Open Access Article
Prasanna Kumar S
Mural†
a,
Maya
Sharma†
a,
Giridhar
Madras
b and
Suryasarathi
Bose
*c
aCenter for Nano Science and Engineering, Indian Institute of Science, Bangalore-560012, India
bDepartment of Chemical Engineering, Indian Institute of Science, Bangalore-560012, India
cDepartment of Materials Engineering, Indian Institute of Science, Bangalore-560012, India. E-mail: sbose@materials.iisc.ernet.in
First published on 30th March 2015
Graphene oxide (GO), prepared by chemical oxidation of graphite, serves as a building block for developing polymeric nanocomposites. However, their application in electrical conductivity is limited by the fact that the oxygen sites on GO trap electrons and impede charge transport. Conducting nanocomposites can be developed by reducing GO. Various strategies have been adopted to either reduce GO ex situ, before the composite preparation, or in situ during the development of the nanocomposites. The current state of research on in situ reduction of GO during the preparation of conducting polymeric nanocomposites is discussed in this review. The mechanism and the efficiency of reduction is discussed with respect to various strategies employed during the preparation of the nanocomposite, the type of polymer used, and the processing conditions employed, etc. Its overall effect on the electrical conductivity of the nanocomposites is also discussed and the future outlook in this area is presented.
000 cm2 V−1 s−1 of charge carrier mobility, high Young's modulus of ∼1100 GPa and extraordinary fracture strength of 125 GPa. Apart from high electrical conductivity, they possess very high specific surface area of 2630 m2 g−1 and interesting transport phenomena such as the quantum Hall effect.1,2 Besides flexibility in properties, mono layer graphene sheets offers myriad possibilities of chemical modification and functionalization.1,3 Thus, graphene can be potentially used as energy-storage materials, polymer composites and sensors etc. Yan et al.4 reported that graphene and few-layer graphene can be utilized for thermal management of advanced electronics wherein graphene acts as heat spreaders in transistors. Luo et al.,5 summarized various graphene based applications in energy storage such as solar cells, lithium ion secondary batteries and supercapacitors. In a recent article, Mittal et al.,6 reported that graphene exhibits restacking of sheets due van der Waals forces and strong π–π stacking effects. The interaction between graphene sheets and polymer matrix plays a vital role in achieving high electrical percolation. Graphene sheets can also act as a cationic initiator for polymerization of polystyrene (PS) and poly(styrene-isoprene).7 Graphene polymer composites can be used for tissue engineering applications such as osteogenesis in 3D scaffolds,8,9 selective gas/vapor sensors10 and in removal of heavy metal/pathogenic bacteria from aqueous media.11–14 GO based composites showed a significant surface enhanced Raman scattering (SERS),15 which can be useful in developing ultrasensitive SERS-based immunosensing platforms.16–18
Graphite to individual 2D graphene conversion is widely carried out by chemical oxidation of graphite layer. The formed oxides serve as a precursor for cost-effective and mass production of graphene-based materials.2 These exfoliated layers of graphite sheets with oxygen moieties are referred to as graphene oxide (GO). GO offers flexibility in the large scale production of graphene based nanocomposite.19 GO is widely synthesized via Brodie,20 Staudenmaier21 and Hummers methods.17,22 All these methods focus on the oxidation of individual layers using strong acids and oxidants for the introduction of oxide groups such as epoxy, hydroxyl or carboxyl.17,23 This oxidation results in the disruption of sp2 bonding. In addition, these oxygen sites trap electrons that results in electrically insulating materials. High electrical conductivity can be achieved by effective removal of oxygen species. Various strategies that are reported in the literature for restoring electrical conductivity are chemical reduction, thermally reduction, microwave assisted reduction,24 photocatalysis2,17,25,26 and in situ reduction of GO during processing. Apart from these techniques, some literature exist that utilized UV treatment for further reduction of GO.27 There are plenty of reviews available on the techniques mentioned above.2 However, the state of research on in situ reduction of GO, during the preparation of the nanocomposite, has not received much attention although the field is gaining significant interest among the scientific community.
Given the brevity of this review, we focus here mainly on the in situ reduction of GO during the preparation of conducting nanocomposites. The in situ reduction of GO is carried out mainly during in situ polymerization or during melt processing.28 The following section will highlight the current state of research on in situ reduction of GO to design and develop conducting polymeric nanocomposites from GO as the starting material.
![]() | ||
| Fig. 1 The synthesis of reduced polymer composite via in situ melt polycondensation (Liu et al.29 reproduced by the permission of The Royal Society of Chemistry). | ||
![]() | ||
| Fig. 2 In situ reduction during compositing process (Tang et al.37 reproduced by permission of American Chemical Society Copyright © 2012, American Chemical Society). | ||
Recently, Tang et al.37 explored this strategy for poly(vinylidene fluoride) PVDF based GO composites, wherein they showed well-isolated reduced GO sheets in PVDF. Initially, they dispersed GO in DMF to yield highly exfoliated GO followed by mixing with PVDF. Thus, obtained PVDF/GO composites were then hot pressed at 200 °C for 2 h. Hot pressing leads to in situ reduction of GO that is evident from the change in color from brown to black. This color change is attributed to the efficient removal of oxygen functional groups and partially restoring the graphitic structure. From Fig. 2, it is evident that PVDF/GO composite was initially brown suggesting uniform dispersion of GO in PVDF. However, after reduction, the films turned black. Restoring graphitic structure can lead to π-conjugation, which renders electrical conductivity in the composites. Further, they demonstrated a three-order increase in conductivity with respect to GO composites at a relatively low fraction of 0.16 vol% of reduced GO. Due to immobilization of GO, thin layer of sheets percolate into composites and lead to low percolation threshold. The single layer of reduced GO was further confirmed by TEM (as shown in Fig. 2) and suggested that polymer with higher thermal stability could even enhance the conductivity by complete reduction of GO to graphene.
In another study,38 PVDF/GO composites were prepared by solution mixing in DMF. The resultant composite was coagulated with water. The obtained PVDF/GO mixture was then hot pressed at 200 °C at 50 MPa into sheets of about 0.4 mm thickness. After hot pressing, the color of the composite changed from grey to black suggesting thermal reduction of GO into reduced GO. These sheets were used as master batch for the preparation of PVDF/GO composites using a melt compounder. They demonstrated a significant improvement in conductivity of the composite i.e. ca. 3 × 10−7 S cm−1 at a loading of 0.17 vol% of reduced GO. Further, at 1 vol% of reduced GO, a conductivity of 2 × 10−3 S cm−1 was obtained and the change in conductivity with processing temperature was studied. The processing temperature affects the percolation threshold of TRG. For instance, PVDF–TRG composite processed (hot pressed) at 200 °C exhibited a percolation threshold of 0.12 ± 0.02 vol% (as shown in Fig. 3) in contrast to the composite processed at 190 and 210 °C which showed similar percolation thresholds of 0.17 ± 0.02 vol% and 0.17 ± 0.01 vol%, respectively (Table 1).
![]() | ||
| Fig. 3 Conductivity plot of PVDF with various vol% of TRG–PVDF (He et al.38 reproduced by the permission of The Royal Society of Chemistry). | ||
| Temperature (°C) | Critical percolation (Pc) (vol%) | Conductivity (σc) (S cm−1) |
|---|---|---|
| 190 | 0.17 ± 0.02 | 75.29 ± 16.23 |
| 200 | 0.12 ± 0.02 | 1496 ± 136.38 |
| 210 | 0.17 ± 0.02 | 573.74 ± 63.54 |
The reduction of GO in situ during compression molding was also studied by Ding-Xiang et al.39 wherein they distributed GO on the surface of ultra-high molecular weight poly(ethylene) (UHMWPE) particles followed by compression molding. It is evident that GO sheets are distributed at the interface of UHMWPE particles and further reduction of GO was assisted by hot pressing at 280 °C and 10 MPa pressure for 30 min. Hot pressing ensured reduction of GO to reduced GO in situ. The distribution of reduced GO sheets at the interface yielded a low percolation threshold of 0.66 vol% in UHMWPE and an electrical conductivity of ca. 3.4 S m−1.
Yan et al.40 utilized segregated architectures to distribute reduced GO at the interface of polymer granules rather than distributing in volume in the matrix. These segregated architectures lead to reduction of percolation threshold hence resulting in an increase in electrical conductivity. They utilized polystyrene (PS) of different particle size and observed PS with higher particle size to yield lower percolation threshold. They reported percolation threshold of 0.09 vol% in PS and conductivity of order of ca. 10−2 S m−1 at ∼1 vol% of reduced GO (Fig. 4).
![]() | ||
| Fig. 4 Schematic of in situ reduction of GO in PS and its composite preparation with optical, TEM micrograph of rGO layers at interface of PS and Conductivity with varying in situ reduced GO in PS (Yan et al.40 reproduced by permission of John Wiley and Sons, Copyright © 2014 Wiley Ltd. All rights reserved). | ||
In another study by Shen et al.,41 GO was mixed with poly(vinyl pyrrolidone) (PVP) and glucose (which acts as a reducing agent). GO–PVP and GO–glucose was then mixed with poly(lactic acid) (PLA) in DMF. The resultant solution containing GO–PVP and GO–glucose was then coagulated using methanol. The formed precipitate was dried, transferred to a mold and hot pressed at 210 °C. This process results in the reduction of GO facilitated by both thermal and chemical processes. The resultant composite with 1.25 vol% of GO–glucose showed a significant high conductivity of 2.2 S m−1 in striking contrast to neat PLA (10−15 S m−1). The high electrical conductivity was attributed to the improved exfoliation and dispersion of reduced GO glucose further facilitating in the formation of interconnected conducting network within the PLA matrix. Fig. 5 shows the electrical conductivity as a function of particle concentration for different PLA composites.
![]() | ||
| Fig. 5 Plots of electrical conductivity versus filler loading for PLA composites (Shen et al.41 reproduced by permission of Elsevier, Copyright © 2012 Elsevier Ltd.). | ||
Zhu et al.42 reported reduction of GO by dispersing in propylene carbonate (PC) at pH 3 by bath sonication. The obtained PC solution was heat treated at 150 °C for 12 h resulting in reduced GO. Composite samples were dried in vacuum at 80 °C, and resulted in a conductivity of 2100 and 1800 S m−1 for the samples, which were heat treated at 150 and 200 °C, respectively. Further, vacuum annealing at 250 °C for 12 h resulted in a conductivity of 5230 S m−1 and 2640 S m−1 for the samples which were subjected to heat treatment at 150 and 200 °C, respectively. In this work, they also discussed the dispersion of GO in the PC matrix, and observed that the dispersion of GO in PC may not be due to electrostatic repulsion but can be due to the high dipole moment of PC. Fig. 6 shows the state of dispersion of GO in PC as well as in the solvent.
![]() | ||
| Fig. 6 (a) Optical images of a graphene oxide suspension in PC (top) before and (bottom) after heating at 150 °C for 12 h. (b) SEM image of graphene oxide platelets deposited on a Si substrate (Zhu et al.42 reproduced by permission of American Chemical Society Copyright © 2010, American Chemical Society). | ||
In a recent study, in situ reduction of GO was reported during the compositing process in the presence of surface functionalized octadecylamine (ODA).43 The presence of long octadecyl chain resulted in the hydrophobicity in GO and also facilitated in its efficient reduction. Further, by hot pressing at 210 °C in the presence of PS matrix, electrically conducting samples were obtained. Thermal reduction in presence of PS resulted in a sharp transition from insulator to conductor. The GO–ODA showed 2 orders of higher electrical conductivity when subjected to thermal treatment as displayed in Fig. 7. Hence, it can be concluded that surface modification of GO along with thermal reduction may assist in both improved dispersion and simultaneous increase in electrical conductivity.
![]() | ||
| Fig. 7 The electrical conductivities of GO and GO–ODA before and after the thermal treatment (Li et al.43 reproduced by permission of Elsevier, Copyright © 2011 Elsevier Ltd.). | ||
Another strategy commonly used to reduce GO is by changing polymer chemistry. The latter can play an important role in assisting thermal reduction of GO during hot pressing. Glover et al.44 studied the in situ reduction of GO in poly(vinyl pyrrolidone), poly(vinyl acetate), and poly(vinyl pyrrolidone/vinyl acetate) and they demonstrated that the chemical architecture exhibited by these polymer had strong influence on the extent of thermal reduction. They compared time/temperature relationship for GO reduction in air and in dimethylformamide under the same temperature conditions. The efficiency of reduction was reported based on the change in the C/O (carbon/oxygen) ratio (see Table 2). They showed that reduction of GO depends on time/temperature history and also on the polymer chemistry. They reported an increase in electrical conductivity by 2 orders in all the composites.
| Temperature (°C) | Exposure time | Medium | C content (%) | O content (%) | Atomic C/O ratio |
|---|---|---|---|---|---|
| 25 | — | — | 56 | 39 | 1.9 |
| 55 | 4 months | Water | 63 | 27 | 3.2 |
| 150 | 10 min | Air | 56 | 39 | 1.9 |
| 150 | 240 min | Air | 60 | 36 | 2.2 |
| 175 | 10 min | Air | 59 | 37 | 2.1 |
| 175 | 240 min | Air | 73 | 23 | 4.2 |
| 200 | 10 min | Air | 69 | 26 | 3.5 |
| 200 | 240 min | Air | 81 | 16 | 6.8 |
| 200 | 240 min | DMF | 72 | 12 | 8.0 |
| 250 | 10 min | Air | 80 | 20 | 5.3 |
| 250 | 240 min | DMF | 73 | 12 | 8.2 |
In addition, the polymer chemistry i.e. polarity and aromaticity dictates efficient reduction and the overall state of dispersion. The exact mechanism of reduction of GO is not yet known, which is due to lack of direct measurement of the reduction process and the chemical reactions that take place during processing in the melt. However, with the limited results, Ye et al.45 arrived at some understanding. The decrease in the reducing temperatures in the presence of polymer is related to the interactions between GO sheets and the polymer matrix. Further, Rourke et al.46 proposed that GO are large sheets of graphene with oxidative debris adhered to sheets by π–π stacking or van der Waals interactions.
Fig. 8 illustrates schematically the GO sheets and its interaction with the polymer. From Fig. 8, it is clear that aromatic polymer such as polystyrene (PS) initially adsorb on the graphene sheets by π–π stacking between the continuous phenyl rings and the conjugated basal planes of graphene. This adsorption facilitates intercalation of GO sheets. Further, the π–π stacking counter balances the interactions between graphene sheets and oxidative debris, thus facilitating strong adhesion of polymer chains on the sheets. This adhesion will eliminate the oxidative debris from the surface of the sheet. Both the attachment of polymer chain and the elimination of oxidative debris leads to the reduction of GO thermally, and requires low energy and low reducing temperatures. It is envisaged that polar polymers, such as PMMA, forms hydrogen bonding with the oxidative debris. This further assists in peeling off the oxidative debris from the sheet. As a result, the reduction of GO in polar polymers require relatively less energy and reducing temperatures.
![]() | ||
| Fig. 8 Schematic of GO sheets model and its interaction with polymer (Ye et al.45 reproduced by the permission of The Royal Society of Chemistry). | ||
Apart from enhanced electrical conductivity during in situ reduction of GO, observed to compatibilize immiscible poly(methyl methacrylate)/polystyrene (PMMA/PS, 80/20) blends.47 The droplet diameter of the dispersed phase (PS) phase significantly reduced in the presence of GO. During processing, at higher temperatures, the in situ thermal reduction of GO renders more hydrophobicity in GO and further suppresses the coalescence of PS droplets. However, the electrical conductivity of the samples were not reported in this study.
Recently, Tan et al.,48 covalently functionalized GO with a copolymer and converted GO to reduced GO in situ during molding of PMMA/PS blends. This facile method resulted in localization of GO at the interface of PMMA/PS blends and further resulted in low percolation threshold of 0.02 vol%. This is the lowest reported value in context to graphene as a nanofillers. Further, from Fig. 9, it is evident that the conductivity increased by 2 orders of magnitude for the composite containing P(St-co-MMA) copolymer grafted GO. The TEM image indicated localization of GO sheets at the interface.
![]() | ||
| Fig. 9 Schematic of covalent functionalization of GO and conductivity plot of PS/PMMA blend with rGO and P(St-co-MMA) copolymer grafted GO (Tan et al.48 reproduced by the permission of The Royal Society of Chemistry). | ||
![]() | ||
| Fig. 10 Shows (a) schematic of DSC scans of GO in various zones (b) exhibits conductivity of neat polystyrene composite, quiescent and sheared polystyrene composite (Ye et al.45 reproduced by the permission of The Royal Society of Chemistry). | ||
Another less explored strategy for in situ reduction of GO is the use of masterbatch. You et al.49 prepared a master batch of graphene oxide with styrene-ethylene/butylene-styrene (SEBS) triblock copolymer by melt mixing at 225 °C and they obtained a high degree of reduction. The resultant master batch was further diluted with polystyrene in the subsequent melt mixing process for improving mechanical properties. They reported three orders of magnitude increase in conductivity in reduced GO that was extracted from SEBS as shown schematically in Fig. 11.
![]() | ||
| Fig. 11 In situ reduction of GO via master batch in polystyrene (You et al.49 reproduced by permission of John Wiley and Sons, Copyright © 2014 Wiley Ltd. All rights reserved). | ||
| Matrix | Reduced GO content (vol%) | Reduction method | Conductivity (S m−1) | Ref. |
|---|---|---|---|---|
| — | 100 | — | 750 | 39 |
| UHMWPE | 0.66 | Compression molding | 3.4 | 39 |
| PVDF | 0.75 | Compression molding | 0.02 | 50 |
| PVDF | 1.70 | Solar electromagnetic radiation | 0.04 | 51 |
| PVDF | 2.75 | Compression molding | 0.0002 | 37 |
| PA6 | 1.07 | In situ polymerization | 0.028 | 52 |
| PMMA | 2.79 | Compression molding | 0.94 | 45 |
| SEBS | 2.13 | Compression molding | 0.34 | 45 |
| PET | 17.4 | In situ polymerization | 0.56 | 29 |
From the existing literature concerning the reduction of GO it is apparent that in situ reduction of GO to design highly conducting polymer composites is a better way as compared to ex situ reduction processes which often involves harsh treatments. More emphasis should be given in this direction and more research should be pursued towards in situ reduction of GO during the composite preparation.
Footnote |
| † Both PKS. Mural and M. Sharma contributed equally to this work. |
| This journal is © The Royal Society of Chemistry 2015 |