Sei sulla pagina 1di 6

View Article Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/materials | Journal of Materials Chemistry

Synthesis of poly(vinyl alcohol)/reduced graphite oxide nanocomposites with improved thermal and electrical properties
Horacio Javier Salavagione,* Gerardo Martnez and Marin A. Gmez a o
Received 9th March 2009, Accepted 23rd April 2009 First published as an Advance Article on the web 1st June 2009 DOI: 10.1039/b904232f Poly(vinylalcohol)/reduced graphite oxide nanocomposites have been synthesised by reducing graphite oxide in the presence of the polymer matrix and coagulating the system with 2-propanol. It has been observed that some interactions occur between the polymer and the reduced graphite oxide layers, mainly by hydrogen bonding. These interactions are responsible for a remarkable change in the thermal behaviour of the nanocomposites. In addition, high electrical conductivity has been achieved at concentrations beyond 7.5 wt% of reduced graphite oxide ($0.1 S cm1), with a percolation threshold between 0.5 and 1 wt%.

Downloaded by King Faisal University on 14 February 2013 Published on 01 June 2009 on http://pubs.rsc.org | doi:10.1039/B904232F

Introduction
Polymer-based nanocomposites have been intensively studied in the last decades with the aim to obtain high-performance lightweight materials. In order to achieve this purpose, clay minerals1 and carbon nanostructures14 have been proposed as candidates for the ller material. Several polymeric matrices have been employed as hosts for carbon nanobers (CNF) and nanotubes (CNT) ller, including polyolens,5 poly(acrylates),6 epoxy,7 elastomers,8 etc. However, the expensive multi-step methods used to prepare and purify CNT limit their production on an industrial scale. Graphene, a two-dimensional monolayer of carbon atoms, the basic building block for naturally occurring graphite, emerges as an interesting alternative to CNT due especially to their unique structure and novel electronic properties. As it occurs with CNT, processing and dispersion of graphene in the polymeric hosts constitutes the main challenge to use it in nanocomposites, the control of the interfacial interaction being crucial. The polymer matrix and the ller are bonded to each other by weak intermolecular forces, and chemical bonding is rarely involved. If the reinforcing material in the composite could be dispersed on a molecular scale and interacted with the matrix by chemical bonding, then signicant improvements in the mechanical properties of the material or unexpected new properties might be attained. In that sense, CNT have been successfully functionalized to improve their interaction with several polymeric matrices.9,10 Thus, amino-functionalized CNT have been used as reinforcement of epoxy matrices,11 nylon-6 and other polyamides have been grafted from 3-caprolactamfunctionalized CNT12 and carboxylic functional groups in CNT have been used for etherication with poly(vinyl alcohol) (PVA).13,14 Due to its recent discovery, just a few attempts to modify graphene layers have been reported. However, graphite oxide
Instituto de Ciencia y Tecnologa de Polmeros, CSIC. Juan de la Cierva 3, 28006 Madrid, Spain. E-mail: horacio@ictp.csic.es; Fax: +34 915 644 853; Tel: +34 915 622 900 Electronic supplementary information (ESI) available: Characterization of graphite oxide and FTIR spectra and DSC curves for PVA and nanocomposites. See DOI: 10.1039/b904232f

(GO) has been treated with octadecylamine to produce octadecyl-amido graphite15 and with different isocyanates to produce amides and carbamate esters on the GO surface.16 The authors demonstrated that this treatment resulted in a new class of GO materials with reduced hydrophilic properties which can be exfoliated in polar aprotic solvents to yield derivatized graphene oxide nanoplatelets. Thus, the modied GO can be intimately mixed with polystyrene which is soluble in the same solvents and reduced in-situ to produce polystyrene/graphene nanocomposites.17 On the other hand, the high capacity of PVA to form both intra- and interchain hydrogen bonding simultaneously makes it an interesting polymer which could interact with GO in the same way and could be used as a model for understanding the interaction of graphite oxide with other polar polymers. PVA/GO1820 and PVA/CNT12,13,2123 composites have been prepared by mixing solutions of both components. Here, we report a strategy for the synthesis of PVA/reduced graphite oxide (RGO) nanocomposites by reducing GO in presence of PVA and coagulating the system with 2-propanol to precipitate the nanocomposite. In fact, the interaction between the polymer and the groups remaining on the graphene sheet signicantly alters the thermal properties of the nanocomposites. In addition, the nanocomposites display good electrical conductivity with a low percolation threshold.

Experimental
Chemicals Graphite powder (45 mm), PVA (99+% hydrolyzed, Mw $ 8900098000) and hydrazine ($35% water solution) were purchased from Aldrich. Potassium permanganate (KMnO4), sulfuric acid (H2SO4), hydrogen peroxide (H2O2) and sodium nitrate (NaNO3) were purchased from Panreac, Spain. Methods Graphite oxide (GO) was obtained using Hummers method.24 Briey, 2 g of graphite was mixed with 1 g of NaNO3 and 50 mL of H2SO4, and the mixture was cooled down to 0  C. Then, 6 g
J. Mater. Chem., 2009, 19, 50275032 | 5027

This journal is The Royal Society of Chemistry 2009

View Article Online

KMnO4 was added slowly maintaining the temperature below 5  C. The cooling bath was removed and the suspension was maintained for 0.5 h. After that, 100 mL of water was added and the temperature increased to 90  C. The mixture was further diluted with 300 mL of water, treated with 50 mL of 5% H2O2, ltered and washed with hot water. The graphite oxide was dried by vacuum obtaining a nal mass of 3.2 g. The GO powder was examined by X-ray diffraction (XRD). It was observed that no total oxidation was achieved suggesting that GO contains some original graphite (Fig S1, ESI). Nanocomposites preparation
Downloaded by King Faisal University on 14 February 2013 Published on 01 June 2009 on http://pubs.rsc.org | doi:10.1039/B904232F

PVA/reduced graphite oxide (RGO) nanocomposites with 0.5, 1, 1.5, 2, 5, 7.5 and 10% of ller were prepared as follows: GO was dissolved in 20 mL of water and treated with ultrasound for 15 min. PVA was dissolved in 80 mL water by heating it for 45 min at 90  C and mixed with the GO solution. The mixture was maintained under magnetic agitation for 3 days. Then, the reduction of GO was accomplished by adding 5 mL of hydrazine to the PVA/GO solution and keeping it under magnetic stirring overnight. The coagulation of the polymer nanocomposites was accomplished by adding the aqueous solutions dropwise to 100 mL of 2-propanol under vigorous stirring. The solid nanocomposites were ltered, washed abundantly with methanol and dried at 40  C under vacuum for ca. 24 h. The same procedure, without the reduction step, was employed to prepare PVA/ graphite and PVA/GO composites Techniques The X-ray diffractograms (XRD) of the powder nanocomposites were obtained using a Bruker D8-Advanced X-ray generator (Cu Ka) operated at 40 kV and 40 mA. The dispersion of the nanoller was examined using a Philips XL30 SEM equipment. The nanocomposite samples were cryofractured from lm specimens and then were coated with ca. 5 nm Au/Pd overlayer to avoid charging during electron irradiation. The thermogravimetric analysis was performed in a TGA Q500 equipment from TA instruments. The samples were dried under dynamic vacuum before the experiments and then placed in a platinum pan. The loss of weight was monitored from room temperature to 950  C using a heating rate of 10  C min1. The experiments were carried out in nitrogen atmosphere. The crystallization and melting behaviour were investigated by DSC using a Mettler TA4000/DSC30. The experiments were carried out in nitrogen atmosphere using $5 mg sample sealed in aluminium pans. The samples were heated from room temperature to 240  C, maintained at this temperature for 5 min, then cooled to room temperature and heated again to 240  C. The heating and cooling rates were 10  C min1 in all cases. The transition temperatures were taken as the peak maximum in the calorimetric curves. The degree of crystallinity (Xc) was calculated from the ratio DHa/DHu, which are the apparent and the 100% crystalline melting enthalpy, respectively. The apparent heat of fusion DHa is taken from the area of the fusion peak. The melting enthalpy of a 100% crystalline PVA, DHu, was taken as 161.6 J g1. The crystallinity of PVA in the nanocomposites was
5028 | J. Mater. Chem., 2009, 19, 50275032

determined by considering the weight fraction of PVA in the nanocomposites. Fourier transform infrared spectra (FTIR) of the nanocomposites in the transmission mode were obtained in KBr pellets using a Perkin-Elmer System 200 spectrometer with a 4 cm1 resolution. DC-Conductivity measurements were carried out using the four-probe method on lms. The hot-pressed lms were cut into rectangles (about 0.6 cm wide and 1.2 cm long) and dried under vacuum for 24 h. The measurements were carried out using a four-probe setup equipped with a dc low-current source (LCS02) and a digital micro-voltmeter (DMV-001) from Scientic Equipment & Services. The conductivity was calculated by using eqn (1).   . V s 1 r 4:5324t f1 f2 (1) I where t is the thickness of the sample, f1 is the nite thickness correction for thick samples on an insulating bottom boundary and f2 is the nite width correction.25 The thickness correction is described by eqn (2): f1 ln2  sinht=s ln sinht=2s  (2)

where s is the probe distance (0.2 cm). Usually, the thickness is about 0.05 cm and f1 takes values close to 1. The width correction f2 is a function of the width (d) and length (l) and is obtained from experimental curves in the literature (in our case, l 2d).

Results and discussion


Nanocomposites of PVA/RGO with 0.5, 1, 1.5, 2, 5, 7.5 and 10% of the ller were prepared. In all cases, the starting brownish solution of PVA/GO changed immediately to black when hydrazine was added, suggesting some graphene formation by reduction of GO. However, no precipitate was observed a few hours after the magnetic stirring had been stopped, conrming that no aggregates of the RGO sheets were formed. This behaviour suggests the presence of interactions between the matrix and the RGO which is critical for avoiding RGO sheets aggregation. Fig. 1 shows the XRD patterns of PVA powder and its nanocomposites with different contents of RGO. PVA shows diffraction peaks at 11.3, 19.4 and 22.4 that correspond to the crystalline phase of the polymer.26 In the nanocomposites, the intensity of the main peaks of the PVA decreased and became broader, showing a decrease in crystallinity and indicating that some interactions between polymer chains and the ller may take place. In order to characterize the nanocomposites and to determine the interaction among polymer chains and with the ller, FTIR experiments were carried out. It is known that both the OH stretching and the COH stretching bands are sensitive to the hydrogen bonding degree. The FTIR spectra of PVA and the nanocomposites contain typical bands of PVA (Fig S2). However, the broad band around 3400 cm1, involving the strong hydroxyl band for free and hydrogen bonded alcohols, shifts to a lower wavenumber as the amount of ller increases.
This journal is The Royal Society of Chemistry 2009

View Article Online

Fig. 1 X-Ray diffraction patterns of PVA and PVA/RGO nanocomposites.

This behaviour has been attributed to the dissociation of the hydrogen bonding among the hydroxyl groups in the polymer.27 Furthermore, the band assigned to the COH stretching (around 1100 cm1) displayed a similar behaviour, indicating hydrogen bonding between the OH in PVA and the oxygenated groups remaining in RGO.20,27 These band displacements are represented in Fig. 2. A significant effect is observed at low ller concentration. The effect decreased when the ller concentration increased over 2% probably due to the association between RGO laminates which became important and should be considered at these loadings. Although the trend in the FTIR bands is clear, factors other than the hydrogen bonding interactions can also be involved (e.g., steric effects). However, it is clear that the addition of RGO decreases the crystallinity of the polymer (Fig. 1). The band around 1155 cm1 has already been related to the crystallinity of

PVA.28 The relative intensity of this band (related to the band at ca. 1100 cm1) decreases as the amount of RGO increases showing the same trend described by the shifts of the COH and OH bands (inset in Fig. 2). These results may reinforce the hypothesis for the existence of hydrogen bonding between the polymer and the ller, to the detriment of hydrogen bonding among polymer chains, that diminishes the crystalinity. Before we analyze the physical properties of the nanocomposites, it is important to note that the ller is composed of graphite and reduced graphite oxide (see ESI). In order to investigate the inuence of non-exfoliated graphite on the physical properties of the nanocomposites, 5 wt% PVA/graphite and PVA/GO composites were prepared. These compounds will be used to compare their properties with the ones of PVA/graphene nanocomposite. In order to determine the degree of exfoliation and the quality of the nanoller dispersion in PVA, scanning electron microscopy (SEM) was employed. Fig. 3 shows the SEM images of the fracture surface of 2 and 5 wt% PVA/RGO nanocomposites. Composites of PVA/graphite and PVA/GO (not reduced) are also shown for comparison. A well-dispersed and densely packed distribution of GO and RGO can be seen in 5 wt% nanocomposites. In both cases, graphene platelets more than 5 mm long and about 5060 nm thick can be observed. However, it seems that the voids between the laminates are slightly higher in PVA/GO. Also, some agglomerates appear (around 0.3 mm) due to not total oxidation of the graphite (Fig S1a). In fact, the 5 wt% PVA/graphite composite shows ller dimensions of about 0.20.3 mm thick and less packed distribution (Fig. 3b). Thus, it can be stated that the agglomerates are due to original graphite materials and the polymer plays a key role to prevent further sheet agglomeration after the reduction step. The 2 wt% PVA/ RGO nanocomposite displayed separated nanoplatelets of graphene (the white dots are due to the low quality of the fracture and the cracks are due to the sensitivity of the PVA to the electron beam)

Downloaded by King Faisal University on 14 February 2013 Published on 01 June 2009 on http://pubs.rsc.org | doi:10.1039/B904232F

Fig. 2 Displacements of the OH and COH stretching bands with the RGO content in the nanocomposites. Inset: variation of the height ratio between 1155 and 1100 cm1 bands with the RGO contents.

Fig. 3 SEM images of PVA/GO 5 wt% (a), PVA/graphite 5 wt% (b), PVA/RGO 2 wt% (c) and 5 wt% (d) nanocomposites. The scale bar corresponds to 2 mm.

This journal is The Royal Society of Chemistry 2009

J. Mater. Chem., 2009, 19, 50275032 | 5029

View Article Online


Table 1 DSC crystallization and melting parameters of PVA and PVA/ RGO nanocomposites Sample PVA PVA/RGO 1 wt% PVA/RGO 2 wt% PVA/RGO 5 wt% PVA/RGO 10 wt% PVA/graphite 5 wt% PVA/GO 5 wt%
a

Tg/ C 82 84 94 94 103 84 87

Tm/ C 225 214 200 189 160 225 221

Tc/ C 203 196 180 169 206 202

DHm/J g1 83 60.9 54.5 33.2 5.2 70.1 55.0

Xc (%)a 51 37 33 19 3 43 34

Obtained from the melting endotherms.

Downloaded by King Faisal University on 14 February 2013 Published on 01 June 2009 on http://pubs.rsc.org | doi:10.1039/B904232F

The crystallization behaviour and crystalline parameters of the polymer matrix in the composite were also investigated because the macroscopic properties of the nanocomposites strongly depend on them. The crystallization and melting behaviour were analyzed by cooling down from the melt to 30  C at 10  C min1 and subsequent heating up, in the same experiment, to 240  C (Fig S3). The parameters derived from the experimental DSC curves are summarized in Table 1. Signicant changes can be observed in the crystallization temperature (Tc), the melting temperature (Tm) and the glass transition temperature (Tg) of PVA in the nanocomposites depending on ller content. Regarding the heating scan, it can be observed that Tm and the melting enthalpy (DHm) are strongly inuenced by the ller loading. As the RGO content increases, the melting endotherm of PVA in the nanocomposite broadens and Tm decreases. In addition, the crystallinity of PVA in the nanocomposites decreases from $51% (for PVA) to an almost amorphous material at 10 wt% of RGO (Table 1), supporting the existence of some interaction between the polymer and the ller to the detriment of interactions among polymer chains. These results are conrmed by the DSC heating scan of PVA/graphite composite, where a sharp melting peak with the same Tm and slightly lower Xc with respect to PVA is observed (Fig S3 in ESI, Table 1). Therefore, the rests of the graphite in the ller behaves as inert material in the nanocomposite and the reduced graphite oxide platelets are responsible for the changes in the properties. However, the melting behaviour of PVA/GO displays a broad melting peak (similar to the nanocomposites), but with Tm values close to that for PVA and PVA/graphite. It appears that the reduction step, which eliminates oxygenated groups, changes the interaction between the ller and the polymer. Therefore, additional effects to the hydrogen bonding are also involved (e.g., steric effects). In addition, the Xc of PVA/GO lies between the values for PVA/RGO 1 and 2 wt%, suggesting that the rigidity of the nanocomposite is altered after the reduction step. These changes should affect the mobility of the chains. In order to investigate this effect, the glass transition region was analyzed (Fig S3, ESI). A remarkable increase in the Tg is observed as the ller loading increases over 1 wt% (Table 1). This effect can be ascribed to an effective attachment of the polymer to the layers of RGO that prevents the segmental motions of the polymer chains, as occurs with different polymer matrices intercalated in laminar montmorillonite (MMT).29 However, the Tg for PVA/GO composite is similar to that for PVA. Therefore, the removal of oxygenated groups causes the steric effect to
5030 | J. Mater. Chem., 2009, 19, 50275032

decrease, effecting more compact packaging, which increases the rigidity of the system. The DSC cooling scans show a strong dependence on the ller nature and content. The Tc decreases signicantly as the ller loading increases in the RGO composites, which indicates that the crystallization is retarded. On the other hand, no changes are observed in Tc for PVA/GO and PVA/graphite nanocomposites. The differences in crystallization behaviour observed for the composites can be explained considering the existence of interactions between the ller and the matrix in the reduced graphite oxide nanocomposites as was previously pointed out from other experimental evidences (FTIR and DRX), and they suggest that some steric effects are eliminated through the reduction step. In order to investigate the effects of the RGO on the thermal stability of the polymer matrix, thermo-gravimetric analysis was performed on all composites (Fig. 4). TGA curves for PVA/RGO suggests that the ller plays an important role to improve the thermal stability of the nanocomposites. The temperature of the maximum degradation rate for the nanocomposites (obtained from the derivative of TGA curves, Fig 4b) increases by more than 100  C. This behaviour has been attributed, in similar systems, to the stability of the hydrogen bonding between the polymer and the GO.20 In our case, most of the oxygenated

Fig. 4 Thermogravimetric curves of PVA and PVA/RGO nanocomposites (a) weight loss percent and (b) derivative curves.

This journal is The Royal Society of Chemistry 2009

View Article Online

groups were reduced to produce graphene sheets. However, it has been proposed that the carboxylic groups on the border of the GO laminates are not reduced under the reaction conditions we used.30,31 In fact, the PVA/GO composite displays the same behaviour with a temperature of maximum degradation rate around 365  C (not shown). Fig 4 also shows that although a small weight loss is observed around 200  C commonly in these systems; the onset of PVA decomposition for the composites is retarded. Similar behaviour has been observed for PVA/CNT composites with less than 20% of ller and has been attributed to the absorption, by the carbon surface, of free-radicals generated during polymer decomposition.21 Here, a low quantity of RGO (0.5 wt%) produces the same effect, due to their higher aspect ratio. One of the main challenges of nanocomposites of insulating polymers and very conductive llers is to achieve materials with improved electrical properties. The llers, normally CNT and graphite nanosheets, exhibited a high aspect ratio which is advantageous for forming conducting networks in a polymer matrix. The conductivity of the PVA/RGO nanocomposites as a function of the ller loading has been measured by the fourprobe method (Fig. 5). Three-stage behaviour can be observed in which two different steps occurred in conductivity. The rst increase can be assigned to the typical insulating-conductive percolation behaviour. That is, conductivity remains at very low values (1013 to 1014 S cm1) until a critical loading content is achieved. Then, it increases quickly by several orders of magnitude up to 8 106 S cm1. The percolation threshold lies between 0.5 and 1 wt% of RGO. If we use the density of PVA (1.312 g cm3) and assume the graphite oxide density as 2.2 g cm3 (similar to ref. 17) to convert the values to vol% this value changes to 0.30.6% which is a little higher than those obtained for polystyrene/graphene,17 and close to poly(methyl methacrylate)(PMMA)/graphite nanosheets.32 The higher percolation threshold in our case seems reasonable because of the presence of non-exfoliated graphite, which diminishes the ller aspect ratio. However, the reported percolation threshold in similar systems23 composed of PVA and CNT (between 5 and 10 wt%) is quite higher than the one obtained in this work. This difference could be attributed to the previous acidic treatment of

CNT, which generates a higher degree of structural defects that causes the material to lose its special electronic properties.33,34 Also, under these conditions very short tubes with a low aspect ratio are produced.34 Therefore, although original graphite is present, the exfoliation and reduction of the oxidised laminates are very important steps. In addition the second increase of conductivity occurs between 5 and 10 wt% ($3 to 6.2 vol%). This transition has also been observed in PMMA/graphite32 and in epoxy/bromine-treated graphite nanoplatelets composites35 and has been attributed to hopping of electrons between ller laminates by tunnelling through the polymer layers.32 We think that tunnelling also occurs in our system but we have not explanation for the fourth order of magnitude increase. Here, it is very important to emphasize that for 5 wt% PVA/ graphite and PVA/GO composites no measurable conductivity values were obtained. It is well known that GO is intrinsically an insulating material. However, although graphite is a conductive material, this did not confer conductivity to the composite. Although the maximum conductivity value obtained for 7.5 wt% PVA/RGO is slightly lower than those reported for other systems (PS/graphene17 and PMMA/graphite32), it is rather higher than the one reported for PVA/CNT composites, which requires more than 20 wt% of CNT to achieve conductivity in the same order of magnitude.23 Taking into account the results described above, it can be noted that (i) the presence of non-exfoliated graphite is not responsible for the increase in the conductivity of PVA/RGO, and (ii) the reduction step is crucial to eliminating the functional groups that limit conductivity, and (iii) the polymer plays a key role in avoiding ller aggregation. In addition, the system described in this paper could be optimized by controlling some other parameters such as the degree of oxygenated groups in the graphene layers and the type of interaction with the matrix (e.g., covalent bonds). In fact, we used partially oxidized graphite as starting material. Thus, by increasing the degree of oxidation (and thus, the degree of exfoliation) better nanocomposite properties are expected.

Downloaded by King Faisal University on 14 February 2013 Published on 01 June 2009 on http://pubs.rsc.org | doi:10.1039/B904232F

Conclusions
A successful method to produce PVA/RGO nanocomposites, based on the reduction of graphite oxide in the presence of the polymer was developed. The nanocomposites prepared by this procedure display interactions among the polymer and the ller due to the remaining oxygenated groups in graphene. These interactions signicantly alter the thermal properties of the nanocomposites, lowering the Tm more than 35  C and decreasing crystallinity down to a half for nanocomposites with 5 wt% of RGO. The materials display good electrical conductivity with a percolation threshold below 1 wt% of ller and conductivities as high as 0.1 S cm1 for composites with 7.5% wt of RGO.

Acknowledgements
Fig. 5 Plot of the electric conductivity of PVA/RGO nanocomposites as a function of the RGO content.

Financial support from the Spanish Ministry of Science and Innovation, MICINN (MAT 2006-13167-C02-01) is gratefully acknowledged. H.J.S. thanks the Consejo Superior de
J. Mater. Chem., 2009, 19, 50275032 | 5031

This journal is The Royal Society of Chemistry 2009

View Article Online

Investigaciones Cientcas (CSIC) for a postdoctoral contract JAE-Doc. The authors thank Dr Gary Ellis and Dr Carlos Marco for helpful discussions.

References
1 D. R. Paul and L. M. Robeson, Polymer, 2008, 49, 31873204. 2 J. N. Coleman, U. Khan and Y. K. Gunko, Adv. Mater., 2006, 18, 689706. 3 M. Moniruzzaman and K. I. Winey, Macromolecules, 2006, 39, 5194 5205. 4 N. Grossiord, J. Loos, O. Regev and C. E. Koning, Chem. Mater., 2006, 18, 10891099. 5 A. Linares, J. C. Canalda, E. Cagiao, M. C. Garca-Gutirrez, e A. Nogales, I. Martn-Gulln, J. Vera and T. A. Ezquerra, o Macromolecules, 2008, 41, 70907097. 6 G. L. Hwang, Y.-T. Shieh and K. C. Hwang, Adv. Funct. Mater., 2004, 14, 487491. 7 B. L. Wardle, D. S. Saito, E. J. Garca, A. J. Hart, R. Guzmn de a Villoria and E. A. Verploegen, Adv. Mater., 2008, 20, 27072714. 8 S. V. Ahir, Y. Y. Huang and E. M. Terentjev, Polymer, 2008, 49, 38413854. 9 S. Wang, R. Liang, B. Wang and C. Zhang, Chem. Phys. Lett., 2008, 457, 371375. 10 L. Dai and A. W. H. Mau, Adv.Mater., 2001, 13, 899913. 11 J. Shen, W. Huang, L. Wu, Y. Hu and M. Ye, Composite Sci. Technol., 2007, 67, 30413050. 12 Y. Lin, M. J. Meziani and Y.-P. Sun, J. Mater. Chem., 2007, 17, 1143 1148. 13 M. C. Paiva, B. Zhou, K. A. S. Fernando, Y. Lin, J. M. Kennedy and y.-P. Sun, Carbon, 2004, 42, 28492854. 14 Y. Lin, B. Zhou, K. A. S. Fernando, P. Liu, L. F. Allard and Y.P. Sun, Macromolecules, 2003, 36, 71997204. 15 S. Nigoyi, E. Bekyarova, M. E. Itkis, J. L. McWilliams, M. A. Hamon and R. C. Haddon, J. Am. Chem. Soc., 2006, 128, 77207721. 16 S. Stankovich, R. D. Piner, S. T. Nguyen and R. S. Ruoff, Carbon, 2006, 44, 33423347.

Downloaded by King Faisal University on 14 February 2013 Published on 01 June 2009 on http://pubs.rsc.org | doi:10.1039/B904232F

17 S. Stankovich, D. A. Dikin, G. H. B. Dommett, k. M. Kohlhass, E. J. Zimney, E. A. Stach, R. D. Piner, S. T Nguyen and R. S. Ruoff, Nature, 2006, 442, 282286. 18 H. Kaczmarek and A. Podgrski, Polym. Deg. & Stab., 2007, 92, 939 o 946. 19 J. Xu, Y. Hu, L. Song, Q. Wang, W. Fan and Z. Chen, Carbon, 2002, 40, 445467. 20 Y. Matsuo, K. Hatase and Y. Sugie, Chem. Mater., 1998, 10, 2266 2269. 21 K. P. Ryan, M. Cadek, V. Nicolosi, S. Walker, M. Ruether, A. Fonseca, J. B. Nagy, W. J. Blau and J. N. Coleman, Synth. Met., 2006, 156, 332335. 22 J. N. Coleman, M. Cadek, R. Blake, V. Nicolosi, K. P. Ryan, C. Belton, A. Fonseca, J. B. Nagy, Y. K. Gunko and W. J. Blau, Adv. Funct. Mater., 2004, 14, 791798. 23 M. S. P. Shaffer and A. H. Windle, Adv. Mater., 1999, 11, 937941. 24 W. S. Hummers, Jr. and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339. 25 Semiconductor Material and Device Characterization, ed. D. K. Scroedeer, John Willey and Sons, New Jersey, 2006. 26 L. A. Garca-Cerda, M. U. Escare~o-Castro and M. Salazarn Zertuche, J. Non-Cryst. Solids, 2007, 353, 808810. 27 L. Lu, H. Sun, F. Peng and Z. Jiang, J. Membrane Sci., 2006, 281, 245252. 28 S. K. Mallapragada and N. A. Peppas, J. Polym Sci. Part B: Polym. Phys., 1996, 34, 13391346. 29 A. Leszczynska and K. Pielichowski, J. Therm. Anal. Calorim., 2008, 93, 677687. 30 S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruoff, Carbon, 2007, 45, 15581565. 31 D. Li, M. B. Mller, S. Gilje, R. B. Kaner and G. G. Wallace, Nat. u Nanotechnol., 2008, 3, 101105. 32 G. Chen, W. Weng, D. Wu and C. Wu, Eur. Polym. J., 2003, 39, 2329 2335. 33 M. A. Hamon, J. Chen, H. Hu, Y. Chen, M. E. Itkis, A. M. Rao, P. C. Eklund and R. C. Haddon, Adv. Mater., 1999, 11, 834840. 34 A. Hirsch, Angew. Chem. Int. Ed., 2002, 41, 18531859. 35 J. Li, L. Vaisman, G. Marom and J.-K. Kim, Carbon, 2007, 45, 744 750.

5032 | J. Mater. Chem., 2009, 19, 50275032

This journal is The Royal Society of Chemistry 2009

Potrebbero piacerti anche