You are on page 1of 13

REVIEW

Covalent Surface Chemistry of Single-Walled Carbon Nanotubes**


By Sarbajit Banerjee, Tirandai Hemraj-Benny, and Stanislaus S. Wong*
Dedicated to the memory of Yi Sun, a Stony Brook graduate student.

In this review article, we explore covalent chemical strategies for the functionalization of carbon-nanotube surfaces. In recent years, nanotubes have been treated as chemical reagents (be it inorganic or organic) in their own right. Indeed, from their inherent structure, one can view nanotubes as sterically bulky, p-conjugated ligands, or conversely as electron-deficient alkenes. Hence, herein we seek to understand, from a structural perspective, the breadth and types of reactions single-walled nanotubes (SWNTs) can undergo in solution phase, not only at the ends and defect sites but also along the sidewalls. Controllable chemical functionalization suggests that the unique electronic and mechanical properties of SWNTs can be tailored in a determinable manner. Moreover, prevailing themes in nanotube functionalization have been involved with dissolution of tubes.

1. Introduction
The remarkable structure-dependent properties of singlewalled carbon nanotubes (SWNTs) have attracted a lot of attention over the last decade due to their potential in applications as varied as molecular electronics, sensing, gas storage, field-emission devices, catalyst supports, probes for scanning probe microscopy, and components in high-performance com-

[*] Prof. S. S. Wong, Dr. S. Banerjee,[+] T. Hemraj-Benny Department of Chemistry, State University of New York at Stony Brook Stony Brook, NY 11794-3400 (USA) E-mail: sswong@notes.cc.sunysb.edu Prof. S. S. Wong Materials and Chemical Sciences Department Brookhaven National Laboratory Building 480, Upton, NY 11973 (USA) E-mail: sswong@bnl.gov [+] Current address: Department of Applied Physics and Applied Mathematics, Columbia University, New York, NY 10027, USA. [**] We acknowledge support of this work through startup funds provided by the State University of New York at Stony Brook as well as Brookhaven National Laboratory. Acknowledgment is also made to the National Science Foundation for a CAREER award (DMR0348239) and to the donors of the Petroleum Research Fund, administered by the American Chemical Society, for support of this research. SSW thanks 3M for a non-tenured faculty award.

posites.[1] Of an intrinsic structural beauty, carbon nanotubes consist of shells of sp2-hybridized (trivalent) carbon atoms forming a hexagonal network that is itself arranged helically within a tubular motif. The diameter and helicity of a defectfree SWNT can be uniquely characterized by the roll-up vector, also called the Hamada vector, which connects crystallographically equivalent sites on a two-dimensional (2D) graphene sheet. Electronic band structure calculations predict that the Hamada vector will determine whether a SWNT will be a metal or a semiconductor.[2] In addition, these tubes possess extremely desirable mechanical properties of strength and flexibility. Many applications utilizing SWNTs require chemical modification of the carbon nanotubes to make them more amenable to rational and predictable manipulation. For example, the generation of high-strength fibers is associated with the individualization of nanotubes and their subsequent dispersion into a polymer matrix. Moreover, the requirements of loadtransfer efficiency demand that nanotube surfaces should be compatible with the host matrix.[3] Secondly, sensor applications involve the tethering onto nanotube surfaces of chemical moieties with specific recognition sites for analytes with ensuing triggering of a predictable response in the optical or transport properties of the nanotubes.[4] Thirdly, gas storage and lithium intercalation applications necessitate the opening of hollow cavities in nanotube sidewalls. To fulfill all of these
 2004 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2005, 17, No. 1, January 6

DOI: 10.1002/adma.200401340

17

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

varied stipulations at the nanoscale requires an intimate understanding of the chemistry and functionality of carbon nanotubes. The main problem with the majority of popular synthetic methods for growing nanotubes (i.e., laser ablation, arc discharge, and chemical vapor deposition) is that they produce samples yielding a mixture of many different diameters and chiralities of nanotubes that are moreover contaminated with metallic and amorphous impurities. Thus, post-synthesis chemical processing protocols[5] that purify tubes and that can also separate individual tubes according to diameter and chirality by taking advantage of their differential reactivities are often the only viable routes to rational and predictable manipulation of the favorable electronic and mechanical properties of these materials.[6] Since important photophysical analyses of nanotubes take place in solution, the development of chemical strategies aimed at solubilizing SWNTs has been an important motivator in driving the surface chemistry of SWNTs.[6a,7] Nanotube solubilization, for instance, a) can facilitate the placement of a huge array of functional groups onto SWNT surfaces, b) is essential for the directed assembly of tubes into functional molecular devices and composites, and c) is critical for generating reproducible nanoscale chemistry (including purification) protocols.

mers such as poly(vinyl pyrrolidone) and poly(styrene sulfonate). In this case, the polymers rigidly and uniformly associate with the sides of nanotube, disrupting the hydrophobic interface with water and the inter-tube interactions within the tubular aggregates.[9a] Wrapping of SWNTs with DNA oligonucleotide sequences has been one methodology aimed at improved separation of metallic versus semiconducting nanotubes.[9b] In addition, wide ranges of biological molecules have been immobilized onto SWNT sidewalls through p-stacking interactions involving the mediation of 1-pyrenebutanoic acid and succinimidyl ester linkers. Such an approach tends to retain biological activity and functionality.[10a] Applications of these functionalized complexes include the development of highly specific biomolecule sensors based on electronic detection.[10b]

1.2. Theme of Review In this review, due to space limitations we focus explicitly on covalent surface chemical modification strategies of SWNTs. Different applications of nanotubes require varied, tailored modifications to improve nanotube processability, accessibility, and performance. Thus, covalent chemical functionalization of nanotubes has the remarkable potential to provide a higher degree of tunability of SWNT properties compared with methods based on non-covalent methods. We effectively divide this review into two sections, based on the spatial target site for chemical modification involved. Hence, we focus on selected a) end and defect site functionalization as well as b) sidewall-functionalization strategies with the broadest potential for understanding the basic science and applications of nanomaterials. We highlight differences between nanotube and fullerene reactivity where appropriate.

1.1. Non-Covalent Interactions Although non-covalent interactions with the surfaces of carbon nanotubes, as well as the filling[8] of the interior of carbon nanotubes with novel individual crystals (such as fullerene peapods), have been extensively researched, we will not focus on these types of processes in this review. Nevertheless, one of the main advantages of non-covalent functionalization, as opposed to defect-site and covalent-sidewall functionalization, has been the conservation of the SWNT electronic structure by preventing disruption of the intrinsic nanotube sp2 structure and conjugation. As examples of non-covalent functionalization, SWNTs can be solubilized in water by wrapping them with linear poly-

2. End and Defect-Site Chemistry


The end caps of nanotubes (when not closed by the catalyst particle) tend to be composed of highly curved (and hence,

Stanislaus S. Wong earned a B.Sc. (First Class Honours) from McGill University in 1994 and subsequently completed his Ph.D. thesis from Harvard University in 1999 under the tutelage of Professor Charles M. Lieber. After finishing a postdoctoral fellowship at Columbia University with Professor Louis E. Brus, he accepted a position as Assistant Professor in Chemistry at the State University of New York at Stony Brook with a joint appointment at Brookhaven National Laboratory in September 2000. He and his group have wide-ranging interests in nanotechnology, including the rational chemical functionalization of carbon nanotubes, the synthesis and characterization of metal oxide nanostructures (such as perovskites), the development of synchrotron-based techniques for nanoscale characterization, and the use of probe microscopy to initiate localized chemistry. Professor Wong has been a 3M non-tenured faculty fellow since 2002 and earned a National Science Foundation CAREER award, commencing in 2004. 18 http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2005, 17, No. 1, January 6

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

unstable) fullerene-like hemispheres, which are therefore highly reactive, as compared with the sidewalls.[6a,11] The sidewalls themselves contain defect sites (Fig. 1) such as pentagonheptagon pairs called StoneWales defects, sp3-hybridized defects, and vacancies in the nanotube lattice.[11a]

Generation of a zwitterionic linkage and amide bond formation with the oxygenated functional groups have been two of the most popular routes towards the attachment of chemical moieties onto the ends and defects of SWNTs. In particular, the main advantage of solubilizing SWNTs by these means is that the electronic structure of the SWNTs is essentially preserved. Figure 2 summarizes typical schemes for this set of reactions.

2.1. Molecular Moieties SWNTs were initially solubilized in organic solutions, such as chloroform, dichloromethane, aromatic solvents (including benzene, toluene, and chlorobenzene), and CS2 by activating acid-terminated, purified nanotubes with thionyl chloride to generate acyl chloride groups that could then be reacted with amines such as octadecylamine (ODA).[12] This work was later extended to amide formation from alkyl-aryl amines such as 4-dodecylaniline.[14] In addition, SWNTs functionalized with esters at defect sites (i.e., SWNTCOO(CH2)17CH3) have been prepared, leading to improved solubilization in solvents such as tetrahydrofuran (THF) and CHCl3.[15] SWNTs have also been functionalized with derivatized pyrenes via a similar reaction, namely the esterification of nanotube-bound carboxylic acids, to create adducts soluble in chloroform, THF, and toluene.[16] Furthermore, the attachment of derivatized lipophilic and hydrophilic dendrons via amidation and esterification reactions can render SWNTs soluble in hexane and chloroform to form colored homogeneous solutions;[17a] a similar reaction arranges SWNTs into aesthetically pleasing starshaped assemblies (Fig. 3).[17b] One example illustrating the utility of these functionalization protocols is the use of ODA-functionalized SWNT solutions as starting materials for further high-pressure liquid chromatography (HPLC) purification on a Styragel HMW7 column.[18] Zwitterion-functionalized SWNTs have been separated according to length using gel permeation chromatography.[19] Aqueous solubilization (in the range of 0.10.3 mg mL1) was achieved by functionalizing SWNTs with glucosamine and 2-aminoethanesulfonic acid.[20] Our group has demonstrated the same feat by linking 2-aminoethyl-18-crown-6 to oxygenated SWNT functional groups through a zwitterionic COONH3+ linkage, with the associated nanotube solubility values found to be on the order of ~ 1 g L1 in water as well as in methanol according to optical measurements.[21] In addition, the nanotubecrown ether adduct could be readily redissolved in ten different organic solvents at substantially high concentrations.

Figure 1. Typical defects in a SWNT. A) Five-or seven-membered rings in the carbon framework, instead of the normal six-membered ring, leads to a bend in the tube. B) sp3-hybridized defects (R = H and OH). C) Carbon framework damaged by oxidative conditions, which leaves a hole lined with COOH groups. D) Open end of the SWNT, terminated with COOH groups. Besides carboxy termini, the existence of which has been unambiguously demonstrated, other terminal groups such as NO2, OH, H, and O are possible. Reprinted with permission from A. Hirsch, Angew. Chem. Int. Ed. 2002, 41, 1853. Copyright 2002 Wiley-VCH.

Frequently, these intrinsic defects are supplemented by oxidative damage to the nanotube framework by strong acids which leave holes functionalized with oxygenated functional groups such as carboxylic acid, ketone, alcohol, and ester groups.[12] In particular, treatment of SWNTs with strong acids such as nitric acid,[13a] or with other strong oxidizing agents, such as KMnO4/H2SO4,[13b] oxygen gas,[13c] K2Cr2O7/H2SO4, or OsO4, tends to open these tubes and to subsequently generate oxygenated functional groups that serve to tether many different types of chemical moieties onto the ends and defect sites of these tubes. We consequently discuss a) attachment of molecular moieties (mainly small molecules), b) linkage of particles, c) bio-inspired moieties, and d) coordination chemistry at these localized sites.

2.2. Particles and Quantum Dots A novel strategy of altering the optical and electronic properties of nanotubes is to chemically functionalize them with a structure whose intrinsic properties are size-dependent and

Adv. Mater. 2005, 17, No. 1, January 6

http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

19

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

Figure 2. Schematic of common functionalization routes used to derivatize SWNTs at ends and defect sites. Pristine SWNTs are treated with acid or other oxidative protocols to generate oxygenated groups, such as carboxylic acid moieties, at the ends and defect sites. These functionalities are then linked to amines, either initially through reaction with SOCl2 followed by a functionalized amine, or through EDC-mediated amidebond formation. When bifunctional amines are used, pendant functional groups can attach to other moieties such as metal colloids. An alternative route to chemical functionalization involves zwitterionic interactions between complementary carboxylic acid and amine groups. Functionalization through ester linkages is also possible. Oxygenated functional groups at the ends and sidewalls are also effective sites for coordination to metal complexes and ions.

hence, configurable. Inorganic nanoparticles have been attached to SWNT surfaces through their ends and defect sites. For example, acid-oxidized nanotubes, linked to aminoethanethiol using 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide hydrochloride (EDC), possessed dangling thiol ends that were subsequently used to bind Au and Ag nanoparticles.[22] Our group has also attached CdSe quantum dots capped by mercaptobenzoic acid or mercaptopropionic acid to functional groups at the ends and defect sites of SWNTs using this linker.[23] Functionalized TiO2 nanoparticles have been similarly attached. Other groups have reported variations of this method to generate nanotubenanocrystal heterostructure assemblies with potential applications in the construction of photovoltaic cells and optoelectronic devices.[24] 2.3. Bio-Inspired Moieties A large amount of effort has been directed towards making carbon nanotubes biologically relevant. For instance, the biocompatibility of water-soluble, poly(ethylene glycol)-bound SWNTs was enhanced by forming amide bonds with bovine serum albumin (BSA) protein. Atomic force microscopy (AFM) height images showed the protein to be strongly associated with the SWNTs.[25] In addition, surface carboxylate groups on SWNT ends and defect sites have also been tagged with fluorescein molecules in the presence of EDC. These
Adv. Mater. 2005, 17, No. 1, January 6

Figure 3. a) AFM height image (5 lm 5 lm) of carbon nanotube stars on mica. The reaction mixture was directly cast on mica without any purification. b) Scanning electron microscopy (SEM) image (5 lm 5 lm) of the same sample as in (a) but coated with Pt/Pd. Reprinted with permission from M. Sano, A. Kamino, S. Shinkai, Angew. Chem. Int. Ed. 2001, 40, 4661. Copyright 2001 Wiley-VCH.

20

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

fluorescently labeled nanotubes were subsequently taken up by human promyelocytic leukemia cells and T cells.[26] Amine-functionalized biotin moieties were immobilized onto nanotube functional groups and further used to link to fluorescein-tagged streptavidin. These functionalized SWNTs were found to be capable of transporting their protein loads into cells, with the internalized conjugates eliciting a functional dose-dependent response.[26] The low toxicity of the nanostructures makes them potentially useful as biocompatible transporters. Peptide nucleic acid (PNA) strands have been linked to the carboxyl end groups of SWNTs through the displacement of N-hydroxysuccinimide in the presence of EDC, as suggested by the AFM height images in Figure 4. The PNA could be hybridized with complementary DNA sequences, suggesting the possibility for recognition-based assembly of functionalized nanotubes, as well as for the potential use of these DNAfunctionalized SWNT adducts as biological sensors.[27]

2.4. Coordination Chemistry Some of the first compounds added onto fullerenes included coordination complexes, such as Ir, Pt, and Pd compounds. In each case, the metal coordinated in an g2 fashion to the carboncarbon double bond.[28] Elimination of pyramidalization of carbon atoms in fullerenes, as well as the

relief of local and global strain on g2 coordination, have been the driving forces for metal complexation reactions of fullerenes.[28a] The reactivity of SWNTs with respect to metal coordination compounds has not been as straightforward, though we have found that coordination at oxidized functional groups appears to be a favorable mode of bonding. In fact, the development of this chemistry has been crucial for SWNT applications, including their usage as reusable catalyst supports, inspired primarily by decades of carbon fiber and activated carbon research,[29] as well as their potential in varying the solubilization and bundling characteristics of nanotubes. Changes in stacking properties of functionalized tubes may be due to changes in hydrogen bonding. That is, intervening chemical moieties may intercalate between bundles of nanotubes. We have found, for example, that coordination of lanthanide ions to the oxygenated functional groups on the SWNT surfaces disrupted stacking interactions by preventing hydrogen bonding amongst the various individual bundles of tubes.[30] We have also been able to coordinate Vaska's compound [Ir(CO)Cl(PPh3)2] as well as Wilkinson's compound [RhCl(PPh3)3] (Fig. 5)[31a] to the ends and sidewalls of SWNTs.[31b] These complexes showed good solubility in organic solvents, originating from exfoliation of large crystalline ropes to produce smaller bundles and individual tubes. Furthermore, the Rh-based system was found to be useful in the recyclable catalytic hydrogenation of cyclohexene.

Figure 4. Attachment of DNA to carbon nanotubes. a,b) N-Hydroxysuccinimide (NHS) esters formed on carboxylated SWNTs are displaced by peptide nucleic acid (PNA), forming an amide linkage. c) A DNA fragment with a single-stranded, sticky end hybridizes by WatsonCrick base-pairing to the PNASWNT. d,e) AFM (tapping mode) height images of PNASWNTs. SWNTs appear as bright lines; the paler strands represent bound DNA. Scale bars: 100 nm. The diameters of the nanotubes are 0.9 nm in (d) and 1.6 nm in (e). Reprinted with permission from K. A. Williams, P. T. M. Veenhuizen, B. G. de la Torre, R. Eritja, C. Dekker, Nature 2002, 420, 761. Copyright 2002 Nature Publishing Group. We are grateful to K. A. Williams (Department of Physics, University of Virginia) for permission to use this figure.

Adv. Mater. 2005, 17, No. 1, January 6

http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

21

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

Figure 5. Schematic of a functionalized SWNTRhCl(PPh3)3 adduct. The image shows a possible mode of coordination wherein oxygenated functionalities, such as carboxylic acids, at the open cap of a (5,5) SWNT are able to coordinate to the Rh-metal center. Oxygenated functionalities are expected at the ends and defect sites.

quantum dots on nanotube surfaces. Essentially, we coordinate metallic precursors (such as Cd) of quantum dots directly onto the oxygenated functionalities of the nanotubes, and then proceed with in-situ growth of nanocrystals upon further injection of additional reactants such as selenium or tellurium.[36] The variable size and shape of the quantum dots generated are hence dictated by the spatial location and distribution of the oxygenated groups on the nanotube surface. This novel strategy generates well-defined interfaces and junctions between nanostructures, demonstrating the potential for generating complex nanoscale architectures using coordination chemistry with potentially useful applications in photovoltaic cells, optoelectronic devices, and sensors.

3. Covalent Sidewall Functionalization


To explain these observations, Amsterdam density functional theory (ADFT) calculations determined that, in the fragments simulating SWNTs, the presence of two hexagonal rings, as opposed to the cyclopentadienyl units noted in fullerenes, increased the highest occupied molecular orbitallowest unoccupied molecular orbital (HOMOLUMO) gap and reduced the affinity of the system towards electron-rich metal species due to reduced back donation.[32] That is, whereas cyclopentadienyl groups in fullerenes helped to stabilize coordination complexes, their absence in SWNTs made it harder for SWNTs to form g2 complexes with metal-containing coordination compounds. Hence, this explained the preference for metal coordination to oxygenated functional groups when oxidized nanotubes were used. Metal coordination at oxygenated functional groups at ends and defect sites thus presents the possibility of tethering metal clusters and nanoparticles at these locations on SWNTs. It is not surprising, then, that well-dispersed platinum nanoparticles could be tethered to functional groups on the SWNT surface through ion exchange from a potassium chloroplatinate precursor in an ethylene glycol medium; these were then used in the selective reduction of prenal to prenol.[33] Similarly, Hg clusters can be deposited on carboxylic acid and alcohol groups on nanotube surfaces by reduction from a Hg(NO3)2 solution.[34] Well-dispersed, homogeneous catalytic Rh clusters (1.5 2.5 nm), attached through the mediation of carboxylate groups on a multiwalled carbon nanotube (MWNT) surface,[35] showed better performance compared with activatedcarbon-supported catalysts in the catalytic hydrogenation of trans-cinnamaldehyde and the catalytic hydroformylation of hex-1-ene in the liquid phase. The grafting methodology used in the creation of this adduct clearly established the need for functional groups on the nanotube surface to obtain well-dispersed catalytic clusters.[29] Recently, our group has used the oxidized defect sites on MWNTs, as well as on heavily oxygenated SWNTs, to demonstrate the in-situ mineralization of crystalline CdTe and CdSe Whereas fullerenes have a well-developed addition chemistry,[28c] sidewall functionalization of carbon nanotubes has only been achieved relatively recently.[11c] The binding energy for covalent addition to SWNT sidewalls was calculated to be much lower than that for fullerenes in the range of diameters of commercially available SWNTs.[37] In effect, it is predicted that the sidewall addition chemistry of SWNTs differs from that of fullerenes, even though both are curved, conjugated carbon systems. While fullerenes can be thought of as being folded in two dimensions (spherical folding), nanotubes are formed from a graphene sheet, folded in one dimension (cyclic folding).[6a,37] The chemical reactivity in strained carbon systems arises from two factors: a) pyramidalization at the carbon atom, and b) p-orbital misalignment between adjacent carbon atoms.[38] In fullerenes, the former is more important, and relief of pyramidalization strain energy results in addition reactions to fullerenes being energetically favorable. In SWNTs, the pyramidalization strain is not as acute. Hence, it turns out that in SWNTs, p-orbital misalignment is expected to have a greater influence.[6a,39] This misalignment, associated with bonds at an angle to the tube circumference (i.e., bonds that are neither parallel nor perpendicular to the tube axis) is the origin of torsional stain in nanotubes, and the relief of this strain controls the extent to which addition reactions occur with nanotubes. Since p-orbital misalignment, as well as pyramidalization, scale inversely with tube diameter,[37] smaller diameter tubes are expected to be more reactive than larger diameter tubes. Pyramidalization and p-orbital-misalignment angles are illustrated for a (5,5) tube in Figure 6. Not surprisingly, most reported sidewall-functionalization reactions of SWNTs need very reactive addends.[11c,37] Moreover, calculations indicate that the concave outer surfaces remain more amenable to covalent addition as compared with the inner convex surfaces, though the differences in reactivity are less pronounced than in fullerenes. Figure 7 shows reaction schemes for several sidewall addition reactions of SWNTs. In particular, we will focus our discussion on a) fluor-

22

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2005, 17, No. 1, January 6

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW
tube sidewall functionalization.[11c] What is important to note is that sidewall functionalization causes a disruption of the pseudo-one-dimensional lattice of SWNTs. This loss of conjugation and disruption of band structure is also reflected in the loss of optical transitions between van Hove singularities in the electronic density of states (DOS) of nanotubes, a phenomenon which can also be observed in the UV-vis/near-IR data of these systems. Far-IR investigations have revealed that upon covalent bond formation at the nanotube sidewalls, the metallic properties of metallic SWNTs are greatly affected, with significant loss of oscillator strength in the far-IR band corresponding to transitions at the Fermi level.[43] In effect, covalent bonding opens up a gap at the Fermi level, changing a metallic tube to a semiconductor.[43b]

Figure 6. Diagrams of a) metallic (5,5) SWNT, b) pyramidalization angle (hP), and c) the p-orbital misalignment angles (u) along the C1C4 bond in the (5,5) SWNT and its capping fullerene, C60. Reprinted with permission from S. Niyogi, M. A. Hamon, H. Hu, B. Zhao, P. Bhowmik, R. Sen, M. E. Itkis, R. C. Haddon, Acc. Chem. Res. 2002, 35, 1105. Copyright 2002 American Chemical Society.

ination, b) ozonolysis, c) organic functionalization, d) osmylation, and e) derivatization with azomethine ylides. The high degree of chemical functionalization possible through covalent sidewall derivatization makes these routes ideal for applications such as composite formation. However, one disadvantage of these reactions is that they usually result in the loss of the intrinsic electronic structure. It has been predicted that covalent chemical attachments can decrease the maximum buckling force of SWNTs by as much as 15 %, regardless of tube helical structure or radius.[40] A threefold decrease in thermal conductivity, due to a decrease in the phonon-scattering length, was calculated to occur upon 1 % sidewall functionalization of SWNTs with phenyl groups.[41] A number of spectroscopic, microscopic, and gravimetric techniques have been used to monitor and characterize sidewall functionalization of SWNTs. A detailed discussion of these techniques is beyond the scope of this review. Notably, Raman spectroscopy is particularly effective at denoting the presence and extent of functionalization, as well as for revealing electronic structure and electronphonon coupling in SWNTs.[42] Importantly, in the context of this review, the shape and intensity of a weak disorder mode peak at 1290 1320 cm1, attributable to the destruction of the sp2-hybridized structure, has been correlated with the extent of nano-

3.1. Fluorination Raman studies of SWNTs fluorinated at temperatures ranging from 150325 C suggested sidewall derivatization.[44] These fluorinated tubes could be solubilized in various alcohol solvents by means of ultrasonication. Scanning tunneling microscopy (STM) images of fluorinated tubes revealed a dramatic banded structure, indicating broad continuous regions

Adv. Mater. 2005, 17, No. 1, January 6

http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

23

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW
Figure 7. Schematic describing various covalent sidewall functionalization reactions of SWNTs. Descriptions of many of these reactions are discussed in the text.

24

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2005, 17, No. 1, January 6

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

of fluorination terminating abruptly in bands orthogonal to the tube axis.[45] These results suggested that fluorine adds more favorably around the circumference of the tube, backed up by semiempirical AM1 (Austin Model 1) calculations.[45] Further calculations showed that the F atoms tended to bond next to each other.[46] Fluorinated SWNTs have been used as cathode materials in a lithium electrochemical cell, and were found to produce a higher cell potential than commercial fluorographite.[47] Because CF bonds in fluorinated nanotubes are weakened due to an eclipsing strain effect,[44] it is no surprise that sidewall-fluorinated nanotubes could be further modified by reacting them with alkyl magnesium bromides in a Grignard synthesis or by reaction with alkyllithium precursors to yield sidewall-alkylated nanotubes via a concerted, allylic displacement mechanism.[48] Oxidizing these alkylated nanotubes in air allows for recovery of pristine nanotubes. As for fluorinated tubes, it was found that hydrazine could be used to regenerate the unfluorinated starting material. A procedure was also established to tune the fluorine content of the tubes by first fluorinating heterogeneously, solvating in alcohol, and then de-fluorinating with substoichiometric quantities of hydrazine to generate tubes with different fluorine contents.[49] Terminal diamines (H2N(CH2)nNH2, where n is 2, 3, 4, or 6) can displace fluorine to form sidewall aminoalkyl-functionalized nanotubes. These amine-terminated groups could be further reacted with adipoyl chloride to form new nylontube types of materials.[50] More recently, fluorinated SWNTs have been reacted with diols and glycerol in the presence of alkali metals or alkali hydroxides to give OH-terminated moieties on nanotube sidewalls.[51]

cation of as-prepared SWNTs to obtain a high-quality product; second, chemical functionalization of nanotube sidewalls, demonstrated by Raman and near-edge X-ray absorption fine structure (NEXAFS) data;[55a] and third, development of a systematic procedure to select for particular distributions of oxygenated functional groups in the resultant purified SWNTs. The last goal is accomplished by generating higher proportions of carboxylic acids, aldehydes/ketones, or alcohols on the surfaces of carbon nanotubes through rational chemical treatment of the primary ozonide intermediate with hydrogen peroxide (H2O2), dimethyl sulfide (DMS), or sodium borohydride (NaBH4), respectively. Thus, the chemistry described at the ends and defect sites of SWNTs, as discussed in Section 2, can be generalized for SWNT sidewalls, and hence a higher degree of functionalization is achievable.[54,55] It is not surprising then that heavily oxygenated ozonized SWNTs have been used as ligands for the in-situ growth of CdSe and CdTe quantum dots, leading to the formation of nanotubenanocrystal assemblies, as shown in Figure 9.[36b,c] Ozonized SWNTs have also been assembled onto selfassembled monolayers, terminating in conjugated oligo(phenylene ethynylene)s.[55b] Our group has recently found, on the basis of resonance Raman data, that the reaction itself is diameter selective; that is, smaller diameter SWNTs react more extensively than their less strained but larger counterparts, yielding the potential for developing methods of diameterselective separation of SWNTs.[56]

3.3. Organic Functionalization SWNTs have been functionalized in situ by a thermally induced reaction with diazonium compounds, generated by the action of isoamyl nitrite on aniline derivatives.[57a] While the reaction was first done electrochemically, it was later shown that the diazotization functionalization could also be achieved at high temperatures by refluxing the reaction mixture.[57b] This procedure allows for a very high degree of functionalization (estimated to be 812 % from thermal gravimetric analysis) to be achieved. A solvent-free analogue of this system has been reported where the reaction proceeds by mechanical stirring to yield highly functionalized and soluble nanotubes.[7a,58] Recent observations have implied that the diazonium salt sidewall functionalization does not follow a simple diameter-

3.2. Ozonolysis A recent ONIOM (Our N-layered Integrated Molecular Orbital plus Molecular Mechanics) two-level calculation predicted that ozone would add to the double bonds of SWNTs[52] through a facile 1,3-dipolar cycloaddition reaction following Criegee's mechanism, analogous to the addition of alkenes and fullerenes.[53] We recently reported an ozonolysis protocol (Fig. 8) involving treatment of a methanolic dispersion of nanotubes at 78 C with ozone.[6c,54] This one-pot oxidative methodology has three major features: first, purifi-

b
Figure 8. a) SEM image of ozonized tubes. Scale bar is 1 lm. b) Proposed reaction scheme for ozonation. The primary ozonide that is initially formed is cleaved to yield oxygenated functional groups on the sidewalls. Reprinted with permission from S. Banerjee, S. S. Wong. J. Phys. Chem. B. 2002, 106, 12 144. Copyright 2002 American Chemical Society.

O3

cleavage

Adv. Mater. 2005, 17, No. 1, January 6

http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

25

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW
Figure 9. CdTeSWNT heterostructures grown on ozonized tubes. a) AFM height image and schematic showing coordination of oxygenated functional groups on nanotube surface to CdTe quantum dots. b) SEM image of CdTeSWNT heterostructures indicating extent of coverage. c) AFM height image showing nanocrystal-mediated junctions between nanotubes. d) AFM height image of various nanotubenanocrystal assemblies. e) High-resolution transmission electron microscopy (HRTEM) image of wurtzite CdTe nanocrystals grown along the sidewall of a bundle of ozonized SWNTs. Reprinted with permission from S. Banerjee, S. S. Wong. Adv. Mater. 2004, 16, 34. Copyright 2004 WileyVCH.

dependence model. Rather, the electronic structure of this moiety likely has a greater influence over reactivity.[5a,59] The extent of electron transfer is dependent on the density of states in that electron density near the Fermi energy level (EF) leads to higher initial activities for metallic and semimetallic nanotubes. That is, diazotization is an example of a functionalization reaction where the transition state (at an energy level ET) is favored by electron donation from the nanotube (Fig. 10). Hence, due to their finite electron density at the Fermi level, metallic SWNTs are better able to stabilize the charge-transfer complex that characterizes the transition state through electron donation, and are consequentially more reactive in these types of reactions. Once a covalent bond has formed, defects are created, the nanotube symmetry is broken. Nearby carbon atoms increase in reactivity, and ultimately, the electronically selective reaction proceeds. SWNTs that had been sidewall functionalized with 4(10-hydroxy)decyl benzoate moieties using this diazotization

Figure 10. IA) Diazonium reagents extract electrons, thereby evolving N2 gas and leaving a stable CC covalent bond with the nanotube surface. IB) The extent of electron transfer is dependent on the density of states in that electron density near EF leads to higher initial activity for metallic and semimetallic nanotubes. IC) The arene-functionalized nanotube may now exist as the delocalized radical cation, which could further receive electrons from neighboring nanotubes or react with fluoride or diazonium salts. Reprinted with permission from Strano et al., Science 2003, 301, 1519. Copyright 2003 AAAS. II) Schematic of the electronic DOS of metallic and semiconducting nanotubes overlaid on a SEM image of osmylated tubes. Reprinted with permission from S. Banerjee, S. S. Wong. J. Am. Chem. Soc. 2004, 126, 2073. Copyright 2004 American Chemical Society. The inset in (II) shows the optimized geometry of an osmate ester functionalizing the nanotube surface. Reprinted with permission from Lu et al., Nano Lett. 2002, 2, 1325. Copyright 2002 American Chemical Society.

method were used to make composites with polystyrene. The composites showed clear evidence for the formation of a percolated filler network at SWNT loadings as low as 1.5 wt.-%. Rheology data suggested that the degree of reinforcement and dispersability were significantly improved over composites created using pristine SWNTs, which did not form a network superstructure even at loadings of 3 wt.-%.[60] Direct addition to the unsaturated p-electron systems of SWNTs can also occur through i) a [2+1] cycloaddition of nitrenes, ii) a [4+2] DielsAlder cycloaddition of o-quinodimethane performed under microwave irradiation, iii) a cyclopropanation accomplished under Bingel reaction conditions that can be visualized using STM, iv) the addition of nu-

26

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2005, 17, No. 1, January 6

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

cleophilic carbenes, and v) the addition of radicals (through the photoinduced addition of perfluorinated alkyl radicals).[11a,61,62] Nitrile ylides and nitrile imines have been theoretically predicted to be the best candidates to undergo 1,3-dipolar cycloaddition onto SWNT sidewalls.[62d] Many of these reactions are useful in rational nanotube manipulation. Specifically, for the addition of nitrenes[61] in the presence of alkyl azidoformates, alkoxycarbonylaziridinoSWNTs can be formed which are soluble in dimethyl sulfoxide (DMSO). Moreover, the Bingel reaction[62a] allows for the introduction of chemical tags such as Au colloids onto SWNT surfaces, thereby enabling a means of spatially locating functional groups and determining their distribution. In addition, the particularly high solubility of carbene adducts[11a] can be explained through the presence of mutual electrostatic repulsion of tubes. In terms of other relevant sidewall reactions, nanotubes have even been hydrogenated via a modified Birch reduction method and reductively alkylated using lithium, with both processes utilizing liquid ammonia as a reaction medium.[63] Tubes functionalized in these various ways enable a large variety of terminal functional groups to be introduced onto nanotube surfaces with implications for nanocomposite formation and biomedical applications. For instance, reaction of SWNTs with succinic or glutaric acid acyl peroxides in o-dichlorobenzene (ODCB) at 8090 C results in sidewall derivatization with pendant carboxylic groups that can be subsequently converted into a wide range of functional, active moieties.[64] In addition, a detailed investigation of the sidewall functionalization of nanotubes using a family of oxycarbonyl nitrenes substituted with alkyl chains, aromatic groups, dendrimers, crown ethers, and oligo(ethylene glycol) units has recently been reported, leading to increased adduct solubility in DMSO and ODCB. The degree of functionalization, though, was < 5 % in that set of reactions.[65]

Achieving a neat separation of metallic versus semiconducting nanotubes is critical for a variety of applications. For instance, semiconducting nanotubes are useful for optical sensing, whereas isolated metallic tubes with quasi-ballistic transport could be useful as leads in these nanoscale devices.[1c] Hence, covalent sidewall functionalization, by exploiting subtle differences in reactivity between different species of SWNTs, offers an important route to generating such fundamentally interesting monodisperse samples of nanotubes.

3.5. Azomethine Ylides Derivatization based on 1,3-dipolar cycloaddition of azomethine ylides, generated by the condensation of an a-amino acid and an aldehyde, has been utilized to solubilize tubes in most organic solvents.[7b,66] This particular functionalization technique has also been employed for the purification of commercial SWNTs synthesized by high-pressure CO conversion (HiPCO). This covalent functionalization of SWNTs with pyrrolidine rings allows for a wide number of other functionalities to be immobilized onto the SWNT surface, as shown in Figure 11.[67] For instance, a ferrocene-modified glycine precursor can be used to functionalize SWNT sidewalls with electron donor ferrocene units. In this case, electron transfer from ferrocene to the SWNTs was noted on photoexcitation, paving the way for the use of these SWNT hybrids in solar energy harvesting applications.[68] Biological moieties can also been introduced onto nanotube surfaces using this chemistry. For instance, N-protected amino acids were linked to SWNT sidewalls and used to attach bioactive peptides through fragment condensation or by selective chemical ligation using a maleimido linker.[69] In particular, a nineteen amino acid epitope from the VP1 envelope protein of the foot-and-mouth disease virus (FMDV) was linked using the latter protocol to generate covalently bound peptideSWNT conjugates that could be subsequently recognized by specific antibodies using surface plasmon resonance (SPR) and enzyme-linked immunoabsorbent assays (ELISA). In-vivo immunization of mice with these conjugates produced a higher antibody response than with the free peptide itself.[69,70] The favorable antigenicity and immunogenicity of these nanotube-based conjugates make them promising for use in diagnostics, vaccine delivery, and gene therapy.[70]

3.4. Osmylation We recently reported an analogous reaction to diazotization involving the osmylation of SWNTs under UV irradiation.[5c] In fact, the interaction of OsO4 in toluene with SWNTs in the presence of UV irradiation demonstrated chemical specificity for metallic nanotubes, with a larger electron density near the Fermi level. The net results of osmylation were a) covalent sidewall functionalization of these nanotubes through disruption of the conjugated p-electron structure, as well as b) reduction of the osmium tetroxide species to OsO2 nanoparticles, which were then templated onto the sidewall surface. A systematic Raman study of our nanotube samples at three different excitation wavelengths, which probed different electronic populations of tubes, provided strong evidence of the higher reactivity of metallic tubes with respect to osmylation, mainly because of the dramatic loss of resonances at 514.5 nm compared with the minimal alterations observed with the peaks of primarily semiconducting tubes at 1064 nm.

4. Conclusions
Rational SWNT functionalization provides for the potential of the manipulation of their properties in a predictive manner. The surface chemistry of SWNTs plays a vital role in enabling the dispersability, purification, solubilization, biocompatibility, and diameter- and chirality-based separation of these unique nanostructures. In addition, derivatization allows for a number of site-selective nanochemistry applications such as

Adv. Mater. 2005, 17, No. 1, January 6

http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

27

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

[2] M. Ouyang, J.-L. Huang, C. M. Lieber, Acc. Chem. Res. 2002, 35, 1018. [3] a) J.-P. Salvetat, J.-M. Bonard, N. H. Thomson, A. J. Kulik, L. Forro, W. Benoit, L. Zuppiroli, Appl. Phys. A: Mater. Sci. Process. 1999, 69, 255. b) P. Calvert, Nature 1999, 399, 210. [4] H. Dai, Acc. Chem. Res. 2002, 35, 1035. [5] a) M. S. Strano, C. A. Dyke, M. L. Usrey, P. W. Barone, M. J. Allen, H. Shan, C. Kittrell, R. H. Hauge, J. M. Tour, R. E. Smalley, Science 2003, 301, 1519. b) D. Chattopadhyay, I. Galeska, F. Papadimitrikopoulos, J. Am. Chem. Soc. 2003, 125, 3370. c) S. Banerjee, S. S. Wong, J. Am. Chem. Soc. 2004, 126, 2073. d) H. Li, B. Zhou, Y. Lin, L. Gu, W. Wang, K. A. S. Fernando, S. Kumar, L. F. Allard, Y.-P. Sun, J. Am. Chem. Soc. 2004, 126, 1014. [6] a) S. Niyogi, M. A. Hamon, H. Hu, B. Zhao, P. Bhowmik, R. Sen, M. E. Itkis, R. C. Haddon, Acc. Chem. Res. 2002, 35, 1105. b) S. S. Wong, S. Banerjee, in Dekker Encyclopedia of Nanoscience and Nanotechnology (Eds: J. A. Schwarz, C. I. Contescu, K. Putyera), Marcel Dekker, New York 2004, p. 1251. c) S. Banerjee, M. G. C. Kahn, S. S. Wong, Chem. Eur. J. 2003, 9, 1898. [7] a) C. A. Dyke, J. M. Tour, Chem. Eur. J. 2004, 10, 81. b) D. Tasis, N. Tagmatarchis, V. Georgakilas, M. Prato, Chem. Eur. J. 2003, 9, 4000. [8] a) R. R. Meyer, J. Sloan, R. E. Dunin-Borkowski, A. I. Kirkland, M. C. Novotny, S. R. Bailey, J. L. Hutchinson, M. L. H. Green, Science 2000, 289, 1324. b) J. Sloan, A. I. Kirkland, J. L. Hutchison, M. L. H. Green, Acc. Chem. Res. 2002, 35, 1054. c) D. J. Hornbaker, S.-J. Kahn, S. Misra, B. W. Smith, A. T. Johnson, E. J. Mele, D. E. Luzzi, A. Yazdani, Science 2002, 295, 828. [9] a) M. O'Connell, P. Boul, L. M. Ericson, C. Huffman, Y. Wang, E. Haroz, C. Kuper, J. Tour, K. D. Ausman, R. E. Smalley, Chem. Phys. Lett. 2001, 342, 265. b) M. Zheng, A. Jagota, M. S. Strano, Figure 11. 1,3-Dipolar cycloaddition of various azomethine ylides onto SWNTs. Reprinted A. P. Santos, P. Barone, S. G. Chou, B. A. Diner, with permission from N. Tagmatarchis, M. Prato, J. Mater. Chem. 2004, 14, 437. ReproM. S. Dresselhaus, R. S. Mclean, G. B. Onoa, G. G. duced with permission of The Royal Society of Chemistry. Samsonidze, E. D. Semke, M. Usrey, D. J. Walls, Science 2003, 302, 1545. [10] a) R. J. Chen, Y. Zhang, D. Wang, H. Dai, J. Am. Chem. Soc. 2001, 123, 3838. b) R. J. Chen, self-assembly of nanotubes with tunable electronic properties, S. Bangsaruntip, K. A. Drouvalakis, N. W. S. Kam, M. Shim, Y. Li, important for advances in molecular electronics. Other derivW. Kim, P. J. Utz, H. Dai, Proc. Nat. Acad. Sci. USA 2003, 100, atized SWNT adducts show potential as catalytic supports and 4984. as biological transport vessels. Moreover, these systems often [11] a) A. Hirsch, Angew. Chem. Int. Ed. 2002, 41, 1853. b) S. B. Sinnott, demonstrate novel charge-transfer characteristics, the developJ. Nanosci. Nanotechnol. 2002, 2, 113. c) J. Bahr, J. M. Tour, J. Mater. Chem. 2002, 12, 1952. ment and understanding of which have implications for photo[12] J. Chen, M. A. Hamon, H. Hu, Y. Chen, A. M. Rao, P. C. Eklund, catalysis and energy storage. Finally, chemical manipulation R. C. Haddon, Science 1998, 282, 95. of SWNTs is critical for the hierarchical building-up of these [13] a) J. Liu, A. G. Rinzler, H. Dai, J. H. Hafner, R. K. Bradley, P. J. nanomaterials into functional architectures, such as nanocomBoul, A. Lu, T. Iverson, K. Shelimov, C. B. Huffman, F. Rodriguezposites and nanocircuits, with unique structural properties. Macias, Y. S. Shon, T. R. Lee, D. T. Colbert, R. E. Smalley, Science 1998, 280, 1253. b) H. Hiura, T. W. Ebbesen, K. Tanigaki, Adv. MaReceived: August 17, 2004 ter. 1995, 7, 275. c) P. M. Ajayan, T. W. Ebbesen, T. Ichihashi, S. IijiFinal version: November 3, 2004 ma, K. Tanigaki, H. Hiura, Nature 1993, 361, 333. [14] M. A. Hamon, J. Chen, H. Hu, Y. Chen, M. E. Itkis, A. M. Rao, [1] a) Carbon Nanotubes: Synthesis, Structure, Properties, and ApplicaP. C. Eklund, R. C. Haddon, Adv. Mater. 1999, 11, 834. tions (Eds: M. S. Dresselhaus, G. Dresselhaus, P. Avouris), Springer, [15] M. A. Hamon, H. Hui, P. Bhowmik, H. M. E. Itkis, R. C. Haddon, Berlin, Germany 2001. b) R. H. Baughman, A. A. Zakhidov, W. A. Appl. Phys. A: Mater. Sci. Process. 2002, 74, 333. de Heer, Science 2002, 297, 787. c) P. Avouris, Acc. Chem. Res. 2002, [16] L. Qu, R. B. Martin, W. Huang, K. Fu, D. Zweifel, Y. Lin, Y.-P. Sun, 35, 1026. J. Chem. Phys. 2002, 117, 8089.

28

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.advmat.de

Adv. Mater. 2005, 17, No. 1, January 6

Banerjee et al./Surface Chemistry of Carbon Nanotubes

REVIEW

[17] a) Y.-P. Sun, W. Huang, Y. Lin, K. Fu A. Kitaygorodskiy, L. A. Riddle, Y. J. Yu, D. L. Carroll, Chem. Mater. 2001, 13, 2864. b) M. Sano, A. Kamino, S. Shinkai, Angew. Chem. Int. Ed. 2001, 40, 4661. [18] a) S. Niyogi, H. Hu, M. A. Hamon, P. Bhowmik, B. Zhao, S. M. Rozenhak, J. Chen, M. E. Itkis, M. S. Meier, R. C. Haddon, J. Am. Chem. Soc. 2001, 123, 733. b) B. Zhao, H. Hu, S. Niyogi, M. E. Itkis, M. A. Hamon, P. Bhowmik, M. S. Meir, R. C. Haddon, J. Am. Chem. Soc. 2001, 123, 11 673. [19] D. Chattopadhyay, S. Lastella, S. Kim, F. Papadimitrakopoulos, J. Am. Chem. Soc. 2002, 124, 728. [20] a) B. Li, Z. Shi, Y. Lian, Z. Gu, Chem. Lett. 2001, 598. b) F. Pompeo, D. E. Resasco, Nano Lett. 2002, 2, 369. [21] M. G. C. Kahn, S. Banerjee, S. S. Wong, Nano Lett. 2002, 2, 1215. [22] B. R. Azamian, K. S. Coleman, J. J. Davis, N. Hanson, M. L. H. Green, Chem. Commun. 2002, 366. [23] S. Banerjee, S. S. Wong, Nano Lett. 2002, 2, 195. [24] a) S. Ravindran, S. Chaudhary, B. Colburn, M. Ozkan, C. S. Ozkan, Nano Lett. 2003, 3, 447. b) J. M. Haremza, M. A. Hahn, T. D. Krauss, S. Chen, J. Calcines, Nano Lett. 2002, 2, 1253. [25] Y.-P. Sun, K. Fu, Y. Lin, W. Huang, Acc. Chem. Res. 2002, 35, 1096. [26] N. W. S. Kam, T. C. Jessop, P. A. Wender, H. Dai, J. Am. Chem. Soc. 2004, 126, 6850. [27] K. A. Williams, P. T. M. Veenhuizen, B. G. de la Torre, R. Eritja, C. Dekker, Nature 2002, 420, 761. [28] a) R. C. Haddon, Science 1993, 261, 1545. b) A. L. Balch, V. J. Catalano, D. A. Costa, W. R. Fawcett, M. Federco, A. S. Ginwalla, J. W. Lee, M. M. Olmstead, B. C. Noll, K. Winkler, J. Phys. Chem. Solids 1997, 58, 1633. c) A. Hirsch, The Chemistry of the Fullerenes, Thieme, Stuttgart, Germany 1994. [29] P. Serp, M. Corrias, P. Kalck, Appl. Catal., A 2003, 253, 332. [30] T. Hemraj-Benny, S. Banerjee, S. S. Wong, Chem. Mater. 2004, 16, 1855. [31] a) S. Banerjee, S. S. Wong, J. Am. Chem. Soc. 2002, 124, 8940. b) S. Banerjee, S. S. Wong, Nano Lett. 2002, 2, 49. [32] F. Nunzi, F. Mercuri, A. Sgamellotti, N. Re, J. Phys. Chem. B 2002, 106, 10 622. [33] V. Lordi, N. Yao, J. Wei, Chem. Mater. 2001, 13, 733. [34] A. M. Bond, W. Miao, C. Raston, Langmuir 2000, 16, 6004. [35] R. Giordano, P. Serp, P. Kalck, Y. Kihn, J. Schreiber, C. Marhic, J.-L. Duvail, Eur. J. Inorg. Chem. 2003, 4, 610. [36] a) S. Banerjee, S. S. Wong, J. Am. Chem. Soc. 2003, 125, 10 342. b) S. Banerjee, S. S. Wong, Adv. Mater. 2004, 16, 34. c) S. Banerjee, S. S. Wong, Chem. Commun. 2004, 1866. [37] Z. Chen, W. Thiel, A. Hirsch, ChemPhysChem 2003, 4, 93. [38] R. C. Haddon, Acc. Chem. Res. 1998, 21, 243. [39] M. A. Hamon, M. E. Itkis, S. Niyogi, T. Alvaraez, C. Kuper, M. Menon, R. C. Haddon, J. Am. Chem. Soc. 2001, 123, 11 292. [40] A. Garg, S. B. Sinnott, Chem. Phys. Lett. 1998, 295, 273. [41] C. W. Padgett, D. W. Brenner, Nano Lett. 2004, 4, 1051. [42] M. S. Dresselhaus, G. Dresselhaus, A. Jorio, A. G. Souza Filho, M. A. Pimenta, R. Saito, Acc. Chem. Res. 2002, 35, 1070. [43] a) H. Hu, B. Zhao, M. A. Hamon, K. Kamaras, M. E. Itkis, R. C. Haddon, J. Am. Chem. Soc. 2003, 125, 14 893. b) K. Kamaras, M. E. Itkis, H. Hu, B. Zhao, R. C. Haddon, Science 2003, 301, 1501. [44] V. N. Khabashesku, W. E. Billups, J. L. Margrave, Acc. Chem. Res. 2002, 35, 1087. [45] K. F. Kelly, I. W. Chiang, E. T. Mickelson, R. H. Hauge, J. L. Margrave, X. Wang, G. E. Scuseria, C. Radloff, N. J. Halas, Chem. Phys. Lett. 1999, 313, 445.

[46] C. W. Bauschlicher, Chem. Phys. Lett. 2000, 322, 237. [47] H. Peng, Z. Gu, J. Yang, J. L. Zimmerman, P. A. Willis, M. J. Bronikowski, R. E. Smalley, R. H. Hauge, J. L. Margrave, Nano Lett. 2001, 1, 625. [48] P. J. Boul, J. Liu, E. T. Mickelson, C. B. Huffman, L. M. Ericson, I. W. Chiang, K. A. Smith, D. T. Colbert, J. L. Margrave, R. E. Smalley, Chem. Phys. Lett. 1999, 310, 367. [49] a) E. T. Mickelson, C. B. Huffman, A. G. Rinzler, R. E. Smalley, R. H. Hauge, J. L. Margrave, Chem. Phys. Lett. 1998, 296, 188. b) E. T. Mickelson, I. W. Chiang, J. L. Zimmerman, P. J. Boul, J. Lozano, J. Liu, R. E. Smalley, R. H. Hauge, J. L. Margrave, J. Phys. Chem. B 1999, 103, 4318. [50] J. L. Stevens, A. Y. Huang, H. Peng, I. W. Chiang, V. N. Khabashesku, J. L. Margrave, Nano Lett. 2003, 3, 331. [51] L. Zhang, V. U. Kiny, H. Peng, J. Zhu, R. F. M. Lobo, J. L. Margarve, V. N. Khabashesku, Chem. Mater. 2004, 16, 2055. [52] X. Lu, L. Zhang, X. Xu, N. Wang, Q. Zhang, J. Phys. Chem. B 2002, 106, 2136. [53] J.-P. Deng, C.-Y. Mou, C.-C. Han, Fullerene Sci. Technol. 1997, 5, 1325. [54] S. Banerjee, S. S. Wong, J. Phys. Chem. B 2002, 106, 12 144. [55] a) S. Banerjee, T. Hemraj-Benny, M. Balasubramanian, D. Fischer, J. A. Misewich, S. S. Wong, Chem. Commun. 2004, 772. b) L. Cai, J. L. Bahr, Y. Yao, J. M. Tour, Chem. Mater. 2002, 14, 4235. [56] S. Banerjee, S. S. Wong, Nano Lett. 2004, 4, 1445. [57] a) J. L. Bahr, J. Yang, D. V. Kosynkin, M. J. Broniskowski, R. E. Smalley, J. M. Tour, J. Am. Chem. Soc. 2001, 123, 6536. b) J. L. Bahr, J. M. Tour, Chem. Mater. 2001, 13, 3823. [58] C. A. Dyke, J. M. Tour, J. Am. Chem. Soc. 2003, 125, 1156. [59] M. Strano, J. Am. Chem. Soc. 2003, 125, 16 148. [60] C. A. Mitchell, J. L. Bahr, S. Arepalli, J. M. Tour, R. Krishnamoorti, Macromolecules 2002, 35, 8825. [61] M. Holzinger, O. Vostrowsky, A. Hirsch, F. Hennrich, M. Kappes, R. Weiss, F. Jellen, Angew. Chem. Int. Ed. 2001, 40, 4002. [62] a) K. S. Coleman, S. R. Bailey, S. Fogden, M. L. H. Green, J. Am. Chem. Soc. 2003, 125, 8722. b) K. A. Worsley, K. R. Moonoosawmy, P. Kruse, Nano Lett. 2004, 4, 1541. c) J. L. Delgado, P. de la Cruz, F. Langa, A. Urbina, J. Casado, J. T. Lopez Navarrete, Chem. Commun. 2004, 1734. d) X. Lu, F. Tian, X. Xu, N. Wang, Q. Zhang, J. Am. Chem. Soc. 2003, 125, 10 459. [63] a) S. Pekker, J.-P Salvetat, E. Jakab, J.-M. Bonard, L. Forro, J. Phys. Chem. B 2001, 105, 7938. b) F. Liang, A. K. Sadana, A. Peera, J. Chattopadhyay, Z. Gu, R. H. Hauge, W. E. Billups, Nano Lett. 2004, 4, 1257. [64] H. Peng, L. B. Alemany, J. L. Margrave, V. N. Khabashesku, J. Am. Chem. Soc. 2003, 125, 15 174. [65] M. Holzinger, J. Abraham, P. Whelan, R. Graupner, L. Ley, F. Hennrich, M. Kappes, A. Hirsch, J. Am. Chem. Soc. 2003, 125, 8566. [66] V. Georgakilas, K. Kordatos, M. Prato, D. M. Guldi, M. Holzinger, A. Hirsch, J. Am. Chem. Soc. 2002, 124, 760. [67] N. Tagmatarchis, M. Prato, J. Mater. Chem. 2004, 14, 437. [68] D. M. Guldi, M. Marcaccio, D. Paolucci, F. Paolucci, N. Tagmatarchis, D. Tasis, E. Vazquez, M. Prato, Angew. Chem. Int. Ed. 2003, 42, 4206. [69] D. Pantorotto, C. D. Partidos, R. Graff, J. Hoebeke, J.-P. Briand, M. Prato, A. Bianco, J. Am. Chem. Soc. 2003, 125, 6160. [70] A. Bianco, M. Prato, Adv. Mater. 2003, 15, 1765.

______________________

Adv. Mater. 2005, 17, No. 1, January 6

http://www.advmat.de

 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

29

You might also like