Sie sind auf Seite 1von 4

LETTERS

PUBLISHED ONLINE: 7 DECEMBER 2008 | DOI: 10.1038/NNANO.2008.365

Gram-scale production of graphene based on solvothermal synthesis and sonication


Mohammad Choucair1, Pall Thordarson1 and John A. Stride1,2 *
Carbon nanostructures have emerged as likely candidates for a wide range of applications, driving research into novel synthetic techniques to produce nanotubes, graphene and other carbonbased materials. Single sheets of pristine graphene have been isolated from bulk graphite in small amounts by micromechanical cleavage1, and larger amounts of chemically modied graphene sheets have been produced by a number of approaches27. Both of these techniques make use of highly oriented pyrolitic graphite as a starting material and involve labour-intensive preparations. Here, we report the direct chemical synthesis of carbon nanosheets in gram-scale quantities in a bottom-up approach based on the common laboratory reagents ethanol and sodium, which are reacted to give an intermediate solid that is then pyrolized, yielding a fused array of graphene sheets that are dispersed by mild sonication. The ability to produce bulk graphene samples from nongraphitic precursors with a scalable, low-cost approach should take us a step closer to real-world applications of graphene. Graphite is the most common allotrope of carbon. The name is derived from the Greek verb graphein, to write, which relates very literally to the compound we now know as graphene, as single sheets of graphene were rst isolated by simply tracing a sample of bulk graphite across a substrate in a process known as micromechanical cleavage1. Although theoretical work by Landau8 and others suggests that true two-dimensional structures cannot be isolated, recent work has resolved this apparent anomaly by showing that free-standing graphene sheets are actually buckled, with the ripples extending into the third spatial dimension9. Further attempts to synthesize isolated graphene have been based on intercalation2, sonication in various solvents3,4 and the chemical reduction of graphite to yield few-layer graphite oxide57,10. Approaches developed for the production of carbon nanotubes have also been explored, but so far they have only been able to produce graphitic lm11. However, single- and few-layer graphene sheets have been grown epitaxially by the chemical vapour deposition of hydrocarbons on metal and nonmetal substrates1214, substrate-free deposition15 and by thermal decomposition of SiC16,17. Given the lack of a reliable top-down approach for the large-scale production of graphene, attention has turned to bottom-up approaches that might be able to deliver the economies of scale that are found in the chemical and pharmaceutical industries. We have demonstrated that single-layer graphene can be synthesized by low-temperature ash pyrolysis of a solvothermal product of sodium and ethanol, followed by gentle sonication of the nanoporous carbon product (see Methods). The product of this reaction is then washed in water and dried to obtain pure graphene (Fig. 1). Solvothermal reactions have been reported to lead to the production of carbon nanosheets, although such materials do not appear to consist of single sheets having both high degrees of planarity
1

Figure 1 | Example of the bulk quantity of graphene product. The image consists of approximately 2 g of sample.

and crystallinity, suggesting an amorphous or disordered graphitic structure18,19. The graphene sheets produced herein have been characterized by a number of physical methods that clearly demonstrate the nanostructured form of the carbon product. Representative transmission electron microscopy (TEM) images of the as-synthesized graphene samples post-sonication are shown in Fig. 2, and scanning electron microscopy (SEM) images of the sample before sonication are shown in Fig. 3. The Raman spectra of the synthesized graphene were recorded using both green (514 nm) and red (633 nm) laser radiation, along with samples of natural graphite and charcoal for comparison (see Supplementary Information, Fig. S2). The spectra for natural graphite and charcoal agree well with the literature20, and the graphene samples gave Raman signals in the green laser similar to that of highly defective graphitea broad D-band centred at 1,353 cm21 and a G-band at 1,590 cm21, which have similar intensities such that IG/ID % 1.16. Samples characterized in the red laser showed similar results with an increased intensity of the D-peak across all samples. The denitive evidence that the carbon sheets obtained by this method are only a single atomic layer thick was obtained with atomic force microscopy (AFM) (Fig. 4; see also Supplementary Information, Figs S15 and S16). Multiple cross-sections were recorded of the sample laid onto a mica substrate, and a number of step heights were measured across the relief of the sample. The step heights measured between the surface of the sheets and the sub strate were consistently found to be 4+1 A, proving them to be only a single atom thick.

School of Chemistry, University of New South Wales, Sydney 2052, Australia, 2 Bragg Institute, Australian Nuclear Science and Technology Organisation, PMB 1, Menai, New South Wales 2234, Australia; *e-mail: j.stride@unsw.edu.au
NATURE NANOTECHNOLOGY | VOL 4 | JANUARY 2009 | www.nature.com/naturenanotechnology 2009 Macmillan Publishers Limited. All rights reserved.

30

NATURE NANOTECHNOLOGY

DOI: 10.1038/NNANO.2008.365

LETTERS

Figure 2 | TEM images of the agglomerated graphene sheets. a,b, The same sample region is viewed at different magnications and clearly shows the extent of sheet formation and the tendency for sheets to coalesce into overlapped regions. Scale bars, 200 nm (a) and 1,000 nm (b). An inherent sheet-like structure displaying an intricate long-range array of folds is evident. As the images are taken in transmission mode, the relative opacity of individual sheets is a result of interfacial regions with overlap between individual sheets. The sheets extend in lateral dimensions over micrometre length scales, ranging from 1 1027 to greater than 1 1025 m.

Figure 3 | SEM image of the as-synthesized graphene structures. a,b, The bulk graphene product obtained from the pyrolysis of the solvothermal product is highly porous, but consists of individual sheets. Scale bars, 15 mm (a) and 6 mm (b). The entire image consists of individual graphene sheets held into a porous structure that typically extends over more than 1 1024 m, with the presence of numerous cavities, or holes. The graphene is therefore initially obtained as fused sheets, weakly held into a foam-like structure that is then easily separated into individual sheets by sonication in ethanol for several minutes.

The crystallinity of the sheets was determined with selected area electron diffraction (SAED; Fig. 5), which showed that over length scales that are at least the order of the coherence length of the electron beam, the graphene sheets are crystalline. Indeed the image shows a number of features that demonstrate the sheet-like nature p of the material. Only the in-plane hk(d/n) and h0( 3d/n) graphite reections were observed, where d is the C C bond length of graph ite (1.35 A), with no indication of the (00l ) peaks that result from the interplanar correlations. The fact that the diffraction spots are relatively diffuse and of constant intensity indicates that the sample consists of free sheets that may not be exactly perpendicular to the incident beam9. The bulk conductivity of the material when pressed into a disc of 15 mm diameter was $0.05 S m21, which is higher than the value for charcoal (1 1027 S m21; ref. 21) but much lower than that for graphite ($3 104 S m21; see Supplementary Information and ref. 22). This is consistent with highly conductive sheets having multiple interfaces rather than a highly disordered amorphous product. In addition, surface area measurements indicate that in the solid state, the fused graphene structure has a Langmuir23 (nitrogen, 77 K) surface area of 612 m2 g21, whereas in dispersed ethanol solutions, methylene blue adsorption24 led to a

value of 1,692 m2 g21. The apparent disparity is due to the dissipation of free sheets in dilute solution in contrast to the solid state. Finally, X-ray photoemission spectroscopy (XPS) indicated that the initial crude carbon product obtained directly from the pyrolysis of the solvothermal ethoxide product contained sodium oxide, Na2O, as a 16.57% impurity by weight. However, further washing with distilled water removed this completely. The elemental analysis of the residual graphene, as determined by XPS, was 86.4% C and 13.6% O by atomic composition (or by mass, 82.65% C and 17.35% O), and natural graphite was determined to be 97.8% C, 1.8% O, 0.2% Cl and 0.2% S by atomic composition (by mass, 96.6% C, 2.3% O, 0.5% Cl and 0.6% S). These gures should be compared with those obtained from elemental analysis by combustion (see Supplementary Information): graphene was 78.5% C, 2.7% H and 18.8% O; graphite was 94.5% C, 0.1% H and 5.4% O. Most of the oxygen present in all of these samples is believed to be in the form of adsorbed water: this conclusion is supported by the release of comparable quantities of oxygen under thermo-gravimetric analysis (TGA), with a 7.8% weight loss of the graphene product being recorded at around 110 8C (see Supplementary Information, Fig. S2). Although the overall synthetic process is conceptually very simple, a delicate interplay has been identied in the solvothermal
31

NATURE NANOTECHNOLOGY | VOL 4 | JANUARY 2009 | www.nature.com/naturenanotechnology 2009 Macmillan Publishers Limited. All rights reserved.

LETTERS

NATURE NANOTECHNOLOGY
250 200 Intensity 150
22

DOI: 10.1038/NNANO.2008.365

11

Sample Substrate

100 50 0 0.5 250 1.0 1.5 2.0 2.5 2 (deg)

33

3.0

3.5

4.0

10 20 30

c Intensity

200 150 100 50

Sample Substrate

10 8 Height () 6 4 2 0 0 100 300 400 200 Distance along x-axis (nm) 500 600

40

50 0 0.5 1.0 1.5 2.0 2.5 2 (deg) 3.0

60 3.5 4.0

10 8 Height () 6 4 2 0 0 100 300 400 200 Distance along x-axis (nm) 500 600

Figure 4 | AFM image of graphene in tapping mode. a, Topography image (600 600 nm; height scale, 010 A; scale bar, 100 nm). b,c, Height proles (600 nm along the x-axis) obtained from positions b and c indicated by white arrows in a. To the bottom of the image, a narrow ridge, of around 1 1027 m width, extends beyond the bulk of the sheet. The prole across this point of the image is fully consistent with those observed elsewhere and clearly shows the narrow valley between the ridge and the neighbouring sheet. The presence of irregularly shaped sheets is a result of the sonication of the fused sheets, which fragment upon sonication.

Figure 5 | SAED pattern of graphene at 300 keV. a,b, Integrated intensities p of the in-plane hk (d/n) reections (a) and the h0 ( 3d/n) reections (b). Peaks are labelled with the Miller indices based upon a hexagonal lattice. c, Digitized image of a photographic image plate of the same data, showing the two sets of overlapping hk and h0 reections, off set by 198 as a result of the exposure of two graphene sheets. The image is a composite of two patterns, rotated relative to one another by 198, with one strong (h0) reection leading to secondary diffraction peaks about one of the reciprocal lattice points, which are also rotated by 198 to one of the main series of lines. This is consistent with the beam striking a region of the sample in which there is a partial overlap of two sheets with no interplanar stacking coherence between them and in which the lower layer (the second sheet in the direction of travel of the beam) is slightly rotated (198) with respect to the rst.

graphene precursor. When prepared from a 1:1 molar ratio of sodium and ethanol, the simple oxidation of sodium and reduction of ethanol to crystalline sodium ethoxide was not found to be the dominant product. Under the conditions of the closed, heated reaction vessel, the alcoholic solution becomes increasingly saturated with the metal alkoxide as it forms, and as a result of the autogenerated pressure ($1 102 bar), the free alcohol is encapsulated into the metal alkoxide in a clathrate-like structure. TGA conducted on the product of the solvothermal step conrmed that the solid contained 29.0+2.5% by weight of water and ethanol. The solid solvothermal product is not purely the ethoxide ethanol clathrate, as some metal hydroxide may also form in the presence of water. In addition, some residual sodium may also initially be dispersed into the solid metal alkoxide, but rapidly hydrolyses to sodium hydroxide when exposed to water vapour in the air, consistent with the discoloration of the solid from light grey to white. Owing
32

to the small isolated nature of the trapped ethanol as ignition points, nucleation of the sheets occurs around those regions rich in ethanol, in what we have labelled a popcorn effect. The presence of excess ethanol ($30%) in the clathrate-like structure has been found to be thermodynamically unstable, having a tendency to evaporate out of the structure when exposed to air. Experiments performed on non-clathrated, crystalline sodium ethoxide were not found to result in the formation of graphene sheets, indicating that the solvothermal reaction conditions are essential to produce the metastable ethanol-rich graphene precursor. The production of large quantities of high-purity graphene sheets will allow for the development of large-scale applications of this unique material. The study of the relative surface areas of at and porous graphenes is currently under way. Chemically synthesized graphene is now available in sufcient quantities to seriously consider its use in a wide range of applications, including composites reinforcement, electronic devices, batteries, sensors, hydrogen storage and catalysis.

Methods
All solvothermal reactions were performed in a Teon-lined Parr Instrument Company 4749 reactor having a maximum volume of 23 ml. A typical synthesis consists of heating a 1:1 molar ratio of sodium (2 g) and ethanol (5 ml) in a sealed reactor vessel at 220 8C for 72 h to yield the solid solvothermal productthe

NATURE NANOTECHNOLOGY | VOL 4 | JANUARY 2009 | www.nature.com/naturenanotechnology 2009 Macmillan Publishers Limited. All rights reserved.

NATURE NANOTECHNOLOGY

DOI: 10.1038/NNANO.2008.365

LETTERS

graphene precursor. This material is then rapidly pyrolysed, and the remaining product washed with deionized water (100 ml). The suspended solid is then vacuum ltered and dried in a vacuum oven at 100 8C for 24 h. The nal yield of graphene is approximately 0.1 g per 1 ml of ethanoltypically yielding $0.5 g per solvothermal reaction. The AFM measurements were performed with a Pico SPM II system (Molecular Imaging) in acoustic a.c. mode. Ethanol suspensions of graphene were applied directly on freshly cleaved mica, allowed to dry, and the measurements performed in air at ambient temperature and pressure. The cantilevers used were silicon cantilevers NSG01 (NT-MDT) with a spring constant of 3 10 N m21, and resonance frequency of 110 190 kHz. Images were analysed using PicoScan 5 image-processing software supplied by Molecular Imaging. The gures shown in this article were obtained in topography mode and have been levelled by rst-order plane subtraction, as indicated. The height proles were all obtained from topography images that have been levelled by rst-order plane subtraction. No other data manipulation was done on the images. X-ray photoemission spectroscopy (XPS) measurements were conducted with a ESCALAB220i-XL (VG Scientic) system. Standard conditions required a vacuum better than 2 1029 mbar, with a monochromatic aluminium Ka source and source power of 200 W (10 kV 20 mA). The spot size used was $1 mm in cross scans, with a pass energy of 100 eV for wide scans. TEM images were obtained using a Philips CM200, and SEM images were obtained using a Hitachi S900. The SAED images were recorded at 300 keV using a JEOL 3000F at sample-to-plate distances of 800 and 1,500 mm for the sample and substrate, with exposure times of 0.1 and 0.3 s, respectively. The sample was deposited from suspension onto amorphous carbon substrates and dried by evaporation. Photographic image plates were taken and subsequently digitized into 16-bit images and treated using the public domain software ImageJ 1.38 (http://rsb.info.nih.gov/ij/). Proles were taken along the preferred orientations for all six symmetry equivalent reection sets for each graphene sheet (the image is a composite of two sets of reections off set by 198) and then numerically averaged. The substrate prole was obtained by performing the same procedure across the diffuse rings and normalized for exposure time. TGA was performed on a HiRes Modulated TGA 2950 Thermo-Gravimetric Analyzer (TA Instruments). Data was interpreted using Thermal Advantage v1.1A software. Raman spectroscopy was performed on an inVia Renishaw Raman microscope using green (514 nm) laser excitation, and on a Renishaw R-2000 Raman microscope for red (633 nm) laser excitation. Scans were taken on an extended range (100 3,500 cm21) and the exposure time was 60 s for the green laser and 10 s for the red laser. Samples were sonicated in ethanol and drops applied to a glass slide for observation. The sample was viewed using a green laser apparatus under a maximum magnication of 50, and a red laser apparatus under a magnication of 100.

7. Li, D., Muller, M. B., Gilje, S., Kaner, R. B. & Wallace, G. G. Processable aqueous dispersions of graphene nanosheets. Nature Nanotech. 3, 101 105 (2008). 8. Landau, L. D. Zur theorie der phasenumwandlungen II. Phys. Z. Sowjetunion 11, 26 35 (1937). 9. Meyer, J. C. et al. The structure of suspended graphene sheets. Nature 446, 60 63 (2007). 10. Dikin, D. A. et al. Preparation and characterization of graphene oxide paper. Nature 448, 457 460 (2007). 11. Krishnan, A. et al. Graphitic cones and the nucleation of curved carbon surfaces. Nature 388, 451 454 (1997). 12. Land, T. A., Michely, T., Behm, R. J., Hemminger, J. C. & Comsa, G. STM investigation of single layer graphite structures produced on Pt(111) by hydrocarbon decomposition. Surf. Sci. 264, 261 270 (1992). 13. Nagashima, A., Nuka, K., Itoh, H., Ichinokawa, T. & Oshima, C. Electronic states of monolayer graphite formed on TiC(111) surface. Surf. Sci. 291, 93 98 (1993). 14. Malesevic, A. et al. Synthesis of few-layer graphene via microwave plasma-enhanced chemical vapour deposition. Nanotechnology 19, 305604 (2008). 15. Dato, A., Radmilovic, V., Lee, Z., Phillips, J. & Frenklach, M. Substrate-free gas-phase synthesis of graphene sheets. Nano Lett. 8, 2012 2016 (2008). 16. Forbeaux, I., Themlin, J. M. & Debever, J. M. Heteroepitaxial graphite on 6H-SiC(0001): Interface formation through conduction-band electronic structure. Phys. Rev. B 58, 16396 16406 (1998). 17. van Bommel, A. J., Crombeen, J. E. & van Tooren, A. LEED and Auger electron observations of the SiC(0001) surface. Surf. Sci. 48, 463 472 (1975). 18. Kuang, Q. et al. Low temperature solvothermal synthesis of crumpled carbon nanosheets. Carbon 42, 1737 1741 (2004). 19. Shen, J.-M. & Feng, Y.-T. Formation of ower-like carbon nanosheet aggregations and their electrochemical application. J. Phys. Chem. C 112, 13114 13120 (2008). 20. Kawakami, M., Karato, T., Takenaka, T. & Yokoyama, S. Structure analysis of coke, wood charcoal and bamboo charcoal by Raman spectroscopy and their reaction rate with CO2. ISIJ Int. 45, 1027 1034 (2005). 21. Coutinho, A. R., Rocha, J. D. & Luengo, C. A. Preparing and characterizing biocarbon electrodes. Fuel Process. Technol. 67, 93 102 (2000). 22. Deprez, N. & McLachlan, D. S. The analysis of the electrical conductivity of graphite powders during compaction. J. Phys. D: Appl. Phys. 21, 101 107 (1988). 23. Gregg, S. J. & Sing, K. S. W. Adsorption, Surface Area and Porosity (Academic Press, 1982). 24. Boehm, H. P., Clauss, A., Fisher, G. O. & Hofmann, U. Das adsorptionsverhalten sehr dunner kohlenstoff-folien. Z. Anorg. Allg. Chem. 316, 119 127 (1962).

Received 24 July 2008; accepted 30 October 2008; published online 7 December 2008 References
1. Novoselov, K. S. et al. Electric eld effect in atomically thin carbon lms. Science 306, 666 669 (2004). 2. Li, X. et al. Highly conducting graphene sheets and Langmuir Blodgett lms. Nature Nanotech. 3, 538 542 (2008). 3. Hernandez, Y. et al. High-yield production of graphene by liquid-phase exfoliation of graphite. Nature Nanotech. 3, 563 568 (2008). 4. Li, X., Wang, X., Zhang, L., Lee, S. & Dai, H. Chemically derived, ultrasmooth graphene nanoribbon semiconductors. Science 319, 1229 1232 (2008). 5. Stankovich, S. et al. Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide. Carbon 45, 1558 1565 (2007). 6. Stankovich, S. et al. Graphene-based composite materials. Nature 442, 282 286 (2006).

Acknowledgements
This research was supported by the University of New South Wales (UNSW), the Faculty Research Grant Program (FRGP), Goldstar grants and an Australian Postgraduate Award (APA) to M.C. AFM measurements were carried out courtesy of P.T. at the University of Sydney.

Author contributions
J.A.S. and M.C. conceived and designed the experiments and developed the interpretation of results. M.C. synthesized all samples. P.T. assisted in the measurement and interpretation of the AFM data. All authors discussed the results and commented on the manuscript.

Additional information
Supplementary Information accompanies this paper at www.nature.com/ naturenanotechnology. Reprints and permission information is available online at http://npg. nature.com/reprintsandpermissions/. Correspondence and requests for materials should be addressed to J.A.S.

NATURE NANOTECHNOLOGY | VOL 4 | JANUARY 2009 | www.nature.com/naturenanotechnology 2009 Macmillan Publishers Limited. All rights reserved.

33

Das könnte Ihnen auch gefallen