This morphology change of the film surface was due to structural rearrangements accompanying release of the solvent, which was entrapped during the film casting, and ingress of potassium cations during the film electroreduction. The nature of the supporting electrolyte greatly affected electrochemical behavior of AA at the C60-[dimethyl-(b-CD)]2/Nafion CME. That is, anodic peak potential in 0.05 M KClO4 was shifted negatively by ca. 0.35 V with respect to that measured at a bare Au electrode. A linear dependence of the anodic peak current on the square root of the potential scan rate indicated that the charge transport in the film was controlled by semiinfinite diffusion. For a mixed solution of AA and dopamine, DA, the corresponding two anodic CV and DPV peaks were separated by ca. 280 mVand 305 mV, respectively, at the C60[dimethyl-(b-CD)]2/Nafion CME, as opposed to the overlapped broad anodic peaks recorded both at the bare and the Nafion film coated Au electrode. While elctrocatalysis and mediation by fullerenes have been mainly focused on alkyl halides and biomolecules, there is a report on the use C60 as a synthetic tool in chemistry. Namely, the 18-crown-6 assisted potassium metal reduction of C60 in benzene at room temperature lead to the coupling of benzene and, ultimately, formation of the biphenyl anion [213].

B. S. Sherigara et al.

3.2. Carbon Nanotubes as Catalyst Supports in Heterogeneous Catalysis and Sensor Development The discovery of the C60 fullerene has sparked a great research effort in the area of carbon related nanomaterials, such as CNTs. SWCNTs and multi-wall carbon nanotubes,
Electroanalysis 2003, 15, No. 9

MWCNTs are the new materials featuring predetermined functionality thus forming an essential bridge between industrially useful graphite fibers of a micrometer size and fullerenes of a molecular size. MWCNTs combine high electrical conductivity with mechanical strength and chemical inertness. Literature on the growth, structure and properties of CNTs has been reviewed [14]. Fullerenes and CNTs can be chemically functionalized to serve as the sites or cells for nano-chemical reactions [214]. The covalent bonding of the carbon atoms and the functionalized CNT tip gives the birth of chemical microscopy [215] which can be used to probe the local bonding, bond-to-bond interactions and chemical forces. A recent review provides status on the research and development of sensors and sensor arrays based on conjugated polymers and carbon nanotubes [15]. Miniaturization of devises into nanometer sizes has dramatically enhanced their ultimate performance. Research in nanomaterials opens many new frontiers both in fundamental science and in technology. The nano structured materials find potential applications in areas, such as electronics, optics, catalysis, ceramics, magnetic data storage and nanosensing [16]. With regards to the latter, aligned CNTs electrode arrays have been grown on silicon micro chips for sensor applications [16]. These arrays have been applied as electroanalytical nanosensors. An average electrode diameter was 200 nm and ca. 13 million CNT electrodes were mounted per square millimeter in the arrays. The individual electrode tips in the array have been insulated from each other by resin infiltration and their ends have been Au coated. In order to improve detection of low concentration analytes, functionalized surface coatings sensitive to these analytes have been applied to the Au tips. For heterogeneous catalysis, different supporting carbon materials have been used to disperse and stabilize metal particles. Catalytic properties of the resulting solids were largely dependent upon interaction between the carbon support and metal particles. For instance, a composite material, comprising the CNT supported Ru particles, is one of such heterogeneous catalysts [216]. The Ru metallic clusters were deposited onto the external surfaces of CNTs and evenly distributed. The catalytic properties of this new material were tested by cinnamaldehyde hydrogenation in a liquid phase. This process, involving parallel and consecutive reactions, appeared to be like that typical for a,bunsaturated aldehydes, as shown in Scheme 9. The product analysis by GC showed 80% cinnamaldehyde conversion and high, up to 92%, selectivity with respect to the cinnamyl alcohol product, as compared to 20 30% selectivity for the Al2O3-supported Ru catalyst and 30 40% selectivity for the active-carbon-supported Ru, both featuring similar Ru particle diameter and dispersion. Most likely, the CNT material induced metal-support interactions different than those between the Ru particles and the Al2O3 or active carbon support. A carbon nanotube powder modified microelectrodes, CNTPMEs, were prepared and applied for nitrite electro-

Applications of Fullerenes and Carbon Nanotubes

767 The ability of CNTs to facilitate the NADH electrooxidation seems prospective for development of the dehydrogenase-based amperometric biosensors. A novel CME was constructed by intercalating CNTs into the graphite electrode, a procedure which is more advantageous than the CNT film coating [219, 220]. The electrochemical behavior of DA and AA on this CME was investigated by CV (Fig. 6). The electrode intercalated with carboxylated CNTs not only showed electrocatalytic oxidation of DA and AA but also the anodic peak for DA well separated from that for AA. That is, the anodic peak potential difference for DA and AA was about 160 mV. Moreover, the sensitivity with respect to DAwas higher than that to AA. So, DA could be determined in the presence of excess of AA. The first-order derivative of the electroreduction peak current was proportional to the DA concentration. In another study, the anodic CV peak potential separation of DA and AA was studied at two kinds of CNT modified electrodes, i.e., coated and intercalated [221]. The anodic peak potential difference for DA and AA was as high as 270 mV. The charge and nature of the surfactants, used as dispersants of CNTs in coatings, strongly affected the electrochemical behavior of DA and AA. When the solution of CNTs used for electrooxidation was changed from the most commonly used mixture of HNO3 and H2SO4 (1 : 3, v : v) to dilute mixture of HNO3 and HCl, the anodic peak potential separation of DA and AA increased significantly. Carboxy groups attached to the CNTs constituted an adverse factor for the discrimination of DA from AA. The resolution of DA and AAwas mainly attributed to the stereo porous interfacial layer formed from aggregated pores and inner cavities of CNTs. Behavior of the prepared CMEs was stable and reproducible allowing to determine DA and AA simultaneously.

Scheme 9.

reduction in 0.05 M H2SO4 [217]. Properties of CNTPMEs were markedly altered by anodic pretreatment under potentiostatic conditions and pre-adsorption of the mediator. Anodic pretreatment resulted in partial cutting of long and tangled CNTs into shorter fragments and in generating their hydrophobicity sufficient to match that of the medi ator, Os(bpy) 2 3 . In effect, the CNTPME surface area and the ability to adsorb the mediator were largely increased towards efficient catalytic electroreduction of nitrite. Detection of NADH at (CNT)/GCE, at low potential, has just been reported [218]. This CME exhibited a pronounced and stable CV signal of the electrocatalytic NADH oxidation. At pH 7.4, potential of the NADH electrooxidation was decreased substantially, i.e., by 490 mV, as compared to that typical for bare GCE. The electrooxidation commenced at ca. 0.05V vs. Ag/AgCl, both for the SWNT and MWCNT coating. Furthermore, the NADH amperometric response of the (CNT)/GCE was extremely stable, with 96% and 0% of the initial activity remaining after 1 h stirring of the 2 104 and 5 103 M NADH solution, respectively, (compared to 20% and 14% for bare GCE).

Fig. 6. A) Cyclic voltammogram for 5 mM dopamine at a carbon nanotube electrode. B) Cyclic voltammogram for 1 mM dopamine at a carbon paste electrode. PBS (pH 7.4); potential scan rate 20 mV s1. (Adapted from [219]).
Electroanalysis 2003, 15, No. 9

768 CNTs, synthesized by the template carbonization of poly(pyrrole), were deposited on an alumina membrane and used as a support for the Pt-WO3, Pt-Ru, or Pt catalyst [222]. The resulting solids have been used as electrode materials for methanol oxidation in more acidic solutions than those used in case of other electrodes. The larger electrochemical surface area of these CNT modified electrode (determined by CV) has been effectively used for better dispersion of the catalyst particles. The morphology of the supported and unsupported CNTs has been characterized by SEM and HRTEM. The particle size of the Pt, Pt-Ru, or Pt-WO3 loaded CNTs was 1.2, 2, and 5 nm, respectively. The XPS spectra indicated that the Pt and Ru catalysts were in the metallic state while that of W was in its 4 oxidation state. The electrochemical activity of the methanol oxidation electrode has been evaluated by CV. The activity and stability evaluated by chronoamperometry was of the highest order. The amount of nitrogen present in the CNT support played an important role, as revealed by the activity and stability increase of the methanol oxidation with the N2 containing CNTs. Presumably, this increase was due to the hydrophilic nature of the resulting CNTs. The 4 to 50 nm in diameter CNTs, produced from highpurity graphite, were purified, activated and used as a support for deposition of a Pt catalyst [223]. After further processing, surface morphology of the resulting catalysts were characterized by TEM whereas the electrochemical activity of the CNT cathode catalysts was tested in a direct methanol fuel cell using a Pt-Ru/C anode, and a Nafion-115 membrane. Performance of the direct methanol fuel cells featuring the CNT catalyst support was better than that containing XC-72 carbon. This high electrocatalytic activity of the former was attributed to the unique structure and superior electrochemical properties of the CNTs as well as to the interaction between Pt and CNTs. The electrochemistry of horse heart cytochrome c was studied by CV at an electrode modified with a film of activated SWNTs [224]. A pair of well-defined redox peaks was obtained for a cytochrome c aqueous solution at this CME. The optimum conditions for activating the SWNT CMEs have been determined. The electrode process of cytochrome c is controlled by diffusion. The peak current increased linearly with the concentration of cytochrome c. The activated SWNT film was characterized by SEM. Furthermore, interaction of cytochrome c with adenine was characterized by electrochemical and spectral methods. A CV behavior of norepinephrine, NE, was studied at GCE modified with SWNTs [225]. In a pH 5.72 buffer solution, the SWNT CMEs show high electrocatalytic activity toward NE oxidation. One well-defined reversible redox couple was obtained at scan rates lower than 0.15 V s1. The peak current increased linearly with the NE concentration. Furthermore, the SWNT CME exhibited favorable electrocatalytic activity with respect to DA, NE and AA. A MWNT CME exhibited high stability and strong catalytic activity toward electrooxidation of nitric oxide in 0.1 M phosphate buffer, pH 7, [226]. Upon subsequent
Electroanalysis 2003, 15, No. 9

B. S. Sherigara et al.

coating with a thin film of Nafion for preventing interference from anions, especially nitrite, this CME can be employed as a fast-response high-selectivity sensor for determination of NO in solution provided that the experimental conditions, such as supporting electrolyte and amounts of Nafion as well as potential scan rate have been optimized. The chronoamperometric currents increased linearly with the NO concentration. Moreover, this determination was free from the interference of nitrite and some biological substances. The experimental results showed that NO might be adsorbed on the electrode surface prior to the transfer of electrons. The voltammetric behavior of 3,4-dihydroxyphenylacetic acid, DOPAC, was studied at GCE modified with SWNTs [227]. In 0.1 M acetic buffer, pH 4.4, this SWNT CME was electrocatalytically active toward oxidation of DOPAC. Under CV conditions, one well-defined redox couple was obtained as DOPAC underwent a two-electron oxidation to 1,2-benzoquinone, followed by dimerization. The peak current increased linearly with the increase of DOPAC concentration. The lower detection limit was 4.0 107 M DOPAC. Electrochemical responses of DOPAC and 5hydroxytryptamine could be also separated at this SWNT CME.

4. Conclusions
Because of the very rich electrochemical properties of fullerenes, their exploitation for the development of new and efficient electrocatalyts as well as chemical sensors for sensitive and selective detection of different analytes commenced a few years ago. As documented in the present review, several chemical reactions can be assisted by the electrode charge transfer to fullerenes. Due to a sequential reversible oxidative and reductive electron transfers, fullerenes act as electron mediators and, hence, operate as electron relays for activation of oxidations or reductions of target substances. Knowledge on the electrochemical behavior of fullerene derivatives in comparison with that of pristine fullerenes is important for understanding the fullerene activity and for exploring their reactivity as mediators in electrocatalytic processes as well as for exploiting their structure-reactivity aspects. Fullerene CMEs have served as excellent electrocatalysts for various chemical and biochemical reactions. Agreeably, discovery of C60 has sparked an outburst of research in the area of carbon-related nanomaterials. The surface activity of the nanomaterials is higher than that of the regular-size materials because of larger number of surface atoms. Hence, catalysis and electrocatalysis became one of the most important areas of the nanomaterial application. Moreover, CNTs have advantages as an electrode material, as they exhibit superior electrochemical properties to traditional carbon electrodes, such as the glassy carbon electrode. Hence, as documented in this review, CNT electrodes have been utilized to detect different types of analytes. Apart from the fullerene and CNT

Applications of Fullerenes and Carbon Nanotubes

769
[19] [20] [21] [22] [23] [24] [25] [26] [27] [28] S. Iijima, Nature, 1991, 354, 56. S. Iijima, T. Ichihashi, Nature, 1999, 363, 603. P. M. Ajayan, Chem. Rev. 1999, 99, 1787. Electrocatalysis (Eds: J. Lipkowski, P. N. Ross,) Wiley, New York 1998. W. Kutner, J. Wang, M. LHer, R. P. Buck, Pure Appl. Chem. 1998, 70, 1301. Q. Xie, E. Perez-Cordero, L. Echegoyen, J. Am. Chem. Soc. 1992, 114, 3978. Y. Ohsava, T. Saji, J. Chem. Soc., Chem. Commun. 1992, 781. K. Meerholz, P. Tschuncky, J. Heinze, J. Electroanal. Chem. 1993, 347, 425. F. Zhou, C. Jehoulet, A. J. Bard, J. Am. Chem. Soc. 1992, 114, 11004. R. C. Haddon, L. E. Brus, K. Raghvachari, Chem. Phys. Lett. 1986, 125, 459. Y. Yang, F. Arias, L. Echegoyen, L. P. F. Chibante, S. Flanagan, A. Robertson, L. J. Wilson, J. Am. Chem. Soc. 1995, 117, 7801. R. Taylor, J. P. Hare, A. K. Abdul-Sada, H. W. Kroto, J. Chem. Soc., Chem. Commun. 1990, 1423. G. E. Scuseria, Chem. Phys. Lett. 1991, 176, 423. Q. Xie, F. Arias, L. Echegoyen, J. Am. Chem. Soc. 1993, 115, 9818. V. Krishnan, G. Moninot, D. Dubois, W. Kutner, K. M. Kadish, J. Electroanal. Chem. 1993, 356, 93. B. Soucaze-Guillous, W. Kutner, M. T. Jones, K. M. Kadish, J. Electrochem. Soc. 1996, 143, 550. I. Noviandri, R. D. Bolskar, P. A. Lay, C. A. Reed, J. Phys. Chem. 1997, 101, 6350. Y. Ohsawa, T. Saji, J. Chem. Soc., Chem. Commun. 1992, 781. D. Dubois, K. M. Kadish, S. Flanagan, R. E. Haufler, L. P. F. Chibante, L. J. Wilson, J. Am. Chem. Soc. 1991, 113, 4363. D. M. Cox, S. Bethal, M. Disko, S. M. Gorun, M. Greaney, C. S. Hsu, E. B. Kollin, J. Millar, J. Robbins, W. Robbins, R. D. Sherwood, P. Tindall, J. Am. Chem. Soc. 1991, 113, 2940. D. R. Lawson, D. L. Feldheim, C. A. Foss, P. K. Dorhout, C. M. Elliott, C. R. Martin, B. Parkinson, J. Electrochem. Soc. 1992, 139, 68. M. M. Khaled, R. T. Carlin, P. C. Trulove, G. R. Eaton, S. S. Eaton, J. Am. Chem. Soc. 1994, 116, 3465. D. E. Cliffel, A. J. Bard, J. Phys. Chem. 1994, 98, 8140. M. E. Niyazymbetov, D. H. Evans, S. A. Lerke, P. A. Cahill, C. C. Henderson, J. Phys. Chem. 1994, 98, 13093. P. L. Boulas, M. T. Jones, K. M. Kadish, R. S. Ruoff, D. C. Lorents, R. Malhotra, D. S. Tse, Proc. Electrochem. Soc. 1994, 1007. J. P. Selegue, J. P. Shaw, T. F. Guarr, M. S. Meier, in Proc. Electrochem. Soc. 1994, 94-24, 1274. Y. Yang, F. Arias, L. Echegoyen, S. Flanagan, A. Robertson, L. J. Wilson, Proc. Electrochem. Soc. 1995, 95-10, 306. D. Dubois, K. M. Kadish, S. Flanagan, L. J. Wilson, J. Am. Chem. Soc. 1991, 113, 7773. A. F. Hebard, M. J. Rosseinsky, R. C. Haddon, D. W. Murphy, S. H. Glarum, T. T. M. Palstra, A. P. Ramirez, A. R. Koran, Nature, 1991, 350, 600. M. T. Jones, P. Boulas, Q. Yan, B. Soucaze-Guillous, W. Koh, W. Kutner, K. M. Kadish, R. Czemuszewicz, Synth. Met. 1995, 70, 1359. P. Boulas, R. Subramanian, W. Kutner, M. T. Jones, K. M. Kadish, J. Electrochem. Soc. 1993, 140, 130. W. Koh, D. Dubois, W. Kutner, M. T. Jones, K. M. Kadish, J. Phys. Chem. 1993, 97, 6871. S. Barazzouk, S. Hotchandani, P. V. Kamat, J. Mater. Chem. 2002, 12, 2021.
Electroanalysis 2003, 15, No. 9

fundamental studies including structure-topology-property relations, their future electrochemical applications will cover, most likely, the areas of biomolecular sensing by the protein- and enzyme-immobilized CNTs and in environmental analysis as miniature gas sensors. Furthermore, using CNTs chemically modified with different functional groups as electrode materials, a field in its infancy, would make the electrodes more sensitive and selective in detection application. Electrodes, chemically modified with functionalized fullerenes or CNTs have a bright prospect as a new generation of sensing devices.

5. Acknowledgements
[29]

The authors are grateful to the ACS-PRF, NIH, NSF-MariaCurie Fund II, NSF-EPSCoR, and NATO-Collaborative Linkage Grant for financial support. BSS is thankful to the Kuvempu University and University Grants Commission, India, for a travel grant.

[30] [31] [32] [33]

6. References
[1] R. E. Haufler, J. Conceicao, L. P. F. Chibante, Y. Chai, N. E. Byrne, S. Flanagan, M. M. Haley, S. C. OBrien, C. Pan, Z. Xiao, W. E. Billups, M. A. Ciufolini, R. H. Hauge, J. L. Margrave, L. J. Wilson, R. F. Curl, R. E. Smalley, J. Phys. Chem. 1990, 94, 8634. [2] P. L. Boulas, L. Echegoyen, Interface 1997, 6, 36. [3] L. Echegoyen, L. E. Echegoyen, Acc. Chem. Res. 1998, 31, 593. [4] J. Chlistunoff, D. Cliffel, A. J. Bard, in Handbook of Conductive Molecules and Polymers, Vol. 1, Ch. 7, Chargetransfer Salts, Fullerenes and Photoconductors (Ed: H. S. Nalwa), Wiley, Chichester 1997, pp. 333 412. [5] M. Prato, J. Mater. Chem. 1997, 7, 1097. [6] C. A. Reed, R. D. Bolskar, Chem. Rev. 2000, 100, 1075. [7] L. Echegoyen, F. Diederich, L. E. Echegoyen, in Fullerenes: Chemistry, Physics, and Technology (Eds: K. M. Kadish, R. S. Ruoff), Wiley, New York 2000, pp. 1 52. [8] R. S. Ruoff, K. M. Kadish, P. Boulas, E. C. M. Chen, J. Phys. Chem. 1995, 99, 8843. [9] D. V. Konarev, R. N. Lyubovskaya, Uspekhi Khimii (Russ. Chem. Rev.) 1999, 68, 23. [10] N. Martin, L. Sanchez, B. Illescas, I. Perez, Chem. Rev. 1998, 98, 2527. [11] D. Dubois, G. Moninot, W. Kutner, M. T. Jones, K. M. Kadish, J. Phys. Chem. 1992, 96, 7137. [12] R. S. Ruoff, K. M. Kadish, P. Boulas, E. C. M. Chen, J. Phys. Chem. 1995, 99, 8843. [13] M. J. Rosseinsky, Ann. Rep. Prog. Chem., Section A: Inorg. Chem. 1995, 91, 577. [14] P. M. Ajayan, T. W. Ebbesen, Rep. Prog. Phys. 1997, 60, 1025. [15] Carbon Nanotubes Special issue, Acc. Chem. Res. 2002, 35, pp. 977 1113. [16] R. O. Loutfy, M. Hecht, in Perspectives of Fullerene Nanotechnology (Ed: E. Osawa), Kluwer, Dordrecht 2002, pp. 311 316. [17] Q. Zhao, Z. Gan, Q. Zhuang, Electroanalysis 2002, 14, 1609. [18] A. Hirsch, Chemistry of the Fullerenes, Georg Thieme, Stuttgart 1994.

[34] [35] [36] [37] [38]

[39]

[40] [41] [42] [43]

[44] [45] [46] [47]

[48]

[49] [50] [51]

770
[52] Q. Xie, F. Arias, L. Echegoyen, J. Am. Chem. Soc. 1993, 115, 9818. [53] C. Boudon, J. P. Gisselbrecht, M. Gross, A. Herrmann, M. Ruttimann, J. Crassous, F. Cardullo, L. Echegoyen, F. Diederich, J. Am. Chem. Soc. 1998, 120, 7860. [54] T. Suzuki, Y. Maruyama, T. Akasaka, W. Ando, K. Kobayashi, S. Nagase, J. Am. Chem. Soc. 1994, 116, 1359. [55] D. Zhou, G. J. Ashwell, R. Rajan, L. Gan, C. Luo, H. Chunhir, J. Chem. Soc., Faraday Trans. 1997, 93, 2077. [56] F. Cardullo, P. Seiler, L. Isaacs, J.-F. Nierengarten, R. F. Haldimann, F. Diederich, T. Mordasini-Denti, N. Thiel, C. Boudon, J.-P. Gisselbrecht, M. Gross, Helv. Chim. Acta 1997, 80, 343. [57] T. Akasaka, T. Suzuki, Y. Maeda, M. Ara, T. Wakahara, K. Kobyashi, S. Nagase, M. Kako, Y. Nakadaira, M. Fujitsuka, O. Ito, J. Org. Chem. 1998, 64, 566. [58] R. Sijbesma, G. Srdanov, W. F., J. Castro, C. Wilkins, S. Fredman, G. Kenyon, J. Am. Chem. Soc. 1993, 115, 6510. [59] H. Tomioka, K. Yamamoto, J. Chem. Soc., Perkin Trans. 2 1995, 63. [60] I. Lamparth, A. Hirsch, J. Chem. Soc., Chem. Commun. 1994, 1727. [61] H. Tokuyama, S. Yamado, E. Nakamura, T. Shiraki, Y. Sugiura, J. Am. Chem. Soc. 1993, 115, 7918. [62] D. M. Guldi, H. Hungerbuhler, K.-D. Asmus, J. Phys. Chem. 1995, 99, 13487. [63] D. M. Guldi, K.-D. Asumus, Proc. Electrochem. Soc. 1997, 97-42, 82. [64] Y. P. Sun, G. Lawson, N. Wang, B. Liu, D. Moton, Proc. Electrochem. Soc. 1997, 97-42, 42. [65] T. Anderson, K. Nillson, M. Sundahl, G. Westman, O. Wennerstrom, J. Chem. Soc., Chem. Commun. 1992, 604. [66] B. Boulas, W. Kutner, M. T. Jones, K. Kadish, J. Phys. Chem. 1994, 98, 1282. [67] A. Buvari-Barcza, L. Barcza, T. Baraun, I. Konkoly-Thege, K. Ludanyi, K. Vekey, Fullerene Sci. Technol. 1997, 5, 311. [68] T. Andersson, G. Westman, O. Wennerstrom, M. Sundahl, J. Chem. Soc., Perkin. Trans. 2 1994, 1097. [69] R. M. Williams, J. W. Verhoeven, Rec. Trav. Chim. Pays-Bas 1992, 111, 531. [70] N. M. Dimitrijevic, P. V. Kamat, J. Phys. Chem. 1993, 97, 7623. [71] Z. Chen, J. M. Fox, P. A. Gale, A. J. Pilgrim, P. D. Beer, M. J. Rosseinsky, J. Electroanal. Chem. 1995, 392, 101. [72] N. Li, B. Zhou, H. Lou, W. He, Z. Shi, Z. Gu, X. Zhou, J. Solid State Electrochem. 1998, 2, 253. [73] H. Luo, N. Li, W. He, Z. Shi, Z. Gu, X. Zhou, Electroanalysis 1998, 10, 576. [74] O. V. Boltalina, D. B. Ponomarev, L. N. Sidorov, Mass Spectrom. Rev. 1997, 16, 333. [75] D. E. Cliffel, A. J. Bard, S. Shinkai, Anal. Chem. 1998, 70, 4146. [76] L. Dunsch, R. Gasiorowski, M. Pietraszkiewicz, W. Kutner, in Molecular Nanostructures. Proceedings of the International Winterschool on Electronic Properties of Novel Materials (Eds: S. H. Kuzmany, J. Fink, M. Mehring, Roth), World Scientific, Singapore, 1996, pp. 505 508. [77] G. V. Andrievsky, M. V. Kosevich, O. M. Vovk, V. S. Shelkovsky, J. Vashchenko, J. Chem. Soc., Chem. Commun. 1995, 1281. [78] A. Beeby, J. Eastoe, R. K. Heenan, J. Chem. Soc., Chem. Commun. 1994, 173. [79] T. Yagami, K. Fakuhara, S. Sueyoshi, N. Miyata, J. Chem. Soc., Chem. Commun. 1994, 517. [80] H. Hungerbuhler, D. M. Guldi, K. D. Asmus, J. Am. Chem. Soc. 1993, 115, 3386.
Electroanalysis 2003, 15, No. 9

B. S. Sherigara et al. [81] S. R. Wilson, D. I. Schuster, B. Nuber, M. S. Meier, M. Magini, M. Prato, R. Taylor, in Fullerene: Chemistry, Physics, and Technology (Eds: K. M. Kadish, R. S. Rouff), Wiley, New York 2000, pp. 91 176. [82] P. J. Fagan, B. Chase, J. C. Calabrese, D. A. Dixon, R. Harlow, P. J. Krusic, N. Matsuzawa, F. N. Tebbe, D. L. Thorn, E. Wasserman, Carbon 1992, 30, 1213. [83] A. L. Balch, M. Olmsead, Chem. Rev. 1998, 98, 2123. [84] S. A. Lerke, B. A. Parkinson, D. H. Evans, P. J. Fagan, J. Am. Chem. Soc. 1992, 114, 7807. [85] S. A. Lerke, D. H. Evans, P. J. Fagan, J. Electranal. Chem. 1995, 383, 127. [86] R. S. Koefod, C. Xu, J. R. Shapley, M. Hill, K. R. Mann, J. Phys. Chem. 1992, 96, 2928. [87] L. I. Denisovich, S. M. Peregudova, A. V. Usatov, A. L. Sigan, Y. N. Novikov, Russ. Chem. Bull. 1997, 46, 1251. [88] C. Caron, R. Subramanian, F. DSouza, J. Kim, W. Kutner, M. T. Jones, K. M. Kadish, J. Am. Chem. Soc. 1993, 115, 8505. [89] R. Subramanian, K. M. Kadish, M. N. Vijayashree, X. Gao, M. T. Jones, M. D. Miller, K. L. Krause, T. Suenobu, S. Fukuzumi, J. Phys. Chem. 1996, 100, 16327. [90] F. DSouza, C. Caron, R. Subramanian, W. Kutner, M. T. Jones, K. .M. Kadish, Proc. Electrochem. Soc. 1994, 94-24, 768. [91] K. Noworyta, P. Kuran, R. Bilewicz, E. A. Nantsis, Dunsch L., W. Kutner, Synth. Met. 2001, 123, 157. [92] C. Boudon, J.-P. Gisselbrecht, M. Gross, L. Isaacs, H. L. Anderson, R. Faust, F. Diederich, Helv. Chim. Acta 1995, 78, 1334. [93] D. H. Evans, S. A. Lerke, Proc. Electrochem. Soc. 1994, 9424, 1087. [94] P. J. Fagan, P. J. Krusic, D. H. Evans, S. A. Lerke, E. Johnston, J. Am. Chem. Soc. 1992, 114, 9697. [95] R. Taylor, G. J. Langley, J. H. Holloway, E. G. Hope, H. W. Kroto, D. M. Walton, J. Chem. Soc., Chem. Commun. 1993, 875. [96] P. R. Birket, A. G. Avent, A. D. Darwish, H. W. Kroto, R. Taylor, D. R. M. Walton, J. Chem. Soc., Chem. Commun. 1993, 1230. [97] R. Taylor, J. H. Holloway, E. G. Hope, G. J. Langley, A. G. Avent, T. J. Dennis, J. P. Hare, H. W. Kroto, D. R. M. Walton, Nature 1992, 355, 27. [98] F. Zhou, G. J. Van Berkel, B. T. Donovan, J. Am. Chem. Soc. 1994, 116, 5485. [99] N. Liu, Y. Morio, F. Okino, H. Touhara, O. V. Boltalina, V. K. Pavlovich, Synth. Met. 1997, 86, 2289 and references therein. [100] M. Fedurco, D. A. Costa, A. L. Balch, W. R. Fawcett, Angew. Chem. Int. Ed. Eng. 1995, 34, 194. [101] K. Winkler, D. A. Costa, A. L. Balch, W. R. Fawcett, J. Phys. Chem. 1995, 99, 17431. [102] A. L. Balch, D. A. Costa, W. R. Fawcett, K. Winkler, J. Phys. Chem. 1996, 100, 4823. [103] Y. Chai, T. Guo, C. Jin, R. E. Haufler, L. P. F. Chibante, J. Fure, L. Wang, J. M. Alford, R. E. Smalley, J. Phys. Chem. 1991, 95, 7564. [104] H. Shinohara, in Fullerenes: Chemistry, Physics, and Technology (Eds: K. M. Kadish, S. Ruoff), Wiley, New York 2000, pp. 357 394. [105] S. Nagase, K. Kobayashi, T. Akasaka, T. Wakahara, in Fullerenes: Chemistry, Physics, and Technology (Eds: K. M. Kadish, S. Ruoff), Wiley, New York 2000, pp. 395 436. [106] R. D. Johnson, M. S. de Vries, J. Salem, D. S. Bethune, C. S. Yannoni, Nature 1992, 355, 239. [107] D. S. Bethune, R. D. Johnson, J. R. Salem, M. S. deVries, C. S. Yannoni, Nature 1993, 366, 123.

Applications of Fullerenes and Carbon Nanotubes [108] S. Nagase, K. Kobayashi, T. Akasaka, Bull. Chem. Soc. Jpn. 1996, 69, 2131. [109] T. Akasaka, S. Okubo, T. Wakahara, K. Yamamoto, T. Kato, T. Suzuki, S. Nagase, K. Kobayashi, Proc. Electrochem. Soc. 1999, 99-12, 771. [110] K. Kikuchi, Y. Nakao, S. Suzuki, Y. Achiba, T. Suzuki, Y. Maruyama, J. Am. Chem. Soc. 1994, 116, 9367. [111] Y. Suzuki, Y. Maruyama, T. Kato, K. Kikuchi, Y. Achiba, J. Am. Chem. Soc. 1993, 115, 11006. [112] T. Suzuki, K. Kikuchi, F. Oguri, Y. Nakao, S. Suzuki, Y. Achiba, K. Yamamoto, H. Funasaka, T. Takahashi, Tetrahedron 1996, 52, 4973. [113] T. Akasaka, S. Nagase, K. Kobayashi, T. Suzuki, T. Kato, K. Yamamoto, H. Funasaka, T. Takahashi, J. Chem. Soc., Chem. Commun. 1995, 1343. [114] K. Yamamoto, H. Funasaka, T. Takahashi, T. Akaska, T. Suzuki, Y. Maruyama, J. Phys. Chem. 1994, 98, 12831. [115] M. Saunders, M. S. Syamala, J. Am. Chem. Soc. 2002, 124, 6216. [116] R. W. Murray, Introduction to the Chemistry of Molecularly Designed Electrode Surfaces, Wiley, New York 1992. [117] G. N. Salaita, A. T. Hubbard, Adsorbed Organic Molecules at Well-defined Electrode Surfaces, Wiley, New York 1992. [118] J. S. Facci, Modification of Electrode Surfaces with Selforganized Electroactive Microstuctures, Wiley, New York 1992. [119] M. Majda, Dynamics of Electron Transport in Polymeric Assemblies of Redox Centers, Wiley, New York 1992. ant, Catalysis at Redox Polymer [120] C. P. Andrieux, J.-M. Save Coated Electrodes, Wiley, New York 1992. [121] A. J. Bard, T. Mallouk, Electrodes Modified With Clays, Zeolites, and Related Microporous Solids, Wiley, New York 1992. [122] C. R. Leidner, Electron Transfer Mediation in Metal Complex Polymer Films, Wiley, New York 1992. [123] N. M. Oyama, T. Ohsaka, Voltammetric Diagnosis of Charge Transport on Polymer Coated Electrodes, Wiley, New York 1992. [124] R. A. Durst, A. J. Baumer, R. W. Murray, R. P. Buck, C. P. Andrieux, Pure Appl. Chem. 1997, 69, 1317. [125] C. A. Mirkin, B. Caldwell, Tetrahedron 1996, 52, 5113. [126] J. Chlistunoff, D. Cliffel, A. J. Bard, Thin Solid Films 1995, 257, 166. [127] R. G. Compton, R. A. Spackman, R. G. Wellington, M. L. H. Green, J. Turner, J. Electroanal. Chem. 1992, 327, 337. [128] C. Jehoulet, Y. S. Obeng, A. J. Bard, F. Wudl, K. C. Khemani, A. Koch, Abstr. Pap. Am. Chem. Soc. 1991, 202, 108. [129] C. Jehoulet, Y. S. Obeng, Y. T. Kim, F. M. Zhou, A. J. Bard, J. Am. Chem. Soc. 1992, 114, 4237. [130] R. G. Compton, R. A. Spackman, D. J. Riley, R. G. Wellington, J. C. Eklund, A. C. Fischer, M. L. H. Green, R. E. Doothwaite, A. H. H. Stephens, J. Turner, J. Electroanal. Chem. 1993, 344, 235. [131] W. Koh, D. Dubois, W. Kutner, T. M. Jones, K. M. Kadish, J. Phys. Chem. 1992, 96, 4163. [132] W. Koh, D. Dubois, W. Kutner, T. M. Jones, K. M. Kadish, J. Phys. Chem. 1993, 97, 6871. [133] M. L. H. Green, R. E. Doothwaite, A. H. H. Stephens, J. Turner, J. Electroanal. Chem. 1993, 353, 329. [134] W. B. Caldwell, K. Chen, C. A. Mirkin, S. J. Babinec, Langmuir 1993, 9, 1945. . Szcs, A. Loix, J. B. Nagy, L. Lamberts, J. Electroanal. [135] A Chem. 1995, 397, 191. . Szcs, A. Loix, J. B. Nagy, L. Lamberts, J. Electroanal. [136] A Chem. 1996, 402, 137.

771
. Szcs, A. Loix, L. Lamberts, J. B. Nagy, Synth. Met. 1996, [137] A 77, 227. [138] B. Miller, J. M. Rosamilia, G. Dabbagh, R. Tycko, R. C. Haddon, A. J. Muller, W. Wilson, D. W. Murphy, A. F. Hebard, J. Am. Chem. Soc. 1991, 113, 6291. [139] T. Atake, T. Tanaka, H. Kawaji, K. Kikuchi, K. Saito, S. Suzuki, Y. Achiba, I. Ikemoto, Chem. Phys. Lett. 1992, 196, 321. [140] J. Milliken, T. M. Keller, A. P. Baronavski, S. W. McElvany, J. H. Callahan, H. H. Nelson, Chem. Mater. 1991, 3, 386. [141] T. Tatsuma, S. Kikuyama, N. Oyama, J. Phys. Chem. 1993, 97, 12067. [142] Y. Achiba, T. Nakagava, Y. Matsui, S. Suzuki, H. Shiromaru, K. Yamauchi, K. Nishiyama, M. Kainosho, H. Hoshi, Y. Marayama, T. Mitani, Chem. Lett. 1991, 1233. [143] W. Koh, W. Kutner, M. T. Jones, K. M. Kadish, Electroanalysis 1993, 5, 209. [144] Y. M. Wang, P. V. Kamat, L. K. Patterson, J. Phys. Chem. 1993, 97, 8793. [145] A. Beeby, J. Eastoe, R. K. Heenan, J. Chem. Soc., Chem. Commun. 1994, 173. [146] Y. P. Sun, B. Ma, C. E. Bunker, B. Liu, J. Am. Chem. Soc. 1995, 117, 12705. [147] S. Nath, H. Pal, D. K. Palit, A. V. Sapre, J. P. Mittal, J. Phys. Chem. 1998, 97, 8793. [148] P. V. Kamat, S. Barazzouk, S. Hotchandani, Proc. Electrochem. Soc. 2001, 2001-14, 41. [149] P. Strasser, M. Ata, J. Phys. Chem. B 1998, 102, 4131. [150] K. Winkler, D. A. Costa, W. R. Fawcett, A. L. Balch, Adv. Mater. 1997, 9, 153. [151] A. P. Moravsky, I. O. Bashkin, I. O. Efimov, E. P. Krinichnaya, E. G. Ponyatovsky, V. V. Strelets, Izv. Acad. Nauk., Ser. Khim. 1997, 863. [152] A. L. Balch, D. A. Costa, K. Winkler, J. Am. Chem. Soc. 1998, 120, 9614. [153] A. de Bettencourt-Dias, K. Winkler, A. Hayashi, A. L. Balch, Proc. Electrochem. Soc. 1999, 99-11, 47. [154] K. Winkler, A. de Bettencourt-Dias, A. L. Balch, Chem. Mater. 1999, 11, 2265. [155] K. Winkler, K. Noworyta, W. Kutner, A. L. Balch, J. Electrochem. Soc. 2000, 147, 2597. [156] K. Winkler, A. de Bettencourt-Dias, A. L. Balch, Chem. Mater. 2000, 12, 1386. [157] H. L. Anderson, C. Boudon, F. Diederich, J.-P. Gisselbrecht, M. Gross, P. Seiler, Angew. Chem. Int. Ed. Eng. 1994, 33, 1628. [158] T. Benincori, E. Brenna, F. Sannicolo, L. Trimarco, G. Zotti, P. Sozzani, Angew. Chem. Int. Ed. Eng. 1996, 35, 648. [159] M. Nishizawa, T. Matsue, I. Uchida, J. Electroanal. Chem. 1993, 353, 329. [160] C. Jehoulet, A. J. Bard, F. Wudl, J. Am. Chem. Soc. 1991, 113, 5456. [161] P. Sharff, W. Bischof, S. Ebinel, C. Ehrdart, R. Gerken, V. Kaiser, F. Menzel, C. Tanke, Carbon, 1994, 32, 709. [162] J. D. Klein, A. Yen, R. D. Rauh, S. L. Clauson, Appl. Phys. Lett. 1993, 63, 599. [163] L. Seger, L. Q. Wen, J. B. Schlenoff, J. Electrochem. Soc. 1991, 138, L81. [164] M. Nishizawa, K. Tomura, T. Matsue, I. Uchida, J. Electroanal. Chem. 1994, 379, 233. [165] K. Chen, W. B. Caldwell, C. A. Mirkin, J. Am. Chem. Soc. 1993, 115, 1193. [166] M. S. Dresselhaus, G. Dresselhaus, P. C. Eklund, Fullerenes and Carbon Nanotubes, Academic Press, San Diego 1996. [167] N. S. Sariciftci, L. Smilowitz, A. J. Heeger, F. Wudl, Science 1992, 258, 144.
Electroanalysis 2003, 15, No. 9

772
[168] J. A. Chupa, S. T. Xu, R. F. Fischetti, R. M. Strongin, J. P. McCauley, A. B. Smith, J. K. Blasie, J. Am. Chem. Soc. 1993, 115, 4383. [169] X. Shi, W. B. Caldwell, K. Chen, C. A. Mirkin, J. Am. Chem. Soc. 1994, 116, 11598. [170] V. V. Tsukruk, L. M. Lader, W. J. Brittain, Langmuir 1994, 10, 996. [171] H. Imahori, T. Azuma, S. Ozawa, H. Yamada, K. Ushida, A. Ajavakom, H. Morieda, Y. Sakata, Chem. Commun. 1999, 557. [172] H. Imahori, H. Yamada, S. Ozawa, K. Ushida, Y. Sakata, Chem. Commun. 1999, 1165. [173] M. Fibbioli, K. Bandyopadyyay, S.-G. Liu, L. Echegoyen, O. Enger, F. Diederich, P. Buhlmann, E. Pretsch, Chem. Commun. 2000, 339. [174] S.-G. Liu, C. Martineau, J.-M. Raimundo, J. Roncali, L. Echegoyen, Chem. Commun. 2001, 913. [175] W. Kutner, K. Noworyta, R. Marczak, F. DSouza, Electrochim. Acta 2002, 47, 2371. [176] V. T. Hoang, L. M. Rogers, F. DSouza, Electrochem. Commun. 2002, 4, 50. [177] N. Nakashima, N. W. B. Wahab, M. Mori, H. Murkami, T. Sagara, Chem. Lett. 2001, 248. [178] Principles of Chemical and Biological Sensors (Ed: D. Diamond), Wiley, New York 1998. [179] Biosensors and Chemical Sensor Technology (Eds: K. R. Rogers, A. Mulchandani, W. Zhou), American Chemical Society, California 1995. [180] Y. Huang, D. M. Wayner, J. Am. Chem. Soc. 1993, 115, 367. [181] M. E. Niyazymbetov, D. H. Evans, J. Electrochem. Soc. 1995, 142, 2655. [182] T. Fuchigami, M. Kasuga, A. Konno, J. Electroanal. Chem. 1996, 411, 115. [183] F. DSouza, J.-P. Choi, Y.-Y. Hsieh, K. Shriver, W. Kutner, J. Phys. Chem. B 1998, 102, 212. [184] Z. Galus, Fundamentals of Electrochemical Analysis, 2nd ed., Ellis Horwood, New York 1994. [185] F. DSouza, J. Choi, W. Kutner, J. Phys. Chem. B 1998, 102, 4247. [186] F. DSouza, J.-P. Choi, W. Kutner, Proc. Electrochem. Soc. 1998, 98-8, 1276. [187] E. Allard, L. Riviere, J. Delaunay, D. Dubois, J. Cousseau, Tetrahedron Lett. 1999, 40, 7223. [188] F. DSouza, J. P. Choi, W. Kutner, J. Phys. Chem. B 1999, 103, 2892. ant, in Adv. Phys. Org. Chem., Vol. 26 (Ed: D. [189] J.-M. Save Bethell), Academic Press, London, 1990, pp. 1 130. [190] Z.-L. Shi, Y.-P. Mao, W. Tong, L. T. Jin, Chinese J. Chem. 1994, 12, 117. [191] H. Qian, J. Ye, L. Jin, Anal. Lett. 1997, 30, 367. [192] N.-Q. Li, M.-X. Li, H.-X. Luo, T. Liu, M. Wei, Proc. Electrochem. Soc. 2000, 2000-8, 99 and references therein. [193] C. Liu, S. Dong, G. Cheng, D. Y. Sun, J. Electrochem. Soc. 1996, 143, 3874. [194] M.-X. Li, N.-Q. Li, Z.-N. Gu, X.-H. Zhou, Y.-L. Sun, Y.-Q. Wu, Talanta 1988, 46, 993. [195] M.-X. Li, N.-Q. Li, Z.-N. Gu, X.-H. Zou, Y.-L. Sun, Y.-Q. Wu, Electroanalysis 1997, 9, 873. [196] M.-X. Li, N.-Q. Li, Z.-N. Gu, X.-H. Zhou, Y.-L. Sun, Y.-Q. Wu, Anal. Chim. Acta 1997, 356, 225. [197] R. V. Bensasson, E. Bienvenne, M. Dellinger, S. Leach, P. J. Seta, J. Phys. Chem. 1994, 98, 3492. [198] M. X. Li, M. Xu, N. Li, Z. Gu, X. Zhou, J. Phys. Chem. B 2002, 106, 4197.

B. S. Sherigara et al. [199] P. N. Bartlet, P. Tebbutt, R. G. Whitaker, Prog. Reaction Kinetics 1991, 16, 55. [200] F. Patolsky, G. Tao, E. Katz, I. Willner, J. Electroanal. Chem. 1998, 454, 9. [201] M. X. Li, N. Q. Li, Z. N. Gu, X. H. Zhou, Y. L. Sun, Y. Q. Wu, Michrochem. J. 1999, 61, 32. [202] T. Braun, A. B. Barcza, L. Barcza, I. K. Thege, M. Fodor, B. Migali, Solid State Ionics 1994, 74, 47. [203] X. Zhang, R. Zhang, J. Shen, G. Zou, Makromol. Chem. Rapid Commun. 1994, 15, 373. [204] T. Liu, M. X. Li, N. Q. Li, Z. J. Shi, Z. N. Gu, X. H. Zhou, Talanta 2000, 50, 1299. [205] T. Liu, M. Li, N. Li, Z. Shi, Z. Gu, X. Zhou, Electrochim. Acta 2000, 45, 2743. [206] T. Liu, M. Li, N. Li, Z. Gu, X. Zhou, Electrochim. Acta 2000, 45, 4457. [207] M. Li, Y. Gao, N. Li, Z. Gu, X. Zhou, Electroanalysis 2001, 13, 1253. [208] H. X. Luo, N. Q. Li, Z. J. Shi, Z. N. Gu, X. H. Zhou, Acta Chim. Sin. 2002, 60, 389. [209] H. Luo, N. Li, Z. Shi, Z. Gu, X. Zhou, Microchem. J. 2000, 65, 17. [210] C. Fang, X. Zhou, Electroanalysis 2001, 13, 949. [211] S. Barazzouk, S. Hotchandani, P. V. Kamat, J. Mater. Chem. 2002, 12, 2021. [212] M. Wei, M. Li, N. Li, Z. Gu, X. Zhou, Electroanalysis 2002, 14, 135. [213] M. P. Eastman, C. L. Wyse, J. P. Abe, R. W. Zoellner, J. Org. Chem. 1994, 59, 7128. [214] W. Q. Han, S. S. Fan, Q. Q. Li, Y. D. Hu, Science 1997, 277, 1287. [215] S. S. Wong, E. Joselevich, A. T. Woolley, C. L. Cheung, C. M. Lieber, Nature 1998, 394, 52. [216] J. M. Planeix, N. Coustel, B. Coq, V. Brotons, P. S. Kumbhar, R. Dutartre, P. Geneste, P. Bernier, P. M. Ajayan, J. Am. Chem. Soc. 1994, 116, 7935. [217] P. F. Liu, J. F. Hu, Chinese Chem. Let. 2002, 13, 79. [218] M. Musameh, J. Wang, A. Merkoci, Y. Lin, Electrochem. Commun. 2002, 4, 743. [219] P. J. Britto, K. S. V. Santhanam, P. M. Ajayan, Bioelectrochem. Bioenergetics, 1996, 41, 121. [220] Z. Wang, J. Liu, L. Yan, Y. Wang, G. Luo, Fenxi Huaxue 2002, 30, 1053. [221] Z. Wang, J. Liu, Q. Liang, Y. Wang, G. Luo, Analyst 2002, 127, 653. [222] B. Rajesh, V. Karthik, S. Karthikeyan, K. Ravindranathan Thampi, J. M. Bonard, B. Viswanathan, Fuel 2002, 81, 2177. [223] W. Li, C. Liang, J. Qiu, W. Zhou, H. Han, Z. Wei, G. Sun, Q. Xin, Carbon 2002, 40, 791. [224] J. Wang, M. Li, Z. Shi, N. Li, Z. Gu, Anal. Chem. 2002, 74, 1993. [225] J. Wang, M. Li, Z. Shi, N. Li, Z. Gu, Electroanalysis 2002, 14, 225. [226] F.-H. Wu, G.-C. Zhao, X.-W. Wei, Electrochem. Commun. 2002, 4, 690. [227] J. Wang, M. Li, Z. Shi, N. Li, Z. Gu, Electrochim. Acta 2001, 47, 651. [228] J. Cioslowski, Electronic Structure Calculations on Fullerenes and their Derivatives, Oxford University Press, Oxford, UK 1995, p. 284. [229] S. Nagase, K. Kobayashi, Chem. Phys. Lett. 1994, 231, 319. [230] Y. Yang, F. Arias, L. Echegoyen, L. P. F. Chibante, S. Flanagan, A. Robertson, L. J. Wilson, J. Am. Chem. Soc. 1995, 117, 7801.

Electroanalysis 2003, 15, No. 9

You might also like