06.01.2013 Views

Graphene Sheets from Graphitized Anthracite Coal: Preparation ...

Graphene Sheets from Graphitized Anthracite Coal: Preparation ...

Graphene Sheets from Graphitized Anthracite Coal: Preparation ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Graphene</strong> <strong>Sheets</strong> <strong>from</strong> <strong>Graphitized</strong> <strong>Anthracite</strong> <strong>Coal</strong>: <strong>Preparation</strong>,<br />

Decoration, and Application<br />

Quan Zhou, † Zongbin Zhao,* ,† Yating Zhang, ‡ Bo Meng, † Anning Zhou, ‡ and Jieshan Qiu* ,†<br />

† Carbon Research Laboratory, Liaoning Key Lab for Energy Materials and Chemical Engineering, State Key Lab of Fine Chemicals,<br />

School of Chemical Engineering, Dalian University of Technology, Dalian 116024, China<br />

‡ School of Chemistry and Chemical Engineering, Xi'an University of Science and Technology, Xi'an 710054, China<br />

ABSTRACT: <strong>Coal</strong> has been used as an important resource for the production of chemicals, conventional carbon materials, as<br />

well as carbon nanomaterials with novel structures, in addition to its main utilization in the energy field. In this work, we present<br />

the synthesis of chemically derived graphene and graphene−noble metal composites with coal as the starting material by means<br />

of catalytic graphitization, chemical oxidation, and dielectric barrier discharge (DBD) plasma-assisted deoxygenation. It is found<br />

that the graphitization degree of the coal-derived carbon remarkably affects the properties of graphene obtained <strong>from</strong> chemical<br />

exfoliation, and high crystallinity of coal-derived carbon is essential for the preparation of high-quality graphene sheets (GS). GS<br />

decorated with highly dispersed noble metallic nanoparticles (NP) on their surface (NP/GS) were successfully fabricated via<br />

simultaneous reduction of graphite oxide (GO) and noble metal salts by H 2 DBD plasma technique. The electrochemical<br />

performance of the GS as electrode in supercapacitor and the catalytic activities of NP/GS composites in selective reduction of<br />

nitrogen oxides (NO x) were investigated. This work demonstrates an alternative approach for the fabrication of graphene and its<br />

composites <strong>from</strong> coal with promising potential in energy storage and environment preservation.<br />

1. INTRODUCTION<br />

As a valuable and abundant solid fuel in nature, coal provides<br />

most of the energy for the economic development and livingused<br />

energy in China. Besides the contribution to energy, coal<br />

can be exploited as an important resource for the production of<br />

chemicals, such as creosote oil, naphthalene, phenol, and<br />

benzene, etc. <strong>Coal</strong>-based carbon materials, such as activated<br />

carbon, carbon blocks, and carbon electrodes, have been<br />

produced worldwide for a long time. 1,2 With the discovery and<br />

development of carbon nanomaterials, the preparation of<br />

carbon nanomaterials <strong>from</strong> the cheapest and most abundant<br />

natural carbon source�coal�has attracted increasing attention<br />

in the last few decades. 3 Wilson and his co-workers first<br />

reported the preparation of C60, C70, 4 and CNTs 5 <strong>from</strong> coal<br />

by using the arc-discharge method with coal-based carbon as<br />

the evaporation anode. Over the past decade, our group has<br />

intensively investigated the possibility and feasibility of<br />

preparing various carbon nanomaterials <strong>from</strong> Chinese coal. 6<br />

As a novel two-dimensional carbon nanomaterial, graphene<br />

has attracted tremendous scientific attention in recent years due<br />

to its excellent physical and chemical properties. 7−9 To further<br />

broaden the application potential of graphene-based materials,<br />

graphene nanocomposites are being developed and receiving<br />

increased attention, especially the decoration with metal or<br />

metal oxide nanoparticles. Various carbon sources, including<br />

graphite, 10 hydrocarbons (CH4, C2H2), 11,12 polymer<br />

(PMMA), 13 natural biomaterials (food, insects), and even<br />

plastic waste 14 have been widely used for preparing graphene.<br />

However, as the most abundant carbon source in the world,<br />

coal has not been investigated for the production of graphene<br />

up to now. <strong>Graphene</strong> layers in the structure of coals have been<br />

investigated in molecular level. 15,16 Tomita et al. evaluated the<br />

size of graphene sheets in several Chinese anthracites by a<br />

temperature-programmed oxidation method. 17 Furthermore,<br />

coal possesses abundant polyaromatic structures that are quite<br />

similar to sp 2 bonding characteristics of graphene; therefore, it<br />

is reasonable to expect that coal can be used as starting material<br />

for the fabrication of graphene and graphene-based composite<br />

materials.<br />

Herein, we present the successful synthesis of chemically<br />

derived graphene and composites of graphene sheets (GS)<br />

decorated with noble metal nanoparticles <strong>from</strong> coal by means<br />

of catalytic graphitization and chemical oxidation combined<br />

with dielectric barrier discharge (DBD) plasma technique. The<br />

graphitization degree of the coal-derived carbon precursor is<br />

controlled by using catalyst, and the correlation between the<br />

graphitization degree of the coal-derived graphite-like carbon<br />

and the resultant graphene quality has been clarified. The<br />

performance of as-prepared chemically derived graphene as<br />

electrode in supercapacitor was investigated, and the graphenebased<br />

composites of GS decorated with noble metal (Pt, Ru,<br />

and PtRu) nanoparticles were used as catalyst in selective<br />

catalytic reduction (SCR) of nitrogen oxides (NO x) with<br />

ammonia.<br />

2. EXPERIMENTAL SECTION<br />

2.1. <strong>Preparation</strong> of <strong>Coal</strong>-Derived Graphite Oxide. Taixi coal<br />

(TX), <strong>from</strong> Shanxi province of China, was used as the starting material<br />

for the preparation of GS. The raw coal was ground and sieved to a<br />

powder, and the obtained particle size of coal powder is less than 0.5<br />

mm. The coal powders were milled for 1.5 h by ball milling (the ratio<br />

of ball and coal is 20:1), and the diameter of more than 90% of them is<br />

less than 15 μm after milling.<br />

Received: October 28, 2011<br />

Revised: July 17, 2012<br />

Published: July 17, 2012<br />

Article<br />

pubs.acs.org/EF<br />

© 2012 American Chemical Society 5186 dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192


Energy & Fuels Article<br />

Table 1. Analysis Data of Taixi <strong>Coal</strong> and Its Demineralization <strong>Coal</strong><br />

proximate analysis (wt %) ultimate analysis (daf, wt %)<br />

samples M ad A d V daf C H N S O a<br />

TX 1.38 15.40 8.56 91.10 3.34 0.89 0.36 4.35<br />

TX-de 1.29 2.52 8.08 93.90 3.47 0.77 0.08 1.81<br />

a By difference.<br />

The raw coal was treated with hydrofluoric acid and concentrated<br />

hydrochloride in plastic beakers at 50 °C for 4 h to make deashed coal<br />

(TX-de), and then it was washed with distilled water until no Cl − and<br />

F − were detected in the filtrate.<br />

The transformation of coal into graphite-like carbon was conducted<br />

by heat treatment in the presence of Fe. First, the TX-de and<br />

Fe 2(SO 4) 3 [TX-de:Fe 2(SO 4) 3 = 16:12.6] was well-mixed by ball<br />

milling for 2 min, and then the mixture was subjected to catalytic<br />

graphitization (C-G) at 2400 °C for 2 h under argon, and the product<br />

obtained was named TX-C-G. For comparison, TX was noncatalytically<br />

graphitized (NC-G, without the additional catalyst) under similar<br />

condition to give rise to the product defined as TX-NC-G. The<br />

proximate analysis and ultimate analysis of TX and TX-de are listed in<br />

Table 1. Both TX-C-G and TX-NC-G described above were oxidized<br />

by Hummers method to get their corresponding graphite-like carbon<br />

oxides, 18 TX-C-GO and TX-NC-GO, respectively.<br />

2.2. <strong>Preparation</strong> of <strong>Coal</strong>-Derived <strong>Graphene</strong> by H 2 DBD<br />

Plasma. Synthesis of chemically derived graphene was carried out in a<br />

DBD reactor with H 2 as working gas at ambient atmosphere<br />

pressure. 19 A weighted amount of TX-C-GO or TX-NC-GO powder<br />

was put into a vertical quartz tube (10 mm in diameter, about 100 mm<br />

in length of discharge) with porous plate in the middle. Prior to the<br />

discharge, the reactor was purged with H 2 to exhaust the air in the<br />

quartz tube. Afterward, the DBD plasma was initiated under 50 V ×<br />

1.2 A input ac power at room temperature and atmospheric pressure<br />

for 5 min; significant volume expansion can be observed in a flash. The<br />

product GS derived <strong>from</strong> TX-C-GO and TX-NC-GO was named TX-<br />

C-GS and TX-NC-GS, respectively.<br />

2.3. <strong>Preparation</strong> of GS Decorated with Noble Metallic<br />

Nanoparticles. GS decorated with noble metal nanoparticles (NP/<br />

GS) were prepared as follows. 20 The TX-C-GO made as in section 2.1<br />

with a high oxidation degree was used as the starting material for the<br />

support of metal nanoparticles. As a typical procedure to synthesize<br />

the NP/GS composite nanostructures, a certain proportion of TX-C-<br />

GO and precursors of noble metal (H 2PtCl 6·6H 2O, RuCl 3·3H 2O)<br />

were dispersed in ethanol with ultrasonication for 1 h. Subsequently,<br />

the mixed solutions were vacuum-dried at 100 °C for 10 h. The assynthesized<br />

solid powder mixtures were exposed in H 2 plasma with 50<br />

V × 1.2 A ac input power for 10 min to simultaneously reduce the GO<br />

and precursors of noble metal ions to produce NP/GS composites.<br />

2.4. Electrochemical Measurement. For the electrochemical<br />

measurement, the fabrication of working electrodes was carried out as<br />

follows. Typically, the electroactive materials and polytetrafluoroethyene<br />

(PTFE) binder (without any other carbon additive) were<br />

mixed in a mass ratio of 9:1 and dispersed in ethanol, and the resulting<br />

mixture was dried at 60 °C. After drying, it was pressed into a wafer of<br />

8 mm diameter and assembled into an electrode with two pieces of<br />

nickel foam substrate (10 mm in diameter).<br />

The electrochemical properties of chemically derived graphene were<br />

measured in a standard three-electrode system with a Pt sheet as<br />

counter electrode and Hg/HgO electrode as reference electrode in a 6<br />

M KOH aqueous electrolyte. CV curves (scanning rates varying <strong>from</strong> 2<br />

to 100 mV/s) were measured with an electrochemical workstation<br />

(CHI 660D). Galvanostatic charge/discharge measurements (current<br />

density ranged <strong>from</strong> 50 to 1000 mA/g) were conducted with a<br />

charge−discharge tester (Arbin BT2000), and the electrochemical<br />

capacitances were obtained <strong>from</strong> charge−discharge curves.<br />

2.5. SCR Catalytic Activity Measurement. SCR of NO with<br />

NH 3 over the catalysts was carried out in a continuous-flow fixed-bed<br />

quartz reactor under atmospheric pressure. Typically, 50 mg catalyst<br />

was packed inside the reaction zone, the initial gas composition was<br />

5187<br />

400 ppm NO x (approximately 92% NO, 8% NO 2), 600 ppm NH 3,2%<br />

O 2, and N 2 (balance), and a total gas flow rate of 100 mL/min (STP)<br />

was used throughout this study. NO x concentrations were analyzed<br />

using a flue gas analyzer (KM9106 Quintox, Kane International<br />

Limited) online.<br />

3. RESULTS AND DISCUSSION<br />

The proximate analysis and ultimate analysis results of TX coal<br />

and its demineralization coal TX-de are shown in Table 1. The<br />

high carbon content (91.10%) and low content of volatile<br />

matter (8.56%) suggest that the coal is a typical anthracite coal<br />

with high metamorphic degree. After the treatment with<br />

hydrofluoric acid and concentrated hydrochloride, the ash<br />

content was decreased to 2.52 wt % <strong>from</strong> an initial value of<br />

15.40 wt %, indicating that the mineral matter in coal has been<br />

efficiently removed.<br />

XRD patterns of samples, including coal-based graphite-like<br />

carbons (TX-NC-G, TX-C-G), their corresponding oxides<br />

(TX-NC-GO, TX-C-GO), and the final product graphene<br />

sheets (TX-NC-GS, TX-C-GS) derived <strong>from</strong> the coal-based<br />

graphite oxides after DBD plasma treatment, are shown in<br />

Figure 1. There are significant differences in crystallite size<br />

Figure 1. XRD patterns of coal-based graphite (TX-NC-G, TX-C-G),<br />

coal-based graphite oxides (TX-NC-GO, TX-C-GO), and graphene<br />

sheets (TX-NC-GS, TX-C-GS) <strong>from</strong> DBD plasma.<br />

(La), stacking height (Lc), and degree of graphitization<br />

between TX-NC-G and TX-C-G. The La and Lc of the<br />

crystallite can be determined using the Scherrer equation 21<br />

La = 1.84 λ/Bacos( φa)<br />

Lc = 0.89 λ/Bccos( φc)<br />

where λ is the wavelength of the radiation used, Ba and Bc are<br />

the width of the (100) and (002) peaks, respectively, at 50%<br />

height, and φa and φc are the corresponding scattering angles<br />

or peak positions. The La of TX-NC-G and TX-C-G is 0.23 and<br />

0.92 nm, and the Lc of TX-NC-G and TX-C-G is 0.27 and 0.48<br />

nm, respectively. The graphitization degree (G) was calculated<br />

by using the following equation<br />

dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192


Energy & Fuels Article<br />

Figure 2. (a, b) Digital photos of samples before and after treatment with H 2 discharge plasma. SEM and TEM images of TX-NC-GS (c, d, e) and<br />

TX-C-GS (f, g, h) with different magnification.<br />

Figure 3. (a) FTIR spectra of coal-based graphite oxides and their chemically derived GS. (b) Nitrogen adsorption and desorption isotherms at 77 K<br />

of GS <strong>from</strong> H 2 discharge plasma (inset: pore-size distribution). (c) Raman spectra of coal-based graphite, GO, and GS.<br />

G = (0.3440 − d )/(0.3440 − 0.3354)<br />

002<br />

where d 002 is the interlayer spacing of (002) calculated <strong>from</strong><br />

XRD patterns. From the XRD patterns, it can be calculated that<br />

the graphitizing degree of TX-C-G (91.98%) is much higher<br />

than that of TX-NC-G (66.28%). It is well-known that the<br />

diffraction peaks of (100), (004), and (110) will appear when<br />

high graphitization degree was achieved, and these peaks can be<br />

observed in the XRD pattern of TX-C-G rather than TX-NC-G.<br />

Therefore, it can be concluded that the applied catalyst<br />

substantially promoted graphitization of the coal.<br />

After oxidation of the coal-based graphite, their sharp (002)<br />

peaks around 26.2° basically disappeared, while the as-prepared<br />

TX-NC-GO and TX-C-GO gave rise to (001) peaks located at<br />

11°, corresponding to an increasing interlayer spacing <strong>from</strong><br />

0.335 to 0.749 nm, indicating that the TX-NC-G and TX-C-G<br />

were efficiently oxidized and oxygen was bonded to their planar<br />

surface. After DBD plasma treatment, (001) peaks of TX-NC-<br />

GO and TX-C-GO disappeared almost completely to give rise<br />

to TX-NC-GS and TX-C-GS, respectively, without any distinct<br />

peaks in their XRD profiles.<br />

5188<br />

Figure 2a,b is the optical photos of coal-based graphite oxides<br />

(TX-NC-GO, TX-C-GO) and their corresponding resultant<br />

products (TX-NC-GS, TX-C-GS) after H 2 discharge plasma<br />

under the same experimental conditions. Obviously, the volume<br />

of TX-C-GS increased remarkably compared with its precursor<br />

TX-C-GO, while the volume change <strong>from</strong> TX-NC-GO to TX-<br />

NC-GS increased only slightly after the DBD plasma treatment.<br />

It can also be observed that TX-C-GS is bulky and fluffy in<br />

appearance while TX-NC-GS is relatively compacted. Taking<br />

into consideration the higher graphitization degree of TX-C-G<br />

compared with TX-NC-G, it can be concluded that the<br />

expansion level of graphite oxide has a close relationship with<br />

the graphitization degree of its precursor. In other words, the<br />

high graphitization degree is of great benefit to the intercalation<br />

of graphite by the oxygen-containing groups during oxidation.<br />

The oxygen intercalated into the interlayer spacing of graphite<br />

will be removed to form gases (H 2O and CO 2) during the<br />

discharge process, and the yielding pressure overcomes the van<br />

der Waals force between the layers, so that the oxides can be<br />

exfoliated immediately during DBD plasma process. For the<br />

low level graphitization carbon, the interaction hardly takes<br />

dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192


Energy & Fuels Article<br />

Figure 4. XPS spectra C1s peaks of coal-based graphite oxides and graphene sheets: TX-C-GO (a), TX-C-GS (b), TX-NC-GO (c), and TX-NC-GS<br />

(d).<br />

place during the oxidation process; therefore, the driving force<br />

resulting <strong>from</strong> gases (H 2O and CO 2) formed during DBD is<br />

not strong enough to overcome interlayer van der Waals force<br />

for complete expansion.<br />

Typical morphologies of TX-NC-GS and TX-C-GS are<br />

shown in Figure 2. It is revealed that layer structures were<br />

formed after expansion via DBD plasma (Figure 2c,f). Highmagnification<br />

SEM images (Figure 2d,g) of GS show a<br />

wrinkled, thin, sheetlike structure similar to that of graphene<br />

prepared <strong>from</strong> natural flake graphite by thermal expansion 22<br />

and sonication-assisted chemical reduction. 9 TEM images<br />

(Figure 2e,h) show that these sheet structures have plenty of<br />

ripples and folded regions in their basal planes and seem to be<br />

transparent, possibly owing to their being ultrathin or even<br />

single sheets. It is obvious that the layer of TX-NC-GS appears<br />

to be thicker than that of TX-C-GS due to the incomplete<br />

exfoliation of the former. This is consistent with the volume<br />

change observation during plasma-assisted exfoliation, as shown<br />

in Figure 2a,b.<br />

There are a variety of oxygen-containing groups on the<br />

surface of TX-NC-GO and TX-C-GO (FTIR spectra; see<br />

Figure 3a), which mainly include the C�O, C−O, and O−H.<br />

After H 2 plasma reduction, the peaks for oxygen functional<br />

groups were significantly reduced with fewer C�O groups left<br />

in TX-NC-GS and TX-C-GS compared with their corresponding<br />

precursors. Generally, the wide peak at 1100−1420 cm −1<br />

could mainly be assigned to the �C−H stretch, and the peak<br />

at 1583 cm −1 could be assigned to the aromatic C�C stretch.<br />

According to the FTIR spectra of coal-based graphite oxides,<br />

the C�C bond almost cannot be detected, while there is a<br />

visible C�C peak located at 1583 cm −1 in their GS. The<br />

typical peak of C�O located at 1731 cm −1 of TX-NC-GS is<br />

5189<br />

stronger than that of TX-C-GS, which may indicate that the<br />

former has more oxygen-containing groups.<br />

The coal-based graphene sheets were further characterized<br />

via N 2 adsorption and desorption at 77 K for deeply<br />

understanding their pore structure. According to the results,<br />

the BET surface area values of TX-C-GS and TX-NC-GS are<br />

306 and 135 m 2 /g, respectively. As shown in Figure 3b, the<br />

isotherms have the characteristics of type IV with a type H3<br />

hysteresis loop at relatively high pressure, revealing that these<br />

materials are composed of aggregated sheets (loose assemblages)<br />

possessing typical mesopore structures, corresponding<br />

with distributions of the most probable pore sizes of 3.69 nm<br />

(TX-NC-GS) and 4.02 nm (TX-C-GS), as shown in the inset<br />

of Figure 3b and the microscopy observations mentioned<br />

above.<br />

The significant structural changes occurring during the<br />

transformation <strong>from</strong> coal-based graphite to the GS are reflected<br />

in their Raman spectra (Figure 3c). As known, the G band<br />

corresponds to the first-order scattering of the E 2g mode<br />

observed for sp 2 carbon domains, and the pronounced D band<br />

is the disordered band associated with structural defects,<br />

amorphous carbon, or edges that can break the symmetry and<br />

selection rule. 23 The Raman spectra of TX-NC-G and TX-C-G<br />

display a prominent G peak as the feature at 1581 cm −1 . The<br />

strong 2D peak at 2663 cm −1 indicates that the Taixi coal has<br />

been transformed into graphite-like carbon after graphitization<br />

treatment. As expected, the I D/I G ratio of TX-C-G is smaller<br />

than that of TX-NC-G, which indicates that TX-C-G possesses<br />

higher graphitization degree than TX-NC-G, consistent with<br />

the results obtained <strong>from</strong> XRD. From the Raman spectra of<br />

coal-based graphite oxides (TX-NC-GO and TX-C-GO), it can<br />

be seen that the G bands seem to be broadened compared with<br />

dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192


Energy & Fuels Article<br />

Figure 5. Electrochemical performance of coal-derived graphene sheets. (a) Cyclic voltammetry curves of GS with the scanning rates of 5 mV/s. (b)<br />

Specific capacitance against different discharge current density with 6 M KOH solution as an electrolyte. (c) Galvanostatic charge/discharge curves<br />

with the current density of 100 mA/g. (d) The cycling performance at a charge/discharge current of 100 mA/g.<br />

their precursors, meanwhile, the D bands at 1337 cm −1 become<br />

much more prominent, indicating the reduction in size of the<br />

in-plane sp 2 domains, possibly due to the extensive oxidation. 10<br />

The Raman spectra of the TX-NC-GS and TX-C-GS also<br />

contain both G and D bands (at 1590 and 1350 cm −1 ,<br />

respectively), however, with a decreased I D/I G intensity ratio<br />

compared to that in TX-NC-GO and TX-C-GO, which<br />

suggests an increase in the average size of the sp 2 domains<br />

and restoration of sp 2 network during reduction of GO in H 2<br />

discharge plasma process. 24,25 It should be noted that no<br />

obvious 2D bands can be observed in the Raman spectra of TX-<br />

NC-GS and TX-C-GS due to the disordered structure (defects,<br />

vacancies, or distortions of the sp 2 domains) caused by the use<br />

of strong acid (H 2SO 4) and strong oxidant (KMnO 4) during<br />

the oxidation stage. We can reasonably believe that it is very<br />

difficult for the graphene sheets to retrieve the graphitic<br />

structures through DBD plasma treatment. Compared with TX-<br />

NC-GS, the D band of TX-C-GS is weaker and wider, which<br />

suggests fewer structural defects in the latter than the former.<br />

As mentioned above, it can be concluded that the crystallinity<br />

of coal-derived carbon remarkably affects the structures and the<br />

properties of the graphene obtained <strong>from</strong> chemical exfoliation.<br />

26<br />

The oxygen-containing functional groups of TX-NC-GO,<br />

TX-NC-GS, as well as TX-C-GO and TX-C-GS were<br />

characterized by X-ray photoelectron spectroscopy (XPS).<br />

The high-resolution C1s spectrum of the TX-C-GO (Figure<br />

4a) and TX-NC-GO (Figure 4c) reveal that there are three<br />

main components arising <strong>from</strong> C�C (aromatic rings), C−O<br />

(alkoxy and epoxy), and C�O (carboxyl) groups. After DBD<br />

plasma treatment, carbon with different chemical valences<br />

remains for TX-NC-GS and TX-C-GS in XPS spectra; however,<br />

the peak intensities of them are much smaller than those in TX-<br />

5190<br />

NC-GO and TX-C-GO. As shown in Figure 4b,d, there are no<br />

O−C�O groups left in TX-NC-GS and TX-C-GS and the<br />

C�C bonds become dominant, while the peak intensities of<br />

C�O and C−O are obviously lower than that of TX-NC-GO<br />

and TX-C-GO, which indicates that most of the oxygencontaining<br />

groups in TX-NC-GO and TX-C-GO have been<br />

removed and the majority of the conjugated graphene networks<br />

are restored 27 by DBD plasma treatment for only 5 min. The<br />

residual oxygen-containing groups on the graphene sheets will<br />

benefit the wettability of the electrode in supercapacitor and<br />

metal or metal oxide catalyst grafting for heterogeneous<br />

catalysis. However, compared with TX-C-GS, it is worth noting<br />

that there are more significant C�O and C−O peaks left in the<br />

XPS spectrum of TX-NC-GS, which is consistent with the<br />

FTIR results.<br />

The electrochemical properties of these coal-derived<br />

graphene were measured in a 6 M KOH aqueous electrolyte<br />

with a three-electrode supercapacitor cell at room temperature<br />

by a CHI 660D electrochemical workstation. Figure 5a shows<br />

the CV measurement results of TX-NC-GS and TX-C-GS at<br />

the scanning rate of 5 mV/s. The CV curve of TX-C-GS<br />

exhibits a more typical rectangular shape than that of TX-NC-<br />

GS, indicating that the TX-C-GS has a better capacitive<br />

behavior and charge propagation at the surface of electrode<br />

following the electric double layer charging mechanism<br />

compared with TX-NC-GS. The specific capacitances of TX-<br />

NC-GS and TX-C-GS against the discharge current density are<br />

shown in Figure 5b, which was calculated <strong>from</strong> the<br />

galvanostatic charge and discharge curves at different current<br />

density, using the following equation 28<br />

C = IΔt/ mΔV dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192


Energy & Fuels Article<br />

Figure 6. TEM images of Pt/GS (a), Ru/GS (b), and PtRu/GS (c). (d) XRD patterns of NP/GS composites. (e) NO conversion as a function of<br />

temperature over chemically derived graphene and noble NP/GS.<br />

where I is charge or discharge current, Δt is the time for a full<br />

charge or discharge, m indicates the mass of the active material,<br />

and ΔV represents the voltage change after a full charge or<br />

discharge. As shown in the Figure 5c, there are almost no<br />

voltage drops observed; this indicates that the electrodes can<br />

carry out fast charge/discharge under the experimental current<br />

density and show good rate capability. 29,30 Moreover, lifecycle<br />

is one of the most important performance indexes for<br />

supercapacitor. The TX-C-GS and TX-NC-GS present the<br />

feature of good cycling performance (as shown in Figure 5d)<br />

tested using the galvanostatic charge−discharge technique. The<br />

capacitance of them basically remains invariant during the 1000<br />

cycles under the operation current density of 100 mA/g after<br />

the unstable cycles at the beginning.<br />

It can be found that the specific capacitance of TX-C-GS is<br />

higher than that of TX-NC-GS at any specified discharge<br />

current density used in the present experiment. This may be<br />

attributed to the high-quality of graphene sheets <strong>from</strong> TX-C-<br />

GS, which has more open structures and higher BET specific<br />

surface area (306 vs 135 m 2 /g) as well as higher conductivity<br />

compared with the TX-NC-GS. These features can significantly<br />

improve the electrical double layer capacitance; thus, TX-G-GS<br />

presented a higher specific capacitance and better electrochemical<br />

performance.<br />

Recently, graphene nanocomposites are receiving tremendously<br />

increasing attention because they possess much more<br />

novel properties compared with pure graphene and therefore<br />

enable more promising potential applications. It is easy to<br />

imagine that the GS can be used as catalyst support due to its<br />

unique two-dimensional basal network plane structure and high<br />

surface area, as well as the presence of surface functional groups<br />

in chemically derived graphene. Noble metal nanoparticles (Pt,<br />

Ru, and PtRu) decorated coal-based graphene sheet (TX-C-<br />

GS) catalysts were successfully fabricated via impregnation and<br />

DBD plasma technique. The loading amounts of Pt, Ru, and<br />

5191<br />

PtRu were kept constant (∼3.5 wt %), and the atomic ratio of<br />

Pt:Ru in PtRu/GS was 3:1.<br />

From the TEM images (Figure 6a−c), it can be clearly seen<br />

that the exfoliated GS was decorated by uniform nanoparticles<br />

less than 2 nm in size. It has been demonstrated that metal<br />

nanoparticles can be deposited on both sides of GS due to their<br />

large surface area and unique basal 2D planar structure, and<br />

furthermore, the attachment of nanoparticles onto the GS<br />

surface can significantly prevent their aggregation and<br />

restacking. 31−33 The highly dispersed metal nanoparticles on<br />

supports with large surface area are favorable to promote their<br />

interfacial contact with the other reactant molecules; therefore,<br />

they have potential advantages in catalytic activity and sensor<br />

sensitivity. 34 The XRD patterns of the NP/GS composites are<br />

shown in Figure 6d. The typical diffraction peaks of the facecentered<br />

cubic (fcc) Pt lattice (111) and (200) in the assynthesized<br />

Pt/GS composites can be clearly observed;<br />

however, the peaks of Ru (Ru/GS) and PtRu alloy (PtRu/<br />

GS) cannot be recognizable in XRD diffraction due to the small<br />

size and low loading. 35,36<br />

Selective catalytic reduction (4NO + 4NH 3 +O 2 =4N 2 +<br />

6H 2O) has been wildly used for the removal of NO in the flue<br />

gas <strong>from</strong> coal combustion. 37 The NH 3−SCR reaction activities<br />

of GS, Pt/GS, Ru/GS, and PtRu/GS are shown in Figure 6e. It<br />

can be seen that GS itself does not show any catalytic activity in<br />

SCR, while high catalytic activities appeared after the<br />

decoration of the graphene sheets with Pt and PtRu<br />

nanoparticles. Although Ru/GS shows negligible catalytic<br />

activity under the present reaction condition, PtRu/GS<br />

bimetallic alloy catalyst shows the highest catalytic activity at<br />

low temperature among the obtained graphene-based catalysts,<br />

which is similar to the performance of the electrocatalytic<br />

reaction in the fuel cell. 38 The light-off temperature (the<br />

temperature at which the conversion of NO reaches 50%) for<br />

this reaction is as low as 150 °C under present experiment<br />

conditions, and more than 90% NO conversion can be achieved<br />

dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192


Energy & Fuels Article<br />

at 165 °C. Moreover, the reaction temperature window of the<br />

alloy catalyst is relatively wide, ranging <strong>from</strong> 150 to 230 °C with<br />

80% NO x conversion. The excellent activity for PtRu/GS<br />

catalyst is possibly due to the synergistic effect between Pt and<br />

Ru caused by the Ru that entered into the lattice of Pt in the<br />

PtRu alloy.<br />

4. CONCLUSION<br />

We demonstrate that coal-derived graphene can be successfully<br />

prepared <strong>from</strong> graphitized coal with the graphite oxide pathway<br />

combined with H 2 DBD plasma technique. Fe can serve as an<br />

effective catalyst to promote the graphitization of coal, which is<br />

critical to the structures, quality, and properties of the final<br />

product of coal-derived graphene. Noble metallic nanoparticles<br />

as small as 2 nm decorated on the surface of GS were<br />

successfully fabricated with high uniformity and dispersity via<br />

the impregnation and H 2 DBD plasma technique. The<br />

electrochemical performance of the produced graphene <strong>from</strong><br />

coal presents a mild capacitance and excellent electrochemical<br />

stability in KOH electrolyte. The noble metal/graphene<br />

composites can be used as catalyst, of which PtRu/GS show<br />

high catalytic activity and wide operating temperature window,<br />

as evidenced in SCR of NO x with ammonia under atmospheric<br />

pressure. Our synthesis strategy may lead to an alternative<br />

approach to the preparation of graphene and graphene-based<br />

composites with novel structures and various applications.<br />

■ AUTHOR<br />

INFORMATION<br />

Corresponding Author<br />

*E-mail: zbzhao@dlut.edu.cn (Z.Z.), jqiu@dlut.edu.cn (J.Q.).<br />

Notes<br />

The authors declare no competing financial interest.<br />

■ ACKNOWLEDGMENTS<br />

This work was partly supported by the NSFC (Nos. 51072028,<br />

20876026, 20836002, 20725619) and the Fundamental<br />

Research Funds for the Central Universities (no. DUT<br />

11ZD120).<br />

■ REFERENCES<br />

(1) Saha, B.; Chingombe, P.; Wakeman, R. J. Carbon 2005, 43 (15),<br />

3132.<br />

(2) Qiu, J. S.; Li, Y. F.; Wang, Y. P.; Liang, C. H.; Wang, T. H.; Wang,<br />

D. Carbon 2003, 41 (4), 767.<br />

(3) Song, C. S.; Schobert, H. H. Fuel 1996, 75 (6), 724.<br />

(4) Pang, L. S. K.; Vassallo, A. M.; Wilson, M. A. Nature 1991, 352<br />

(6335), 480.<br />

(5) Pang, L. S. K.; Wilson, M. A. Energy Fuels 1993, 7 (3), 436.<br />

(6) Qiu, J. S.; Zhang, F.; Zhou, Y.; Han, H. M.; Hu, D. S.; Tsang, S.<br />

C.; Harris, P. J. F. Fuel 2002, 81 (11−12), 1509.<br />

(7) Geim, A. K.; Novoselov, K. S. Nat. Mater. 2007, 6 (3), 183.<br />

(8) Li, D.; Kaner, R. B. Science 2008, 320 (5880), 1170.<br />

(9) Stankovich, S.; Dikin, D. A.; Dommett, G. H. B.; Kohlhaas, K. M.;<br />

Zimney, E. J.; Stach, E. A.; Piner, R. D.; Nguyen, S. T.; Ruoff, R. S.<br />

Nature 2006, 442 (7100), 282.<br />

(10) Nguyen, S. T.; Stankovich, S.; Dikin, D. A.; Piner, R. D.;<br />

Kohlhaas, K. A.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Ruoff, R. S. Carbon<br />

2007, 45 (7), 1558.<br />

(11) Li, X. S.; Cai, W. W.; An, J. H.; Kim, S.; Nah, J.; Yang, D. X.;<br />

Piner, R.; Velamakanni, A.; Jung, I.; Tutuc, E.; Banerjee, S. K.;<br />

Colombo, L.; Ruoff, R. S. Science 2009, 324 (5932), 1312.<br />

(12) Kim, K. S.; Zhao, Y.; Jang, H.; Lee, S. Y.; Kim, J. M.; Kim, K. S.;<br />

Ahn, J. H.; Kim, P.; Choi, J. Y.; Hong, B. H. Nature 2009, 457 (7230),<br />

706.<br />

5192<br />

(13) Sun, Z. Z.; Yan, Z.; Yao, J.; Beitler, E.; Zhu, Y.; Tour, J. M.<br />

Nature 2010, 468 (7323), 549.<br />

(14) Ruan, G. D.; Sun, Z. Z.; Peng, Z. W.; Tour, J. M. Acs Nano<br />

2011, 5 (9), 7601.<br />

(15) Wertz, D. L.; Bissell, M. Energy Fuels 1994, 8 (3), 613.<br />

(16) Saikia, B. K.; Boruah, R. K.; Gogoi, P. K. J. Chem. Sci. 2009, 121<br />

(1), 103.<br />

(17) Aso, H.; Matsuoka, K.; Sharma, A.; Tomita, A. Energy Fuels<br />

2004, 18 (5), 1309.<br />

(18) Hummers, W. S.; Offeman, R. E. J. Am. Chem. Soc. 1958, 80 (6),<br />

1339.<br />

(19) Zhou, Q.; Zhao, Z. B.; Chen, Y. S.; Hu, H.; Qiu, J. S. J. Mater.<br />

Chem. 2012, 22 (13), 6061.<br />

(20) Xu, W. Y.; Wang, X. Z.; Zhou, Q.; Meng, B.; Zhao, J. T.; Qiu, J.<br />

S.; Gogotsi, Y. J. Mater. Chem. 2012, 22 (29), 14364.<br />

(21) Sonibare, O. O.; Haeger, T.; Foley, S. F. Energy 2010, 35 (12),<br />

5347.<br />

(22) Schniepp, H. C.; Li, J. L.; McAllister, M. J.; Sai, H.; Herrera-<br />

Alonso, M.; Adamson, D. H.; Prud’homme, R. K.; Car, R.; Saville, D.<br />

A.; Aksay, I. A. J. Phys. Chem. B 2006, 110 (17), 8535.<br />

(23) Ferrari, A. C.; Meyer, J. C.; Scardaci, V.; Casiraghi, C.; Lazzeri,<br />

M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K. S.; Roth, S.; Geim,<br />

A. K. Phys. Rev. Lett. 2006, 97 (18), 187401.<br />

(24) Ramaprabhu, S.; Eswaraiah, V.; Aravind, S. S. J. J. Mater. Chem.<br />

2011, 21 (19), 6800.<br />

(25) Tuinstra, F.; Koenig, J. L. J. Chem. Phys. 1970, 53 (3), 1126.<br />

(26) Wu, Z. S.; Ren, W. C.; Gao, L. B.; Liu, B. L.; Jiang, C. B.; Cheng,<br />

H. M. Carbon 2009, 47 (2), 493.<br />

(27) Liu, Y. Z.; Li, Y. F.; Zhong, M.; Yang, Y. G.; Wen, Y. F.; Wang,<br />

M. Z. J. Mater. Chem. 2011, 21 (39), 15449.<br />

(28) Wang, H. L.; Casalongue, H. S.; Liang, Y. Y.; Dai, H. J. J. Am.<br />

Chem. Soc. 2010, 132 (21), 7472.<br />

(29) Li, Y. G.; Tan, B.; Wu, Y. Y. Nano Lett. 2008, 8 (1), 265.<br />

(30) Lin, R.; Taberna, P. L.; Chmiola, J.; Guay, D.; Gogotsi, Y.;<br />

Simon, P. J. Electrochem. Soc. 2009, 156 (1), A7.<br />

(31) Wang, X.; Xu, C.; Zhu, J. W. J. Phys. Chem. C 2008, 112 (50),<br />

19841.<br />

(32) Samulski, E. T.; Si, Y. C. Chem. Mater. 2008, 20 (21), 6792.<br />

(33) Yuge, R.; Zhang, M. F.; Tomonari, M.; Yoshitake, T.; Iijima, S.;<br />

Yudasaka, M. ACS Nano 2008, 2 (9), 1865.<br />

(34) Xing, Y. C. J. Phys. Chem. B 2004, 108 (50), 19255.<br />

(35) Abu Bakar, N. H. H.; Bettahar, M. M.; Abu Bakar, M.;<br />

Monteverdi, S.; Ismail, J.; Alnot, M. J. Catal. 2009, 265 (1), 63.<br />

(36) Bock, C.; Paquet, C.; Couillard, M.; Botton, G. A.; MacDougall,<br />

B. R. J. Am. Chem. Soc. 2004, 126 (25), 8028.<br />

(37) Galvez, M. E.; Lazaro, M. J.; Moliner, R. Catal. Today 2005, 102,<br />

142.<br />

(38) Liu, Z. L.; Ling, X. Y.; Su, X. D.; Lee, J. Y. J. Phys. Chem. B 2004,<br />

108 (24), 8234.<br />

dx.doi.org/10.1021/ef300919d | Energy Fuels 2012, 26, 5186−5192

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!