Chanderpratap
Singh
a,
Ashish Kumar
Mishra
*b and
Amit
Paul
*a
aDepartment of Chemistry, Indian Institute of Science Education and Research (IISER) Bhopal, Madhya Pradesh, India 462066. E-mail: apaul@iiserb.ac.in
bDepartment of Physics, Indian Institute of Science Education and Research (IISER) Bhopal, Madhya Pradesh, India 462066. E-mail: akmishra@iiserb.ac.in
First published on 3rd August 2015
We report a facile approach towards mass production of few layered reduced graphene by low-temperature (160 °C) exfoliation of graphite oxide under ambient atmospheric conditions with the aid of formic acid in a short duration (∼14 min). The obtained reduced graphene showed very high bulk electrical conductivity (1.6 × 103 S cm−1) at room temperature due to restoration of extended conjugation of the sp2 network during rapid exfoliation. Protonation followed by the hydride transfer mechanism has been proposed for the restoration of extended conjugation and high electrical conductivity. BET surface areas of 789 and 1130 m2 g−1 with narrow mesopore distribution (1.9–2.5 nm) were obtained for two different samples prepared by modification in the synthetic methodology. The reduced graphenes were tested as supercapacitors and specific capacitances of 152 and 157 F g−1 with excellent cyclic stability were observed for two samples in aqueous electrolytes.
Herein, we report an example of a chemically assisted low-temperature exfoliation route for mass production of graphene. In this work, judiciously formic acid has been chosen, which governs exfoliation of GO at low temperature (160 °C) within a short duration of 14 min with sudden eruption. The produced graphene exhibits excellent bulk electrical conductivity (1.6 × 103 S cm−1), high surface area (1130 m2 g−1), and good electrochemical energy storage capacity (∼157 F g−1). Moreover, we propose plausible mechanisms based on fundamental organic chemistry reactions for excellent electrical conductivity of RG and also suggest the usage of as-prepared RG as the electrode material for supercapacitors. This approach will lead to new facile ways for the mass production of quality graphene under ambient conditions which can be used as energy devices.
![]() | ||
Fig. 1 Schematic representation of chemically assisted low temperature exfoliation of GO with the aid of formic acid. |
The exfoliation of GO was first confirmed by powder X-ray diffraction (PXRD) and is shown in Fig. 2. The observed shift in 2θ maxima for the (002) plane from 10.8° (GO) to 23.4° (RG) corresponds to interlayer distances of 0.82 nm and 0.38 nm, respectively. Peak broadening of RGs compared to GOs indicates decreases in crystallite lengths along the c-axis (Lc) i.e. decrease in the number of layers.14 Synthesized RGs had around 5 layers of graphene sheets, calculated using Scherrer's formula and the Voigt function16 for peak fitting, as given in the ESI.† As an example, utilizing a low temperature chemically assisted exfoliation method, Zheng and co-workers obtained 5–6 layers of graphene sheets.11 Weak peaks were also observed around a 2θ maximum of 42.5° for the (100) plane which were fitted using the Voigt function and crystallite lengths along the a-axis (La) were calculated using Scherrer's formula. La values of RGs were found to be ∼7.5 nm. Table 1 shows the decrease in crystallite length (25 times decrease along the c-axis i.e. Lc and 11 times decrease along the a-axis i.e. La) for reduced graphenes compared to the parent graphite.
![]() | ||
Fig. 2 PXRD patterns for pristine graphite, GO (150) and RG (150) (a) and graphite, GO (150 + N) and RG (150 + N) (b). |
Sample code | 2θ maxima (°) | (FWHM)Lc (°) | L c (nm) | d (Å) | Number of layers | (FWHM)La (°) | L a (nm) |
---|---|---|---|---|---|---|---|
Graphite (150 μm) | 26.4 | 0.19 | 43 | 3.4 | 127 | 0.19 | 89 |
GO (150) | 10.8 | 0.96 | 8.6 | 8.2 | 11 | 0.60 | 27 |
GO (150 + N) | 10.8 | 0.98 | 8.1 | 8.2 | 10 | 0.66 | 25 |
RG (150) | 23.4 | 4.62 | 1.7 | 3.8 | 5 | 2.32 | 7.0 |
RG (150 + N) | 23.4 | 4.53 | 1.8 | 3.8 | 5 | 2.01 | 8.1 |
Layered structures of RGs were confirmed using scanning and transmission electron microscopy (SEM & TEM) techniques. Fig. 3a and e show the SEM images of RG (150) and RG (150 + N) samples, respectively, depicting few layered structures of RGs after rapid exfoliation. RG (150) shows more separated layers compared to those for RG (150 + N). TEM images (Fig. 3b and f) indicate entangled few layer sheets for RG (150) and RG (150 + N), respectively. High resolution TEM images (Fig. 3c and g) indicate 3–5 layer distribution in each stack of RG samples. Fig. 3d and h show the selected area electron diffraction (SAED) pattern for RG samples. The SAED pattern for RG (150) indicates well-defined diffraction spots corresponding to the hexagonal structure of planar graphene sheets, while RG (150 + N) shows a diffused circular ring due to different orientation and large distribution of layers.8 It reveals that the degree of exfoliation was more controlled in RG (150) compared to RG (150 + N) leading to retention of the hexagonal symmetry in RG (150). Furthermore, qualities of synthesized RGs were evaluated by TGA and the results are shown in the ESI (Fig. S2†). RG (150) and RG (150 + N) showed a weight loss of 8 and 11%, respectively up to 250 °C, which can be attributed to the loss of water and labile oxygen functionalities. Furthermore, 7.5 and 11% weight losses were observed for RG (150) and RG (150 + N), respectively, in the temperature range 250–510 °C, which can be attributed to the loss of residual functional groups. The higher weight loss for RG (150 + N) indicates the presence of more functional groups in RG (150 + N). In a previous report by our group,14 we concluded that the addition of NaNO3 during GO synthesis favors strong basal plane oxidation on graphene sheets which generates more functional groups on basal planes and hence slightly more functional groups may remain in graphene sheets for RG (150 + N) compared to RG (150). This conclusion can be further supported by the lower conductivity of RG (150 + N) compared to RG (150) (vide infra).
As a strong tool to characterize sp2 carbon domains, Raman spectroscopy was performed for synthesized RG samples and the spectra are shown in the ESI (Fig. S3†). RGs exhibit a D band around 1330 and a G band around 1589 cm−1. The D band corresponds to disorder while the G band indicates the in-plane vibrations of carbon atoms. A large shift in the G band of RGs compared to parent graphite was observed, indicating few layered structures of RGs.17 The intensity ratio of D band to G band (ID/IG) directly measures the degree of disorder.18 Decreased ID/IG ratios for RGs compared to those of GOs indicate the reduced disorder due to thermal exfoliation as described by Kudin and co-workers.19
BET adsorption/desorption profiles for N2 are shown in Fig. 4a which exhibit characteristics of type-III and IV isotherms with a hysteresis loop of H2 type (according to IUPAC classification) in the range (0.45–1.0) for P/P0, implying the mesoporous nature of synthesized RGs. BET surface areas of 789 and 1130 m2 g−1 with pore volumes of 5.5 and 11.84 cm3 g−1 were observed for RG (150) and RG (150 + N), respectively. RGs showed narrow pore size distribution in the range of 1.9–2.5 nm, calculated using the Barrett–Joyner–Halenda (BJH) method.20 For comparison, Zheng and co-workers11 reported a surface area of 500 m2 g−1 by thermal exfoliation at 130 °C, Aksay and co-workers9 reported 650 m2 g−1 at 1050 °C, and Ruoff and co-workers21 reported a surface area of 705 m2 g−1 for chemically exfoliated graphene sheets. Surface areas for RGs achieved by the present method are higher than those by other thermal and chemical exfoliation methods.
FT-IR spectra of GOs and RGs are compared in Fig. 4b. A significant decrease in the intensity of stretching vibrations of CO (1654 cm−1), C–O (1080 cm−1) and an increase in the stretching vibration of C
C (1466 cm−1) were observed from GOs to RGs. These results indicate the reducing nature of formic acid towards selective oxygen functionalities. Restoration of π conjugation in synthesized RGs was evident from electrical conductivity measurements. High bulk electrical conductivities of (1.6 ± 0.09) × 103 and (1.4 ± 0.07) × 103 S cm−1 were observed for RG (150) and RG (150 + N) respectively, measured by a four probe method. Bulk electrical conductivity achieved in the present work is found to be superior to most of the reports as summarized in Table 2. The excellent electrical conductivity of synthesized RGs can be attributed to the significant decrease in the C
O stretching (Fig. 4b), as described in other reports.7,22
Sample | Preparation method | Electrical conductivity (S cm−1) | Reference |
---|---|---|---|
Reduced graphene | Hydrogen arc discharge at 1050 °C | 2 × 103 | 7 |
Reduced graphene | Hydrazine reduction at 100 °C | 2 | 6 |
Reduced graphene | H2SO4 assisted exfoliation at 120 °C | 17 | 12 |
Reduced graphene | HCl assisted exfoliation at 130 °C | 12 | 11 |
Reduced graphene | Formic acid reduction at 160 °C | 1.6 × 10 3 | Present study |
The plausible mechanisms for chemically assisted thermal exfoliation are shown in Scheme 1. Basal planes of graphene sheets in GOs contain functional groups such as hydroxyl and epoxy attached with sp3 hybridized carbons and sheet edges contains the carbonyl functionality attached with sp2 hybridized carbons. Protonation followed by hydride (H−) transfer from formic acid (HCOOH) and thereafter release of gases such as H2, CO2 and H2O will lead to removal of these functional groups from graphene sheets.15,23 Released gases created extra pressure for the expansion of graphene sheets at low temperature. The most important feature of formic acid assisted exfoliation is that not only the functional groups were removed but also the driving forces for these reactions were to change the hybridization of carbon atoms from sp3 to sp2 for restoration of π networks. Hence, the mechanism also explains the superior conductivity of RGs.
![]() | ||
Scheme 1 Mechanisms for the reactions between different functionalities of GO and formic acid. (i) Hydroxyl, (ii) epoxy, and (iii) carbonyl. |
Inspired by these observations, electrochemical experiments were performed to test these RGs as electrochemical supercapacitors using a regular three electrode cell by depositing the materials on a platinum (Pt) electrode with poly(vinylidene fluoride) binder in 2 M H2SO4 as electrolyte. Nearly rectangular cyclic voltammograms (CV) were obtained (Fig. 5a) even at a faster scan rate of 100 mV s−1 indicating quality charge propagation.21 The CVs at slower scan rates are shown in the ESI (Fig. S6†). Specific capacitances of 157 and 152 F g−1 were observed for RG (150) and RG (150 + N), respectively at a voltage sweep rate of 5 mV s−1. Retention of nearly 86 and 78% for specific capacitances of RG (150) and RG (150 + N), respectively, were observed at a high voltage scan rate of 100 mV s−1 (Fig. 5b). The drop in specific capacitances with increasing scan rate can be attributed to the pseudocapacitance (redox) contribution from the remaining oxygen functionality. These redox contributions in RGs arise due to the presence of hydroxyl functionalities on the sheet edges of graphene sheets24,25 and according to Scheme 1, formic acid cannot remove these functionalities from graphene sheets. TGA results suggest that the functionalities present in RG (150 + N) were higher than those in RG (150) but the electrochemical results show that the redox contribution for RG (150) was higher than that for RG (150 + N). Apparently, these results may seem to contradict with each other but can be understood utilizing the observations of our previous studies14 where we reported that GO synthesis in the absence of NaNO3 results in higher sheet edge oxidation whereas the presence of NaNO3 during GO synthesis results in higher basal plane oxidation on graphene sheets. Hence, RG (150) may have more hydroxyl functionalities on the edges of graphene sheets compared to RG (150 + N) which is responsible for the enhanced redox response of RG (150). It is also important to mention that specific capacitances of RGs in the present study were comparable or greater than most of the other reports for graphene, as shown in Table 3. The high surface area and good electrical conductivity of as-synthesized RGs result in good capacitance performance.
Sample | Preparation method | Specific capacitance (F g−1) | References |
---|---|---|---|
Reduced graphene | Hydrazine reduction (100 °C) | 101 F g−1 at 20 mV s−1 | 21 |
Reduced graphene | Thermal exfoliation (200 °C) in H2 | 80 F g−1 at 10 mV s−1 | 29 |
Reduced graphene | Thermal exfoliation (1050 °C) | 117 F g−1 at 5 mV s−1 | 30 |
Reduced graphene | Thermal exfoliation (1000 °C) | 150 F g−1 at 5 mV s−1 | 31 |
Reduced graphene | Formic acid reduction (160 °C) | 157 F g −1 at 5 mV s −1 | Present work |
The specific capacitances of RGs were also calculated using galvanostatic charge/discharge measurements (Fig. 5c). Calculated specific capacitances with both methods indicate strong agreement between two electrochemical techniques (Table 4). Nyquist plots were analyzed to understand the rate of mass transfer in RGs (Fig. 5d). In the case of RG (150), diffusion of the electrolyte started at 215 Hz corresponding to 0.9 Ω resistance and completed at 1.8 Hz corresponding to 1.5 Ω resistance and thereafter complete capacitor behavior was achieved.26 In the case of RG (150 + N) diffusion of the electrolyte started at 175 Hz corresponding to 0.9 Ω resistance and completed at 1.2 Hz corresponding to 1.7 Ω resistance and thereafter capacitor behavior was completely achieved. These results indicate excellent mass transfer inside the RGs. The excellent mass transport can be explained by the narrow mesopore distribution in the range (1.9–2.5 nm) (vide supra) in RGs which could decrease the ion-diffusion length, hence minimizing inner-pore resistance and also providing large accessible surface area for ion transport/charge storage.27,28 The cyclic performance (Fig. 5e) indicates nearly 100% stability for both the samples after 1000 cycles, indicating the robustness of the material and excellent ion accessibility in the graphene framework during long term cycles.27 The phase angle in the EIS experiment was found to be 83° very close to 90° at low frequencies, indicating the capacitive behaviour of produced RGs (Fig. 5f).
Scan rate (mV s−1) | RG (150) | RG (150 + N) | Current densities (A g−1) | RG (150) | RG (150 + N) |
---|---|---|---|---|---|
5 | 157 | 152 | 1 | 163 | 157 |
10 | 151 | 143 | 2 | 145 | 150 |
20 | 151 | 135 | 5 | 134 | 136 |
50 | 139 | 126 | 10 | 127 | 123 |
100 | 135 | 118 |
The higher capacitance and low cost production of prepared RGs compared to other reports29–31 clearly indicate the suitability of prepared RGs for supercapacitor electrodes. The narrow mesoporous texture of RGs reduces the inner pore resistance and facilitates high mass transfer even at higher frequencies. Additionally, the higher electrical conductivity of RGs can reduce the ohmic loss and cut down the internal resistance for better charge propagation, as observed in the near square shape of CV curves at high scan rates (Fig. 5a).
Footnote |
† Electronic supplementary information (ESI) available: Additional experimental details, electrochemical data, TGA results, Raman spectra, EDS, PXRD and elemental analysis. See DOI: 10.1039/c5ta04655f |
This journal is © The Royal Society of Chemistry 2015 |