Gold nanoparticle decorated reduced graphene oxide sheets with high catalytic activity for Ullmann homocoupling

Siyavash Kazemi Movahed, Mahsa Fakharian, Minoo Dabiri* and Ayoob Bazgir*
Shahid Beheshti University, Tehran, Islamic Republic of Iran. E-mail: a_bazgir@sbu.ac.ir; Fax: +98 21 22431661; Tel: +98 21 2990 3104

Received 1st October 2013 , Accepted 28th October 2013

First published on 29th October 2013


Abstract

A Au nanoparticle-reduced graphene oxide nanocomposite (Au NPs–RGO) was used as an effective and reusable heterogeneous catalyst for the Ullmann homocoupling of aryl iodides with short reaction times and good yields. For the first time, aryl bromides bearing electron-withdrawing groups were coupled in the presence of a catalyst giving moderate yields.


Supported Au nanoparticles (Au NPs) and their catalytic applications in heterogeneous organic reactions have attracted considerable attention in recent years. Aldehyde–alkyne–amine coupling (A3-coupling),1 oxidation of alcohols,2 secondary amines3 and carbon monoxide,4 hydrogenation of olefins,5 Suzuki6 and Sonogashira7 reactions, formation of nitrogen containing compounds,8 etc. are some of the most important applications of supported Au NP catalysts. The catalytic performance of Au NPs–support strongly depends on the size and shape of the Au NPs, the nature of the support, and the Au NPs-support interface interaction.9 Supporting carriers may function by dispersing and fixing the Au NPs. Supports such as oxides (CeO2,10 SiO2,11 Al2O3,12 MgO13 and TiO2 (ref. 14)), mixed oxides (Mg–Al–O15 and Ga–Al–O16), polymers (PVP17 and PS derivatives18) and ordered mesoporous carbon19 have been used as supports for Au NPs. Among these, it has been reported that the π-interaction of aromatic rings with gold nanoparticles leads to nanoparticle stabilization and improves the catalyst performance in a variety of gold-catalyzed reactions.9,18,20

Graphene is a two-dimensional sheet of sp2 bonded carbon atoms, which can be viewed as an extra-large polycyclic aromatic molecule.21 Recently, graphene has been used as the support for metal and metal oxide that is mainly due to its large surface area, excellent electrical and thermal conductivity, low price, high chemical inertness, ease of modification and strong interactions with metal clusters.22 However, only a few reports have involved the application of graphene based nanocomposites as heterogeneous catalysts in organic transformations.23 We therefore reasoned that graphene as a polycyclic aromatic molecule is a potential candidate as both a support and a stabilizer of Au NPs for chemical transformations.

Aryl–aryl bond formation is one of the most important tools of modern organic synthesis. These bonds are very often found in commercial dyes, natural products and organic conductors or semiconductors as well as being the backbone of some ligands used for asymmetric catalysis. The Ullmann type reaction, which is predominately catalyzed by copper, nickel and palladium catalysts, is a powerful and convenient synthetic method in organic chemistry for the aryl–aryl bond formation.24 However, only a few studies have involved the application of free or supported Au NPs as catalysts in the Ullmann type reaction of aryl iodides.20,25–27 Aryl bromides have not been used in gold catalyzed Ullmann homocoupling because the oxidative addition is more difficult relative to that of aryl iodides.

Herein, we report a novel protocol for the Ullmann homocoupling of aryl halides catalysed by Au NPs–RGO. Additionally, the effects of solvent polarity, base, and temperature on yield, and recycling potential of the catalyst have all been assessed.

Graphene oxide (GO) was prepared using a modified Hummers method,28 and subsequent reduction using NaBH4.29 Gold nanoparticles were deposited on the surface of the reduced graphene oxide (RGO) through spontaneous chemical reduction of HAuCl4 by RGO30 (Scheme 1). The mechanism of the spontaneous reductive deposition process that generates Au NPs on the RGO sheets likely involves a galvanic displacement and redox reaction because of the relative potential difference between RGO and HAuCl4 (ref. 30) (Scheme 1). In addition, a similar catalyst was prepared by an alternative route involving the physical mixture of Au NPs from the Turkevich–Frens method31 with RGO. This would enable a direct comparison of the two methodologies upon the catalytic activity.


image file: c3ra45518a-s1.tif
Scheme 1 Schematic route for preparation of Au NPs–RGO.

The Au NPs–RGO nanocomposite was characterized by FT-IR, XRD, TEM, Raman, XPS and EDS measurements. The electronic properties of the nanocomposite were probed by X-ray photoelectron spectroscopy (XPS). The XPS spectrum of Au 4f core for the region of the Au NPs–RGO level displays main peaks at 83.8 and 87.4[thin space (1/6-em)]eV, which corresponds to the binding energies of Au 4f7/2 and Au 4f5/2, respectively, and was shifted to the lower binding energy compared with the characteristic peaks for metallic Au0 at 84.0 and 87.7[thin space (1/6-em)]eV (Fig. 1). The negative shift (0.3[thin space (1/6-em)]eV) arises from the electron transfer from the graphene sheet to the Au NPs.32 The XPS spectrum of the Au 4f core for the region of the physical mixture of the Au NPs and RGO level displays its main peaks at 83.9 and 87.6[thin space (1/6-em)]eV, which correspond to the binding energies of Au 4f7/2 and Au 4f5/2, respectively, and were shifted to the higher binding energies compared with the characteristic peaks for Au NPs–RGO. Sodium citrate used as the reducing and capping agent for synthesis of Au NPs in Turkevich–Frens method inhibits the electron transfer from the graphene sheet to Au NPs, which causes a low shift (0.1[thin space (1/6-em)]eV).


image file: c3ra45518a-f1.tif
Fig. 1 XPS spectra of the Au 4f core level region for Au NPs–RGO and the physical mixture of Au NPs and RGO nanocomposites.

Fig. 2 shows the Raman spectra of GO, RGO, Au NPs–RGO and the physical mixture of Au NPs and RGO. In the Raman spectra of GO, RGO and Au NPs–RGO two similar fundamental peaks about 1308 and 1590[thin space (1/6-em)]cm−1 were observed, which correspond to the disorder induced by the D and G bands of the carbon atoms, respectively. The I(D)/I(G) intensity ratio of RGO (1.48) was slightly higher than that of the GO (1.40). This result may reflect the reduction in size of the graphene.33 The I(D)/I(G) intensity ratio of Au NPs–RGO (1.77) and the physical mixture of Au NPs and RGO (1.49) was much larger than that of the RGO. Such an enhancement has also been observed for metal nanoparticle composites of GO, indicating a probable chemical interaction or bond between the metal nanoparticles and graphene.34 The I(D)/I(G) intensity ratio for the physical mixture of Au NPs and RGO was lower than that of Au NPs–RGO, which indicates the presence of a weak chemical interaction or bond between the metal nanoparticles and graphene. Additionally, after conjugation of the Au NPs onto the RGO sheets, the intensity of the Raman signals of the nanocomposite was enhanced, which was due to the surface enhanced Raman spectroscopy (SERS) of Au NPs. SERS can be via an electromagnetic enhancement (excitation of localized surface plasmons involving a physical interaction) or chemical enhancement (formation of charge-transfer complexes involving chemical interaction) with enhancement factors of ∼1012 and ∼10 to 100, respectively. The low enhancement factor for the Au NPs–RGO nanocomposite indicates the presence of a chemical interaction or bond between Au NPs and RGO.35 The enhancement factor for the physical mixture of Au NPs with RGO was lower than that of Au NPs–RGO, which indicates the presence of a week chemical interaction or bond between the Au NPs and RGO.


image file: c3ra45518a-f2.tif
Fig. 2 Raman spectra of GO, RGO, Au NPs–RGO and physical mixture of Au NPs and RGO nanocomposites.

The morphology of GO and the Au NPs–RGO nanocomposite was determined by transmission electron microscopy (TEM). A TEM image of the GO displayed a crumpled and layer-like structure (Fig. 3a). As Fig. 3b shows, the surfaces of graphene are covered by distributed Au NPs with an average size of 10–20[thin space (1/6-em)]nm. Additionally, Au nanoparticles are not found outside of the RGO sheets. The chemical composition of the Au NPs–RGO nanocomposite was determined using energy dispersive spectroscopy (EDS) analysis. The EDS analysis in Fig. 3c clearly shows the presence of gold in the composite and the sample consists mainly of carbon with an insignificant amount of oxygen probably due to the presence of some unreduced oxygen functional groups. Additionally, the absence of the chlorine in the EDS analysis of the Au NPs–RGO composite corroborates the reduction of HAuCl4 by RGO.


image file: c3ra45518a-f3.tif
Fig. 3 (a) TEM image of GO (b) TEM image of Au NPs–RGO nanocomposite (c) EDS result for Au NPs–RGO nanocomposite.

The catalytic activity of the Au NPs–RGO catalyst was then tested in the Ullmann homocoupling reaction. To optimize the reaction conditions, iodobenzene was selected as a model substrate in the presence of various bases and solvents (Table 1). By using DMSO as the solvent, the reaction was tested employing various bases such as K3PO4, Cs2CO3, K2CO3, and KOH (entries 1–4). A superior yield was obtained when K3PO4 was used as the base (entry 1). Then, different solvents were screened in the model reaction using K3PO4 as the base. It was found that the reaction using NMP after 6[thin space (1/6-em)]h resulted in a higher yield. To optimize the reaction temperature, we also performed several experiments at 90[thin space (1/6-em)]°C, 100[thin space (1/6-em)]°C and 110[thin space (1/6-em)]°C in the presence of K3PO4 in NMP using 1.0[thin space (1/6-em)]mol% Au NPs–RGO (Table 1, entries 5, 9 and 10). As can be seen from Table 1, the most suitable reaction temperature was 100[thin space (1/6-em)]°C (entry 5). The effect of the Au NPs–RGO loading was also investigated under the optimum reaction conditions (Table 1, entries 5, 11 and 12). It was found that 1[thin space (1/6-em)]mol% Au was sufficient to push this reaction forward (entry 5). Additionally, the catalytic activity of the physical mixture of Au NPs with the RGO catalyst was tested in the Ullmann homocoupling reaction and afforded a relatively poor yield of the corresponding coupled product (Table 1, entry 13).

Table 1 Screening of the reaction conditionsa

image file: c3ra45518a-u1.tif

Entry Solvent Base Temp. (°C) Yield (%)
a Iodobenzene (1.0[thin space (1/6-em)]mmol), base (3.0[thin space (1/6-em)]mmol), Au NPs–RGO (1.0[thin space (1/6-em)]mol% Au), 6[thin space (1/6-em)]h and solvent (2[thin space (1/6-em)]ml). GC yield, n-dodecane was used as an internal standard.b Au NPs–RGO (0.5[thin space (1/6-em)]mol% Au).c Catalyst (1.5[thin space (1/6-em)]mol% Au).d Physical mixture Au NPs with RGO (1[thin space (1/6-em)]mol% Au).
1 DMSO K3PO4 100 94
2 DMSO Cs2CO3 100 85
3 DMSO K2CO3 100 67
4 DMSO KOH 100 39
5 NMP K3PO4 100 97
6 DMF K3PO4 100 72
7 PhCH3 K3PO4 100 15
8 CH3CN K3PO4 Reflux 56
9 NMP K3PO4 110 97
10 NMP K3PO4 90 74
11b NMP K3PO4 100 68
12c NMP K3PO4 100 97
13d NMP K3PO4 100 48


After the best reaction conditions were set, various aryl halides were screened to explore the scope of the Ullmann homocoupling reaction. As shown in Table 2, the aryl iodides bearing electron-donating and electron-withdrawing groups reacted well and gave good yields (Table 2, entries 1–4). The aryl iodides possessing electron-donating groups (p-OMe and p-Me) led to higher yields compared to the aryl iodide possessing an electron-withdrawing group (p-COCH3) (Table 2, entries 2–4). The hindered 2-iodotoluene substrate converted to the corresponding homocoupling product with a lower yield (Table 2, entry 5). Under the same reaction conditions the homocoupling of bromobenzene and aryl bromide bearing an electron-donating group (p-Me) failed to provide the target product (Table 2, entries 6–7). The aryl bromides bearing electron-withdrawing groups (p-COCH3 and p-CHO) were coupled in the presence of catalyst giving moderate yields (Table 2, entries 8–9).

Table 2 Au NPs–RGO nanocomposite catalyzed Ullmann homocoupling

image file: c3ra45518a-u2.tif

Entry Aryl halide Product Yielda
a Isolated yields.
1 image file: c3ra45518a-u3.tif image file: c3ra45518a-u4.tif 97%
2 image file: c3ra45518a-u5.tif image file: c3ra45518a-u6.tif 94%
3 image file: c3ra45518a-u7.tif image file: c3ra45518a-u8.tif 88%
4 image file: c3ra45518a-u9.tif image file: c3ra45518a-u10.tif 81%
5 image file: c3ra45518a-u11.tif image file: c3ra45518a-u12.tif 80%
6 image file: c3ra45518a-u13.tif 2a Trace
7 image file: c3ra45518a-u14.tif 2c Trace
8 image file: c3ra45518a-u15.tif 2d 68%
9 image file: c3ra45518a-u16.tif image file: c3ra45518a-u17.tif 41%


We proposed the following four-step mechanism for the Ullmann homocoupling reaction of an aryl halide catalyzed by the Au NPs–RGO nanocomposite (Scheme 2). First, the aryl halide was absorbed on the surface of the Au NP on the reduced graphene oxide sheet to form a (ArX) fragment. The C–X bond activation follow by the dissociation of C–X bond by the neutral Au0 atoms, gave rise to [Au NP Ar] [Au NP X] fragments.25,36 The homocoupling of [Au NP Ar]2 fragments generated the final product. Also, the spontaneous chemical reduction of Au+ by RGO regenerated the Au0–RGO nanocomposite. We assumed that the electron transfer from graphene sheets to Au nanoparticles caused an increase in the reactivity of the Au NPs–RGO nanocomposite toward the Ullmann homocoupling of aryl bromides containing electron withdrawing groups.


image file: c3ra45518a-s2.tif
Scheme 2 Proposed mechanism for the Ullmann homocoupling reaction of an aryl halide catalyzed by Au NPs–RGO nanocomposite.

The recyclability of the Au NPs–RGO nanocomposite was also examined by the Ullmann homocoupling of iodobenzene. It was found that the recovery can be successfully achieved in six successive reaction runs (Table 3).

Table 3 Reusability of the Au NPs–RGO nanocomposite in the homocoupling reaction of iodobenzenea
a Iodobenzene (1.0[thin space (1/6-em)]mmol), K3PO4 (3.0[thin space (1/6-em)]mmol), Au NPs–RGO (1.0[thin space (1/6-em)]mol% Au), 6[thin space (1/6-em)]h and NMP (2[thin space (1/6-em)]ml).b GC yield, n-dodecane was used as an internal standard.
Reaction cycle 1st 2nd 3rd 4th 5th 6th
Yieldb (%) 97 94 92 88 85 83


Table 4, compares the efficiency of the Au NPs–RGO nanocomposite (time, yield, reaction conditions) with the efficiency of other reported heterogeneous gold nanoparticle catalysts in the Ullmann homocoupling reaction of aryl iodides. It is clear from Table 4 that our method is simpler, more efficient, and less time consuming for the Ullmann homocoupling reaction.

Table 4 Comparison of the efficiency of various gold nanoparticle catalysts in the Ullmann homocoupling reaction of aryl iodides
Catalyst Condition Yield Time Ref.
Au NPs–RGO K3PO4, NMP, 100[thin space (1/6-em)]°C 80–97 (%) 6[thin space (1/6-em)]h This work
Au@PMO K3PO4, NMP, 100[thin space (1/6-em)]°C 80–95 (%) 16[thin space (1/6-em)]h 20
Au NP H2O/TBAOH, glucose, 100[thin space (1/6-em)]°C 45–99 (%) 7–15[thin space (1/6-em)]h 25
Au25(SR)18/CeO2 K2CO3, DMF, 130[thin space (1/6-em)]°C 67.5–99.8 (%) 48[thin space (1/6-em)]h 26
NAP–Mg–Au K2CO3 DMF, 140[thin space (1/6-em)]°C 45–92 (%) 48[thin space (1/6-em)]h 27


Conclusions

In conclusion, we have demonstrated that the Au NPs–RGO nanocomposite is an effective and reusable heterogeneous catalyst for the Ullmann homocoupling reaction with a short reaction time leading to good yields. The catalyst was recovered by simple filtration and reused for several cycles without a significant loss of catalytic activity. In this paper for the first time, aryl bromides bearing electron-withdrawing groups (p-COCH3 and p-CHO) were coupled in the presence of catalyst giving moderate yields.

Notes and references

  1. K. K. R. Datta, B. V. S. Reddy, K. Ariga and A. Vinu, Angew. Chem., Int. Ed., 2010, 49, 5961 CrossRef CAS PubMed.
  2. C. Marsden, E. Taarning, D. Hansen, L. Johansen, S. K. Klitgaard, K. Egeblad and C. H. Christensen, Green Chem., 2008, 10, 168 RSC.
  3. B. Zhu, M. Lazar, B. G. Trewyn and R. J. Angelici, J. Catal., 2008, 260, 1 CrossRef CAS PubMed.
  4. Y. Taia, J. Murakamia, K. Tajirib, F. Ohashib, M. Datéc and S. Tsubotac, Appl. Catal., A, 2004, 268, 183 CrossRef PubMed.
  5. P. Claus, Appl. Catal., A, 2005, 291, 222 CrossRef CAS PubMed.
  6. P. Garcia, M. Malacria, C. Aubert, V. Gandon and L. Fensterbank, ChemCatChem, 2010, 2, 493 CrossRef CAS; N. Zhang, H. Qiu, Y. Liu, W. Wang, Y. Li, X. Wang and J. Gao, J. Mater. Chem., 2011, 21, 11080 RSC.
  7. G. Li, D.-e. Jiang, C. Liu, C. Yu and R. Jin, J. Catal., 2013, 306, 177 CrossRef CAS PubMed.
  8. J. Mielby, S. Kegnæs and P. Fristrup, ChemCatChem, 2012, 4, 1037–1047 CrossRef CAS.
  9. A. Corma and H. Garcia, Chem. Soc. Rev., 2008, 37, 2096 RSC.
  10. S. Carrettin, P. Concepción, A. Corma, J. M. López Nieto and V. F. Puntes, Angew. Chem., Int. Ed., 2004, 43, 2538 CrossRef CAS PubMed.
  11. R. Resch, S. Meltzer, T. Vallant, H. Hoffmann, B. E. Koel, A. Madhukar, A. A. G. Requicha and P. Will, Langmuir, 2001, 17, 5666 CrossRef CAS; G. Rizza, Y. Ramjauny, T. Gacoin, L. Vieille and S. Henry, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 245414 CrossRef.
  12. E. Bus, J. T. Miller and J. A. van Bokhoven, J. Phys. Chem. B, 2005, 109, 14581 CrossRef CAS PubMed; Y.-F. Han, Z. Zhong, K. Ramesh, F. Chen and L. Chen, J. Phys. Chem. C, 2007, 111, 3163 Search PubMed.
  13. C.-J. Jia, Y. Liu, H. Bongard and F. Schüth, J. Am. Chem. Soc., 2010, 132, 1520 CrossRef CAS PubMed; N. S. Patil, B. S. Uphade, P. Jana, S. K. Bharagav and V. R. Choudhary, J. Catal., 2004, 223, 236 CrossRef PubMed.
  14. A. Dawson and P. V. Kamat, J. Phys. Chem. B, 2001, 105, 960 CrossRef CAS; R. Zanella, S. Giorgio, C. R. Henry and C. Louis, J. Phys. Chem. B, 2002, 106, 7634 CrossRef; X. Wang and R. A. Caruso, J. Mater. Chem., 2011, 21, 20 RSC.
  15. W. Fang, J. Chen, Q. Zhang, W. Deng and Y. Wang, Chem.–Eur. J., 2011, 17, 1247 CrossRef CAS PubMed.
  16. F. Z. Su, Y. M. Liu, L. C. Wang, Y. Cao, H. Y. He and K. N. Fan, Angew. Chem., 2008, 120, 340 CrossRef.
  17. H. Tsunoyama, H. Sakurai, Y. Negishi and T. Tsukuda, J. Am. Chem. Soc., 2005, 127, 9374 CrossRef CAS PubMed.
  18. H. Miyamura, R. Matsubara, Y. Miyazaki and S. Kobayashi, Angew. Chem., Int. Ed., 2007, 46, 4151 CrossRef CAS PubMed.
  19. S. Wang, Q. Zhao, H. Wei, J.-Q. Wang, M. Cho, H. S. Cho, O. Terasaki and Y. Wan, J. Am. Chem. Soc., 2013, 135, 11849 CrossRef CAS PubMed.
  20. B. Karimi and F. K. Esfahani, Chem. Commun., 2011, 47, 10452 RSC.
  21. M. Pumera, Chem. Soc. Rev., 2010, 39, 4146 RSC.
  22. X. Huang, X. Qi, F. Boey and H. Zhang, Chem. Soc. Rev., 2012, 41, 666 RSC.
  23. S. Bai and X. Shen, RSC Adv., 2012, 2, 64 RSC.
  24. J. Hassan, M. Sevignon, C. Gozzi, E. Schulz and M. Lemaire, Chem. Rev., 2002, 102, 1359–1469 CrossRef CAS PubMed.
  25. A. Monopoli, P. Cotugno, G. Palazzo, N. Ditaranto, B. Mariano, N. Cioffi, F. Ciminale and A. Nacci, Adv. Synth. Catal., 2012, 354, 2777 CrossRef CAS.
  26. G. Li, C. Liu, Y. Lei and R. Jin, Chem. Commun., 2012, 48, 12005 RSC.
  27. K. Layek, H. Maheswaran and M. L. Kantam, Catal. Sci. Technol., 2013, 3, 1147 CAS.
  28. W. S. Hummers and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef CAS; N. I. Kovtyukhova, P. J. Ollivier, B. R. Martin, T. E. Mallouk, S. A. Chizhik, E. V. Buzaneva and A. D. Gorchinskiy, Chem. Mater., 1999, 11, 771 CrossRef.
  29. S. Pei and H.-M. Cheng, Carbon, 2012, 50, 3210 CrossRef CAS PubMed.
  30. B.-S. Kong, J. Geng and H.-T. Jung, Chem. Commun., 2009, 2174 RSC.
  31. J. Turkevich, P. C. Stevenson and J. Hillier, Discuss. Faraday Soc., 1951, 11, 55 RSC; G. Frens, Nature Phys. Sci., 1973, 241, 20 CrossRef CAS.
  32. J. Li, C.-Y. Liu and Y. Liu, J. Mater. Chem., 2012, 22, 8426–8430 RSC; Z. Xiong, L. L. Zhang, J. Ma and X. S. Zhao, Chem. Commun., 2010, 46, 6099 RSC.
  33. D. Yang, A. Velamakanni, G. Bozoklu, S. Park, M. Stoller, R. D. Piner, S. Stankovich, I. Jung, D. A. Field, C. A. V. Jr and R. S. Ruoff, Carbon, 2009, 47, 145 CrossRef CAS PubMed.
  34. K. Jasuja and V. Berry, ACS Nano, 2009, 3, 2358 CrossRef CAS PubMed; K. Jasuja, J. Linn, S. Melton and V. Berry, J. Phys. Chem. Lett., 2010, 1, 1853 CrossRef; Y. Li, X. Fan, J. Qi, J. Ji, S. Wang, G. Zhang and F. Zhang, Nano Res., 2010, 3, 429 CrossRef.
  35. G. Goncalves, P. A. A. P. Marques, C. M. Granadeiro, H. I. S. Nogueira, M. K. Singh and J. Gracio, Chem. Mater., 2009, 21, 4796 CrossRef CAS; A. Campion, J. E. Ivanecky and C. M. Child, J. Am. Chem. Soc., 1995, 117, 11807 CrossRef; H. Ko, S. Singamaneni and V. V. Tsukruk, Small, 2008, 4, 1576 CrossRef PubMed.
  36. S. K. Beaumont, G. Kyriakou and R. M. Lambert, J. Am. Chem. Soc., 2010, 132, 12246 CrossRef CAS PubMed; V. K. Kanuru, G. Kyriakou, S. K. Beaumont, A. C. Papageorgiou, D. J. Watson and R. M. Lambert, J. Am. Chem. Soc., 2010, 132, 8081 CrossRef PubMed; P. S. D. Robinson, G. N. Khairallah, G. da Silva, H. Lioe and R. A. J. O'Hair, Angew. Chem., 2012, 51, 3878 CrossRef; M. Boronat, D. Combita, P. Concepción, A. Corma, H. García, R. Juárez, S. Laursen and J. Dios López-Castro, J. Phys. Chem. C, 2012, 116, 24855 Search PubMed.

Footnote

Electronic supplementary information (ESI) available: FT-IR spectra of GO, RGO and Au NPs–RGO, XRD spectra of GO, RGO Au NPs–RGO and a physical mixture of Au NPs with RGO. See DOI: 10.1039/c3ra45518a

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.