You are on page 1of 8

Colloid Polym Sci (2013) 291:20612068 DOI 10.

1007/s00396-013-2943-8

ORIGINAL CONTRIBUTION

Sulfonated graphene oxide: the new and effective material for synthesis of polystyrene-based nanocomposites
Liyuan Zhang & Tiejun Shi & Shengli Wu & Haiou Zhou

Received: 22 May 2012 / Revised: 1 March 2013 / Accepted: 13 March 2013 / Published online: 26 March 2013 # Springer-Verlag Berlin Heidelberg 2013

Abstract The new sulfonated graphene oxide (S-GO) was prepared and firstly used as effective materials for the synthesis of polystyrene/graphene nanocomposites via Pickering emulsion polymerization. The functionalized, chemically modified GO nanosheets were obtained via facile covalent functionalization with a reactive surfactant, sulfanilic acid. It was found that Pickering emulsion could be formed by simple self-assembly method using the S-GO as a stabilizer (just need 1 wt% relative to the oil phase), which could be adsorbed at the oilwater interface to stabilize the emulsion effectively. After the Pickering emulsion polymerization of styrene, the polystyrene/S-GO nanocomposites were prepared successfully. It is noteworthy that the S-GO not only could be used as a highly effective surfactant for styrene monomers but also could be homogeneously dispersed and incorporated into the polymeric matrix. Keywords Pickering emulsion . Graphene oxide . Surfactant . Polystyrene . Polymerization

Introduction Graphite oxide sheet, now called graphene oxide (GO), has attracted a great deal of attention in recent years for the

Electronic supplementary material The online version of this article (doi:10.1007/s00396-013-2943-8) contains supplementary material, which is available to authorized users. L. Zhang : T. Shi (*) : S. Wu : H. Zhou School of Chemical Engineering, Hefei University of Technology, Hefei, Anhui 230009, Peoples Republic of China e-mail: hfutpolymer@163.com L. Zhang Department of Chemistry and Environmental Engineering, Bengbu College, Bengbu, Anhui 233000, Peoples Republic of China

production of graphene-based materials [14]. Due to excellent water dispersity, GO has long been considered hydrophilic since its discovery a century ago [58]. However, Huang et al. recently revealed that GO are actually amphiphilic with an edge-to-center distribution of hydrophilic and hydrophobic domains [911]. Furthermore, they revealed that GO nanosheets could be adsorbed at the oilwater interface and used as a stabilizer in Pickering emulsions (often referred to solid-stabilized emulsion). As a versatile approach for the design and production of nanostructured materials, the Pickering emulsions have received great interest in recent years because of their practical applications, such as in cosmetics, pharmaceutics, food, oil recovery, and waste water treatment [12]. Compared with conventional emulsifying agents, particulate emulsifiers provide many advantages: robustness, reusability, less toxicity, and lower cost [13]. The Pickering emulsion polymerization provides a simple and effective way to prepare inorganic/organic composite materials [14, 15]. From this point of view, a few examples of Pickering emulsion polymerization stabilized by GO have been reported recently, and a venue to the development of a broad range of graphene-based nanostructures was opened because of GOs novel characteristic [16, 17]. However, the production of polymer/graphene nanocomposites by Pickering emulsion polymerization has not been well investigated and needs a large number of GO usually or under specific circumstances [16]. When incorporated appropriately, graphene can significantly improve physical properties of matrix at extremely small loading [18]. However, the preparation of graphenebased nanocomposites is still a challenge because of the aggregation of bulk-quantity graphene sheets in the synthesis and processing [19]. On the other hand, the covalent functionalization of graphene has attracted increasing attention because it is an effective method in achieving homogeneous and stable dispersions of graphene [20, 21]. The oxygen-containing groups on the GO sheets, such as

2062

Colloid Polym Sci (2013) 291:20612068

epoxide, hydroxyl, and carboxylic acid groups, enable graphene to be further modified by chemical reactions, and many different functional groups can be introduced onto the graphene sheets [22]. For example, isocyanate-treated graphene oxide nanosheets, which can form a stable dispersion in polar aprotic solvents and be used to prepare conducting polystyrenegraphene nanocomposites, have been synthesized by Stankovich [23]. Niyogi and his co-workers prepared octadecylamido-graphite oxide, G-CONH(CH2)17CH3; by functionalization, graphite oxide with a long-chain alkylamine gives rise to stable solutions of this material in organic solvents [24]. 3-Aminopropyltriethoxysilanegraphene nanosheets, which can be homogeneously distributed in polar solvents, were obtained via facile covalent functionalization with 3aminopropyltriethoxysilane by Yang and his colleagues [19]. Si et al. functionalized graphene by treatment with aryl diazonium salts, for attachment of SO3H group onto the graphene, allows these nanosheets to be homogeneously dispersed in aqueous solution [25]. The covalent functionalization of graphene nanosheets not only improves dispersibility of graphene nanosheets in different media but also enhances interfacial interactions between the graphene and the matrix [19]. Herein, we report a new method for the synthesis of polystyrene/graphene nanocomposites via the Pickering emulsion polymerization using the novel functionalized, chemically modified graphene oxide nanosheets as stabilizer. First, the sulfonated graphene oxide (S-GO) nanosheets were prepared via facile covalent functionalization with sulfanilic acid. The functionalized GO was then used as an effective stabilizer (just need 1 % relative to the monomers) for preparation Pickering emulsions. After the Pickering emulsion polymerization of styrene, the polystyrene (PS)/S-GO nanocomposites were prepared successfully. To the best of our knowledge, this is the first report on the polymer/graphene nanocomposites originating from Pickering emulsion polymerization using GO nanosheets with reactive groups as stabilizer. It is noteworthy that the S-GO not only could be used as a highly effective surfactant for organic monomers but also could be homogeneously dispersed and incorporated into the polymeric matrix.

(NHS) were purchased from Aladdin Chemistry Co. Ltd (China). Styrene (St) and potassium persulfate (K2S2O8) were obtained from Shanghai Chemical Company (China). Before use, St was washed with the 5 % aqueous solution of sodium hydroxide and deionized water, dried over anhydrous sodium sulfate, and then distilled at reduced pressure. The other reagents were used as received without further purification. Preparation of GO GO was synthesized from graphite power using a modified Hummers method [26]. In a typical synthesis process, concentrated sulfuric acid (230 mL) was transferred to a 2-L Erlenmeyer flask and was cooled to 0 C by the ice bath. Graphite powder (10 g) and sodium nitrate (5 g) were then slowly added to the flask, with mechanical stirring. Potassium permanganate (30 g) was gradually added to the suspension. The temperature was not allowed to surpass 20 C. The reaction mixture was then stirred at 1015 C for 2 h. The mixture was then brought to 35 C and stirred for about 2 h, forming a thick paste. Subsequently, deionized water (230 mL) was slowly added to the reaction vessel and kept the temperature below 98 C. The diluted suspension was maintained at this temperature for 30 min and further diluted to 1.4 L with warm deionized water, followed by treatment with 30 % hydrogen peroxide (25 mL) to reduce the residual permanganate and manganese dioxide to colorless soluble manganese sulfate. After that, the resulting slurry was dialyzed against water for 7 days and sonicated using a KQ100E ultrasonic bath cleaner (100 W) for 1 h to exfoliate graphitic oxide, and unexfoliated graphitic oxide was removed by centrifugation (10,000 rpm, 5 min). The dry GO was obtained after having been dried in vacuum oven at 50 C for 24 h. Preparation of functionalized sulfonated graphene oxide The process of synthesizing sulfonated graphene oxide from graphene oxide was carried following a literature [22] with some modification. In a typical procedure, GO was treated with sulfanilic acid as follows: 200 mg graphene oxide was dispersed into 100 mL of dry DMF and then sonicated with the KQ-100E ultrasonic bath cleaner (100 W) for 2 h at room temperature to yield a clear solution. In this process, graphene oxide was completely exfoliated down to individual graphene sheets to form a stable dispersion of graphene oxide solution. NHS (0.68 g) and EDCHCl (1.15 g) were then added to the solution at room temperature. After stirring for 2 h, sulfanilic acid (0.76 g) was added, and the solution was stirred 24 h at room temperature to produce the functionalized S-GO. After that, the resulting slurry was dialyzed against water for 2 days, and the product was obtained after having been dried in vacuum oven at 50 C

Experimental Materials Graphite was supplied by Shandong Qingdao Company (China). Sulfuric acid (98 %), potassium permanganate, hydrogen peroxide (30 %), hydrochloric acid, and sulfanilic acid were purchased from Sinopharm Chemical Reagent Co. Ltd (China). N-(3-(dimethylamino)propyl)-N-ethylcarbodiimide hydrochloride (EDCHCl) and N -Hydroxysuccinimide

Colloid Polym Sci (2013) 291:20612068

2063

for 24 h. The elemental analysis of S-GO showed that the loading was readily quantified via S microanalysis and was 0.47 mmol g1 of SO3H based on the sulfur percentage (1.51 wt%). Preparation of Pickering emulsion stabilized by S-GO and PS/S-GO nanocomposites In a typical procedure, S-GO (15 mg) and 80 mL distilled water were loaded in a 250-mL round-bottomed flask, sonicated by the KQ-100E ultrasonic bath cleaner (100 W) until it became homogeneous with yellow-brown dispersion. St (1.5 g) and potassium persulfate (K2S2O8, 10 mg) were then added to the flask. After the mixture was sonicated for 10 min, a styrene-in-water milky emulsion stabilized by SGO was then prepared. The Pickering emulsion was flushed with nitrogen 20 min and then polymerized at 70 C for 4 h under stirring. The resulting slurry was dialyzed against water for 2 days, and the dried PS/S-GO nanocomposites were obtained after having been dried in vacuum oven at 50 C for 24 h. Preparation of Pickering emulsion stabilized by GO and PS/GO nanocomposites Pickering emulsion stabilized by GO and PS/GO nanocomposites were prepared with the same procedures as described above. In a typical procedure, GO (15 mg) and 80 mL distilled water were loaded in a 250-mL roundbottomed flask, sonicated by the ultrasonic bath cleaner until it became homogeneous with yellow-brown dispersion. St (1.5 g) and K2S2O8 (10 mg) were then added to the flask. After the mixture was sonicated and stirred for 30 min, a styrene-in-water milky emulsion stabilized by GO was then prepared. The Pickering emulsion was flushed with nitrogen 20 min and then polymerized at 70 C for 4 h under stirring. The resulting slurry was dialyzed against water for 2 days, and the dried PS/GO nanocomposites were obtained after having been dried in vacuum oven at 50 C for 24 h. Characterization and measurements Fourier transform infrared (FT-IR, Perkin Elmer Spectrum 100) spectra were recorded by using KBr as a background on a Nicolet 8700 FT-IR spectrometer. The samples for AFM measurements were prepared by depositing the aqueous dispersion on a freshly cleaved mica surface and dried at room temperature. X-ray photoelectron spectroscopy (XPS, ESCALAB 250) was used to characterize the structure of the GO and S-GO. Transmission electron microscope (TEM, JEOL-2010) and scanning electron microscopy (SEM, Sirion 200) were used to observe the morphology of the S-GO and nanocomposite samples. Optical micrographs

were obtained on an Olympus BX41 laboratory microscope. The images of the emulsions were obtained using a digital camera.

Results and discussion Covalent functionalization of GO with sulfanilic acid The functionalized, chemically converted graphene nanosheets (S-GO) can be synthesized via the covalent interaction between GO and sulfanilic acid. The amino group of sulfanilic acid can facilely interact with the epoxide and carboxyl groups on the GO nanosheet under the action of NHS and EDC. The similar reactions of amino group with the epoxide and carboxyl groups have been reported recently [19, 22]. Figure 1 shows the FT-IR spectra of GO (Fig. 1a) and its modified form (S-GO) (Fig. 1b). In Fig. 1a, the adsorption bands corresponding to the C=O carbonyl stretching at 1,720 cm1, the OH deformation vibration at 1,373 cm1, and the CO stretching at 1,049 cm1 [27, 28]. Besides the ubiquitous OH stretches which appear at 3,416 cm1 as a broad and intense signal, the resonance at 1,612 cm1 can be assigned to the vibrations of the adsorbed water molecules but may also contain components from the skeletal vibrations of un-oxidized graphitic domains [2830]. All of these confirmed that the graphite was oxided and functional groups were introduced onto the GO sheets [31]. Figure 1b shows the characteristic peaks of the S-GO. After sulfonation, the peaks at 1,720 and 1,049 cm1 are severely attenuated. The peaks at 1,173, 1,124, and 1,036 cm1 (two SO and one S-phenyl) confirm the presence of a sulfonic acid group, and the peaks at 1,008 cm1 (CH in-plane bending) and 838 cm1 (out-ofplane hydrogen wagging) are characteristic vibrations of a p-disubstituted phenyl group [25, 32].

Fig. 1 FT-IR spectra of GO (a) and S-GO (b)

2064

Colloid Polym Sci (2013) 291:20612068

Fig. 2 XPS results of GO (a) and S-GO (b)

The XPS spectrum provides information on the type and number of different species of a given atom in the molecules. The chemical compositions of GO and S-GO were characterized by XPS. Figure 2 shows wide XPS spectra of GO and SGO. For S-GO, there are peaks at 399.78 and 167.75 eV, respectively, corresponding to the N1s and S2p binding energy of sulfanilic acid chains (Fig. 2b). However, the XPS spectrum of GO does not show peaks at 399.78 and 167.75 eV (Fig. 2a), which implies that the surfaces of the flakes of S-GO have been successfully functionalized with sulfanilic acid. The higher resolution data of C1s area of the GO and SGO are shown in Fig. 3a b, respectively. The C1s XPS spectrum of GO showed binding energies at 284.6 eV (C C), 286.67 eV (CO), 287.91 eV (C=O), and 288.87 eV (O C=O) [33, 34]. Accompanied with the amount of C (epoxy/ether), GO dramatically decreases after chemically converted by sulfanilic acid, and there is an additional component at 286.0 eV assigned to C bound to nitrogen [1], strongly indicating that the amino moiety reacts with the epoxy groups on GO sheets. The peak at 288.87 eV (O C=O) turns into 288.79 eV (NC=O), which also proves the grafting of sulfanilic acid chains onto the surface of GO sheets. This is consistent with that in the FT-IR analysis.

Atomic force microscopy (AFM) is one of the most direct methods for quantifying the degree of exfoliation to GO sheet level [35]. The samples for AFM measurements were prepared by ultrasonic treatment of GO and S-GO (in water) dispersions of 0.05 mg/mL, respectively. The samples were prepared through drop-casting on freshly cleaved mica surfaces. The micas were dried at ambient conditions and directly examined using AFM. Individual GO nanosheet was imaged using AFM as shown in Fig. 4a. On average, the height of the GO nanosheet is 1.08 nm, indicating that the individual GO nanosheet was indeed achieved [1]. We also find that the size of GO is fairly polydispersed with lateral dimension ranging from nanometers to micrometers. Figure 4b reveals that exfoliated S-GO nanosheets with average thickness 1.85 nm have been obtained in our work and S-GO is expected to be thicker, owing to the presence of functionalized sulfanilic acid chains grafted on the GO nanosheets. Similar results have also been observed for the thickness of the well-exfoliated functionalized graphene sheets by AFM measurement [19, 36]. Huang et al. recently uncovered the surface activity of GO which may be applied in Pickering emulsions. However, using GO as a stabilizer in Pickering emulsion polymerization meets two main problem. First, the general lateral size of the GO sheets is in the range of tens of microns, which are large enough to endanger the stability of the resulting emulsion. The second is that the sizes of resulting emulsion particles range from few hundreds of micrometers to a few millimeters, which are several orders of magnitude higher than the typical colloidal particles, and thus, the large particles are prone to instability and precipitation [16]. These challenge the fabrication of polymer/graphene nanocomposites by utilizing these Pickering emulsions. As a matter of fact, Huang et al. used simple shaking for the emulsification which was insufficient to disperse the oil phase into fine particles [911]. The high power sonication method was used by Sharif et al. to employ GO for assembly at the liquidliquid interface for Pickering emulsion polymerization [16]. In addition, it is noteworthy that the styrene-in-water emulsions and the PS/GO nanocomposites were prepared also using a sonicator by Song and his

Fig. 3 C1s XPS spectra of GO (a) and S-GO (b)

Colloid Polym Sci (2013) 291:20612068 Fig. 4 AFM images and height profiles of GO (a) and S-GO nanosheets (b)

2065

colleagues [17], but we found that it could not be reproduced by only using a lower-power sonicator. Lack of stability has limited the use of GO as a stabilizer in Pickering emulsions. Specifically, GO needs modification to develop desirable functional properties such as adhesion and dispersion. In particular, GO needs to be modified to enhance its emulsifying capacity. Herein, efforts have been made to introduce a reactive surfactant, sulfonic acid, onto graphene oxide nanosheets in a controlled way. All the above results of FT-IR, XPS, and AFM indicated that the amino moiety of sulfanilic acid was successfully reacted with epoxide and carboxyl groups of the GO nanosheets, and the novel S-GO was prepared successfully. In this study, we have introduced a novel and more practical method for a stable Pickering emulsion stabilized by the novel, functionalized, and chemically converted graphene oxide nanosheets (S-GO) and the production of PS/S-GO nanocomposites. The Pickering emulsion stabilized by GO and S-GO To obtaining stable Pickering emulsions, the colloidal particles, solid surfactants, are required to adsorb at the oilwater interface [37]. Classical colloid science tells us that emulsification generally does not occur spontaneously, and some extra energy is required to initiate the emulsion formation [16]. Herein, we also employed sonication for the emulsification of the monomer through the aqueous phase to obtain droplets within the size of conventional emulsions.

Figure 5 shows the optical micrographs of emulsions stabilized by different concentrations of GO and S-GO. To screen the emulsifying capacity of S-GO, a systematic experiment has been done, and optical micrographs were taken to explore the formation of styrene droplets. The result of the experiment indicates that, in the absence of any stabilizing agent, the droplets of oil coalesced, and complete phase separation occurred after a while in the styrene-in-water emulsion. After the addition of the 1 wt% GO (relative to the oil phase),

Fig. 5 Optical micrographs of O/W emulsions stabilized by 1 wt% of GO (a), 1 wt% of S-GO (b), 5 wt% GO (c), and 5 wt% S-GO (d)

2066 Fig. 6 TEM images of S-GO (a) and PS/S-GO nanocomposites (b)

Colloid Polym Sci (2013) 291:20612068

oil droplets with sizes ranging from 1 to 7 m were observed (Fig. 5a). In fact, the formed emulsions were very unstable, and the emulsion droplets coalesced, and phase separation also occurred. Using the emulsion for polymerization, the dispersion of the PS/GO hybrid was also unstable and the coagulation at the early stage of polymerization and precipitation at the end. This can be explained by the fact that, in order to minimize the free energy of the system, the emulsion particles merge with each other and decrease the surface-to-volume ratio [38]. However, on the other hand, incorporating the 1 wt% S-GO (relative to the oil phase) into the solution led to the formation of a stable dispersion and remarkably decreases the initial size of oil droplets. Figure 5b shows the optical microscope images of emulsions where the ratio of S-GO to organic phase is 1 wt%. In the images, the sizes of styrene droplets are less than 1 m in diameter. This confirms the high surface activity of S-GO

sheets, which are able to reduce the interfacial tension more effectively than GO. Figure 5c, d shows two optical microscope images of emulsions where the ratio of GO and S-GO to organic phase is 5 wt%. When the ratio of GO to oil phase was up to 5 wt%, the stability of emulsions enhanced, and the particle size of the emulsion droplets notably decreased, reaching up to 13 m (Fig. 5c). However, on the other hand, at higher S-GO concentration, the particle size of the emulsion droplets almost unchanged (Fig. 5d). This further evidences that the S-GO (just need 1 wt% relative to the oil phase) could have assisted the Pickering emulsions more effectively than GO. Therefore, in the subsequent Pickering emulsion polymerization of styrene, the SGO concentration was kept as 1 wt% relative to the oil phase. Recently, using GO as surfactant to guarantee the stability of Pickering emulsion, Gudarzi et al. used more than 4 wt% of

Fig. 7 SEM images of PS/SGO nanocomposites (a, b) and PS/GO nanocomposites (c, d) at different magnifications

Colloid Polym Sci (2013) 291:20612068

2067

GO to form Pickering emulsion under specific circumstances, and Song et al. also reported a critical concentration of GO up to 4.5 wt% [16, 17]. In this study, the huge surface area of SGO (Fig. 6a) and the large number of functionalized sulfanilic acid groups on the nanosheets make it possible to have a low critical concentration, and the concentration of S-GO (1 wt%) for a stable dispersion was remarkably low compared with the previous reported works. In addition to the surface area of S-GO, the sizes of polymer particles and the extent of the coverage of polymer particles with S-GO nanosheet also affect the critical concentration of nanosheets for stabilization [16]. Therefore, the morphology of the resulting polymer/graphene nanocomposites was examined to study the arrangement of S-GO nanosheets and polymer particles. Nanocomposite morphology The morphologies of S-GO and PS/S-GO nanocomposites were characterized by TEM technique. As shown in Fig. 6a, the ultrathin platy sheet with rolled edges can be observed. The transparent nanosheet ranges from 200 nm to more than 2 m in dimension. With the aid of sulfanilic acid groups on the sheet, the S-GO nanosheets could stabilize PS particles with a relatively low concentration (1 wt%). Figure 6b shows images of PS/S-GO nanocomposites, and the PS particles dressed on the surface of the S-GO nanosheets were clearly observed. It was noticed that the PS particles with the average size of 200 nm in diameter could dress on the two sides of S-GO nanosheet surface. This can be explained by the fact that the two sides of basal planes of GO nanosheet have been chemically modified and functionalized with sulfanilic acid. When PS colloidal particles dress on the surface of S-GO nanosheet, the particles can share the S-GO nanosheet on two sides, and hence, the efficiency of the S-GO as a stabilizer is improved. To investigate the morphology of PS/S-GO nanocomposites, SEM measurements have also been performed. As shown in Fig. 7a, b, we can distinctly find that PS microspheres are firmly immobilized on the surface of S-GO nanosheets, which further confirm the successful decoration of S-GO by PS microspheres. The size of polymer particles observed in the micrograph is consistent with the results from the TEM. From the figure, we clearly found that the lateral size of S-GO was larger than that of the PS microspheres wherein highly crumbled and wrinkled nanosheets were observed. This may arise from the fact that larger S-GO nanosheet stabilized more than one microsphere which is not fully covered by the nanosheet. Therefore, upon drying the emulsion, the S-GO nanosheets crumble. A control experiment was conducted to prove that the SGO is more effective than GO as surfactant and filler for polymer composites. The Pickering emulsion stabilized by

GO and PS/GO nanocomposites were prepared with the same procedures, and SEM specimen was prepared (Fig. 7c, d). The SEM result indicated that the PS colloidal particles with different size are not homogeneously dispersed on the GO surface, and the diameters of PS colloidal particles range from 450 to 100 nm. Furthermore, it was noticed that the aggregation and restack of GO sheets can be clearly seen in this case. As we all know, aggregation of graphene may be very detrimental to composite properties and is one of the great challenges in the preparation of graphene-based nanocomposites [39]. Compared with GO, with aid of functionalized groups on the nanosheets, the S-GO could overcome the aggregation in the synthesis and could be homogeneously dispersed into the polymeric matrix finally. The S-GO showed to be more compatible and incorporated with the polystyrene matrix. Furthermore, the particle size distribution of the polystyrene is fairly narrow using S-GO as surfactant.

Conclusions In summary, the novel S-GO was prepared successfully and firstly used as effective materials for the synthesis of polystyrene/graphene nanocomposites via the Pickering emulsion polymerization. The functionalized chemically converted graphene oxide was facilely prepared via covalent functionalization with sulfanilic acid. The S-GO was found to be a highly effective surfactant for styrene monomers in aqueous media, resulting in an even formation of submicron polymer particles. It is noteworthy that the S-GO not only could be used as a highly effective surfactant for styrene monomers but also could be homogeneously dispersed and incorporated into the polymeric matrix.
Acknowledgments We gratefully acknowledge financial support by the National Natural Science Foundation of China (grant nos. 51273054 and 50973024) and the Natural Science Foundation of Anhui Provinces Higher Education of China (no. KJ2009B212Z).

References
1. Stankovich S, Dikin DA, Piner RD et al (2007) Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide. Carbon 45:15581565 2. Dreyer DR, Park S, Bielawski CW, Ruoff RS (2010) The chemistry of graphene oxide. Chem Soc Rev 39(1):228240 3. Compton OC, Nguyen ST (2010) Graphene oxide, highly reduced graphene oxide, and graphene: versatile building blocks for carbon-based materials. Small 6(6):711723 4. Zhu YW, Murali S, Cai WW, Li XS et al (2010) Graphene and graphene oxide: synthesis, properties, and applications. Adv Mater 22:39063924 5. Brodie BC (1860) Surle poids atomique du graphite. Ann Chim Phys 59:466472

2068 6. Paredes JI, Villar-Rodil S, Martinez-Alonso A, Tascon JMD (2008) Graphene oxide dispersions in organic solvents. Langmuir 24:1056010564 7. Marcano DC, Kosynkin DV, Berlin JM, Sinitskii A et al (2010) Improved synthesis of graphene oxide. Nano 4(8):48064814 8. Wang H, Hu YH (2011) Effect of oxygen content on structures of graphite oxides. Ind Eng Chem Res 50:61326137 9. Kim J, Cote LJ, Kim F, Yuan W, Shull KR, Huang JX (2010) Graphene oxide sheets at interfaces. J Am Chem Soc 132:81808186 10. Kim F, Cote LJ, Huang JX (2010) Graphene oxide: surface activity and two-dimensional assembly. Adv Mater 22(17):19541958 11. Cote LJ, Kim J, Tung VC, Luo JY, Kim F, Huang JX (2011) Graphene oxide as surfactant sheets. Pure Appl Chem 83(1):95110 12. Fujii S, Okada MSH, Furuzono T, Nakamura Y (2009) Hydroxyapatite nanoparticles as particulate emulsifier: fabrication of hydroxyapatite-coated biodegradable microspheres. Langmuir 25(10):97599766 13. Yang F, Liu S, Xu J, Lan Q, Wei F, Sun DJ (2006) Pickering emulsions stabilized solely by layered double hydroxides particles: the effect of salt on emulsion formation and stability. J Colloid Interface Sci 302:159169 14. Chen JH, Cheng CY, Chiu WY, Lee CF, Liang NY (2008) Synthesis of ZnO/polystyrene composites particles by Pickering emulsion polymerization. Eur Polym J 44(10):32713279 15. Zhang K, Wu W, Meng H, Guo K, Chen JF (2009) Pickering emulsion polymerization: preparation of polystyrene/nano-SiO2 composite microspheres with core-shell structure. Powder Technol 190(3):393400 16. Gudarzi MM, Sharif F (2011) Self assembly of graphene oxide at the liquidliquid interface: a new route to the fabrication of graphene based composites. Soft Matter 7:34323440 17. Song XH, Yang YF, Liu JC, Zhao HY (2011) PS colloidal particles stabilized by graphene oxide. Langmuir 27:11861191 18. Hyunwoo K, Ahmed AA, Christopher WM (2010) Graphene/ polymer nanocomposites. Macromolecules 43(16):65156530 19. Yang HF, Li FH, Shan CS et al (2009) Covalent functionalization of chemically converted graphene sheets via silane and its reinforcement. J Mater Chem 19:46324638 20. Stankovich S, Dikin DA, Dommett GHB et al (2006) Graphenebased composite materials. Nature 442:282286 21. Liang JJ, Xu YF, Huang Y et al (2009) Infrared-triggered actuators from graphene-based nanocomposites. J Phys Chem C 113:99219927 22. Yang YF, Wang J, Zhang J, Liu J, Yang XL, Zhao HY (2009) Exfoliated graphite oxide decorated by PDMAEMA chains and polymer particles. Langmuir 25(19):1180811814

Colloid Polym Sci (2013) 291:20612068 23. Stankovich S, Piner RD, Nguyen ST, Ruoff RS (2006) Synthesis and exfoliation of isocyanate-treated graphene oxide nanoplatelets. Carbon 44:33423347 24. Niyogi S, Bekyarova E, Itkis ME et al (2006) Solution properties of graphite and graphene. J Am Chem Soc 128(24):77207721 25. Si YC, Samulski ET (2008) Synthesis of water soluble graphene. Nano Lett 8(6):16791682 26. Hummers WS, Offeman RE (1958) Preparation of graphitic oxide. J Am Chem Soc 80(6):1339 27. Hontoria LC, Lopez PAJ, Lopez GJD, Rojas CM, Martin ARM (1995) Study of oxygen-containing groups in a series of graphite oxides: physical and chemical characterization. Carbon 33(11):15851592 28. Szabo T, Berkesi O, Dekany I (2005) DRIFT study of deuteriumexchanged graphite oxide. Carbon 43(15):31863189 29. Mermoux M, Chabre Y, Rousseau A (1991) FTIR and 13C NMR study of graphite oxide. Carbon 29(3):469474 30. Cataldo F (2003) Structural analogies and differences between graphite oxide and C60 and C70 polymeric oxides (fullerene ozopolymers). Fullerenes Nanotubes Carbon Nanostruct 11(1):1 13 31. Lerf A, He HY, Forster M, Klinowski J (1998) Structure of graphite oxide revisited. J Phys Chem B 102:44774482 32. Colthup NB, Daly LH, Wiberley SE (1990) Introduction to infrared and Raman spectroscopy. Academic, London 33. Liu JB, Fu SH, Yuan B, Li YL, Deng ZX (2010) Toward a universal adhesive nanosheet for the assembly of multiple nanoparticles based on a protein-induced reduction/decoration of graphene oxide. J Am Chem Soc 132(21):72797281 34. Yoon SH, Park JH, Kim EY, Kim BK (2011) Preparations and properties of waterborne polyurethane/allyl isocyanated-modified graphene oxide nanocomposites. Colloid Polym Sci 289:1809 1814 35. Shen JF, Hu YZ, Li C, Qin C, Ye MX (2009) Synthesis of amphiphilic graphene nanoplatelets. Small 5(1):8285 36. Xu YX, Bai H, Lu GW, Li C, Shi GQ (2008) Flexible graphene films via the filtration of water-soluble noncovalent functionalized graphene sheets. J Am Chem Soc 130(18):58565857 37. Binks BP (2002) Particles as surfactants-similarities and differences. Curr Opin Colloid Interface Sci 7:2141 38. Colver PJ, Colard CAL, Bon SAF (2008) Multi-layered nanocomposite polymer colloids using emulsion polymerization stabilized by solid particles. J Am Chem Soc 130(50):16850 16851 39. Potts JR, Dreyer DR, Bielawski CW, Ruoff RS (2011) Graphenebased polymer nanocomposites. Polymer 52(1):525

You might also like