Interlinked Porous Carbon Nanoflakes Derived from Hydrolyzate

Dec 13, 2016 - (11, 32) The normalized capacitance was obtained from the ratio of .... of less than 10 nm and the thin nanoflake structures (4.6 nm) o...
8 downloads 0 Views 5MB Size
Research Article pubs.acs.org/journal/ascecg

Interlinked Porous Carbon Nanoflakes Derived from Hydrolyzate Residue during Cellulosic Bioethanol Production for Ultrahigh-Rate Supercapacitors in Nonaqueous Electrolytes Weiqian Tian, Qiuming Gao,* and Weiwei Qian

Downloaded via KAOHSIUNG MEDICAL UNIV on June 30, 2018 at 16:33:31 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Key Laboratory of Bio-inspired Smart Interfacial Science and Technology of Ministry of Education, Beijing Key Laboratory of Bio-inspired Energy Materials and Devices, School of Chemistry and Environment, Beihang University, No 37 Xueyuan Road, Beijing 100191, China S Supporting Information *

ABSTRACT: The rapid development of cellulosic bioethanol has produced a mass of hydrolysate residue as byproducts during the pretreatment process of lignocellulose. The high value-added utilization of hydrolysate residue plays a key role in the economic viability of large-scale/green industrial production of lignocellulose into bioethanol. Here the hydrolyzate residue was exploited as a carbon precursor for the fabrication of an interlinked graphitized porous carbon nanoflake (GPCNF) by an in situ carbonization−activation process. The final GPCNF presents an optimum integration of a large surface area of 2026 m2 g−1, bimodal pore systems (86% of mesopore volume), and an excellent electric conductivity of 5.4 S cm−1. These characteristics favorably endow that the GPCNF is ideally suited for nonaqueous electrolyte-based supercapacitor applications. In organic electrolyte of 1 M TEA BF4/AN, the GPCNF-based supercapacitor exhibits a high rate capability of 74% initial capacitance at a high current density of 100 A g−1. Notably, in an ionic liquid electrolyte of EMI TFSI the GPCNF-based supercapacitor displays an integrated high energy−power property at an energy density of 37.7 Wh kg−1 corresponding to a high power density of 77.5 kW kg−1, which puts the GPCNF on the Ragone plot among the best energy−power characteristics in the reported twodimensional biomass-derived carbon electrodes for nonaqueous electrolyte-based supercapcitors. KEYWORDS: Porous carbon nanoflakes, Supercapacitors, Ultrahigh-rate, Nonaqueous electrolytes, Hydrolyzate residue, Cellulosic bioethanol



INTRODUCTION

that the pretreatment process, e.g., alkaline peroxide, acid, ionic liquid or enzymatic hydrolysis, liquid hot water, mechanical comminution, and steam/ammonia explosion, is the crucial step to break down the lignocellulosic structures into high-content accessible cellulose for efficient bioethanol production.6,7 Compared to other technologies, the alkaline peroxide hydrolysis has high degree of delignification and slight loss of cellulose. This process hardly generates the fermentation inhibitors of furfurals7 and also is an efficient/economical approach.3 During this pretreatment, the hemicellulose is decomposed into soluble

Cellulosic bioethanol as one of most promising substitutes for imminently depleted fossil fuels has been extensively explored from renewable lignocellulose via biorefinery processes.1−3 About 442 billion L y−1 of bioethanol can be manufactured from the world’s lignocellulosic wastes, which is much more than 90 billion L of the actual world bioethanol production in 2014.4 Cellulosic bioethanol can also avoid the food insecurity caused by cereal crop-dependent bioethanol production (including that from corn, sugar cane, wheat, rice, sorghum, etc.). The efficient production of lignocellulose to bioethanol usually required four stages, such as pretreatment of lignocellulose feedstock, enzymatic hydrolysis of cellulose into sugars, microbial fermentation of sugars to bioethanol, and distillation.5 Note © 2016 American Chemical Society

Received: June 20, 2016 Revised: December 6, 2016 Published: December 13, 2016 1297

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering

until neutral pH. Finally, the resulting sample was dried at 100 °C overnight and was denoted as GPCNF. The yield rate of GPCNF is 24% based on the weight of HR. The synthesis condition such as the mass ratio of HR/ZnCl2, the concentration of FeCl2 solution, and the carbonization−activation temperatures were optimized based on the electrochemical performance of the obtained samples (Figure S2). Characterization of GPCNF. Field emission scanning electron microscopy (FESEM, S-3400N, Hitachi) and field emission transmission electron microscopy (FETEM, JEM2100F, JEOL) were used to observe the morphologies of the samples. Atomic force microscope (AFM) was conducted on the Multimode Nanoscope IIIa, Bruker, and the ethanol suspension of sample was sonicated to obtain a highly dilute solution before drop-casting onto a silicon substrate. Powder X-ray diffractometer (XRD) was performed on the Shimadzu X-6000. Raman spectrum was tested by using a 514 nm excitation on a LabRAM HR800 apparatus. The electrical conductivity of the samples was measured using a standard four-point-probe system (Jandel-RM3) with a pindistance of about 1 mm. The powder materials were pressed between two plungers into a hollow stainless-steel cylinder at a pressure of 15 MPa for 30 s at room temperature. The obtained disclike samples have a diameter of 13 mm and a thickness of 0.8 mm. The average values out of three measurements at different positions were taken within an error range of ±1%. X-ray photoelectron spectroscopy (XPS) of Thermo Escalab 250 was used to analyze the surface chemical compositions. The bulk elemental composition of samples was tested on the elemental analyzer Vario EL III, Elementar. N2 adsorption−desorption isotherm was conducted on the Quantachrome Instruments Autosorb-1 (77 K). The specific surface area is determined by the multipoints Brumauer− Emmett−Teller (BET) method. The nonlocal density functional theory (NLDFT) with slit model is chosen to determine the pore size distribution. Fabrication of GPCNF-Supercapacitor (GPCNF-SC). The electrodes were prepared by pasting a slurry containing 95 wt % GPCNF and 5 wt % PTFE binder without conductive additive onto a conducting carboncoated aluminum foil as the current collector. The as-prepared electrodes were dried at 100 °C for 12 h in vacuum oven and then were pressed at 10 MPa with a final mass loading around 5 mg cm−2. The symmetrical supercapacitors were assembled in an argon-filled glovebox of 800 °C) was

slits between the neighboring nanoflakes are also shown, which can effectively improve ion-transport kinetics, decrease iondiffusion resistances, and deminish the ion-sieving effect during the electrochemical process.33,34 The high-resolution FETEM image (Figure 1c) highlights partially ordered graphitized structures in GPCNF with a thin thickness. Moreover, some small worm-like pores can be observed in Figure 1c. The AFM image (Figure 1d) shows some single nanoflakes as a result of the ultrasonic fragmentation of interlinked nanoflake structures. The corresponding height analyses along the lines marked in the AFM image exhibit a uniform thickness of around 4.6 nm across the carbon nanoflakes (Figure 1e), which are consistent with the SEM and TEM observations. Note that the interlinked porous graphitized carbon nanoflake structures constructed successfully can be ascribed to the combined effect of the composition and structure of the HR precursor and the in situ carbonization−activation process. The fabrication process of GPCNF is illustrated in Scheme 1. The poplar microfibril is composed of the cellulose tightly embedded in a matrix of the lignin and hemicellulose.4 The alkaline peroxide pretreatment disrupted the complicated microfibril structures and unwound cellulose from lignin/hemicellulose matrix.7 The isolated cellulose was used to the bioethanol production, while the degraded lignin and hemicellulose were dissolved in hydrolysate thus generating vast HR wastes (Figure S1). The FESEM image of HR (Figure S2) exhibits a particlelike structure. The particlelike HR consists of 46.6 wt % hemicellulose-derived saccharides (xylose 86.5%, Glu 5.3%, Rha 1.7%, Ara 1.1%, and Gal 1.4%) and 53.4 wt % degraded lignin based on the National Renewable Energy Laboratory (NREL, Golden, CO) analytical methods.3 The degraded lignin and hemicellulose of HR have a bulk elemental composition of 52.36 wt % carbon, 41.21 wt % oxygen, and 6.43 wt % hydrogen, which contain abundantly polar oxygen-containing functional groups such as hydroxyl, ether, and carbonyl groups. During the in situ carbonization−activation 1299

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering

Figure 2. (a) XRD pattern, (b) Raman spectrogram, (c) XPS, (d) C 1s and O 1s spectrum of XPS, (e) N2 adsorption−desorption isotherm, and (f) pore size distribution of GPCNF. (e inset) Cumulative pore volume of GPCNF.

angle region reveals a high porosity in GPCNF, which is consistent with the FETEM analysis. Raman spectrum (Figure 2b) of GPCNF displays two typical peak, assigned to D-band (1350 cm−1) for defective sp3-phase and G-band (1587 cm−1) for ordered sp2-phase,11,24 respectively. The second order 2D peak originating from the second-order zone boundary phonons,47 was also observed from the Raman spectrum, which further attests the formation of local few-layer graphene structures.40,48 The integral ratio IG/ID of 1.13 is well comparable with that of the functionalized graphene (1.07− 1.20)20 and is higher than activated carbon Norit (0.52).12 The outcome further demonstrates a wide graphitic domain in the GPCNF. An excellent electric conductivity of 5.4 S cm−1 was obtained by four-point probe conductivity measurements, which is higher than that of commercial activated carbons for supercapacitor (e.g., Supra 50, 4.3 S cm−1).11 XPS (Figure 2c) was used to offer insight into surface chemical composition of GPCNF. The result manifests 94.6 at % carbon and 5.2 at % oxygen with a high C/O ratio of 18, which is consistent with the bulk chemical compositions measured by combustion elemental analysis of 94.12 wt % carbon, 5.74 wt % oxygen, and 0.14 wt % hydrogen. The outcome is due to XPS analysis only affording the surface composition of less than 10 nm and the thin nanoflake structures (4.6 nm) of GPCNF.34 From the deconvolution of high resolution C 1s spectrum of XPS

needed to build the in situ self-template and simultaneously grow the carbon nanoflakes. The crystal structures of GPCNF were characterized by XRD and Raman spectroscopy. The XRD pattern (Figure 2a) shows two characteristic peak, corresponding to the (002) reflection at 20.3° stemming from the stacking intergraphene layers and the (100) reflection at 43.6° related to ordered graphitic domains, respectively.28,43 Based on the (002) reflection peak, a layer-tolayer distance (d-spacing, d002) of 0.44 nm was obtained,14 which is larger than that of graphite (0.34 nm) and even functionalized graphene (0.37 nm).20 The high interlayer spacing may result from the formation of graphite lattice through nucleation or growth from biomass-based precursors instead of reordering of the existing graphene layers.44 The large d-spacing leads to a loose structure with a high accessibility of electrolyte ions into the nanoflake surfaces and boundaries which may contribute to a high capacitance performance.45 According to the full width at half-maximum values of the (002) peak, the average domain thickness (Lc) in the graphitic lattice can be calculated to 3.39 nm, which indicates the carbon nanoflakes consisting of 7−8 stacked graphene-like sheets (i.e., Lc/d002 ≈ 7.7).12,46 Based on the (100) peak, the average domain width (La) is deduced to 9.2 nm. The detail calculation could be found in the Supporting Information. Note that an obvious rise of the intensity in the low1300

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering

Figure 3. (a) CV curves of GPCNF-SC and HR-SC in EMI TFSI electrolyte at 50 mV s−1, (b) CV curves of GPCNF-SC in EMI TFSI electrolyte at 200 mV s−1 tested at different temperatures, (c) CV curves of GPCNF-SC under different scan rates in 1 M TEA BF4/AN, and (d) EMI TFSI electrolytes.

Figure 4. Charge/discharge voltage profiles of GPCNF-SC in (a) 1 M TEA BF4/AN and (b) EMI TFSI electrolytes at different current densities. (c) Gravimetric-normalized specific capacitances at different current densities. (d) Specific capacitance versus square root of charge/discharge time of GPCNF-SC in 1 M TEA BF4/AN and EMI TFSI electrolytes.

(Figure 2d), a high percentage of sp2 C (284.7 eV, 72%) peak was displayed and several small peaks of sp3 C (285.7 eV), C−O (286.8 eV), CO (288.4 eV) and OC−O (289.8 eV) were also observed at high binding energy region.49 The sp2/sp3 ratio is up to 4.8, which indicates a mass of conjugated graphitic

domain constructed in GPCNF in good agreement with the FETEM, XRD, and Raman analyses.50 The O 1s spectrum of XPS is also shown in Figure 2d. The high ratios of sp2/sp3 and C/ O in GPCNF are in favor of the electrochemical process in organic or ionic-liquid electrolytes.51 1301

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering

Figure 5. (a) Nyquist plots and (b) frequency response versus normalized capacitance of GPCNF-SC in 1 M TEA BF4/AN and EMI TFSI electrolytes. (c) Cycling stability of GPCNF-SC in 1 M TEA BF4/AN and EMI TFSI electrolytes at 5 A g−1. (a inset) Expanded view in the high frequency region.

comparable to the reported graphene-based electrode.56 The result can be attributed to the combined effect of high mesopore volume, fast conducting, and thin interlinked nanoflake structures of GPCNF electrodes, and the low viscosity (the melting point of EMI TFSI is about −15 °C), high ionic conductivity, and small ionic size (the maximum dimension of cation size is 0.76 nm and that of the anion is 0.79 nm50) of EMI TFSI electrolyte. Such effects result in a low ion sieving effect and small diffusion resistance even at low operating temperatures.50 Figure 3c and d highlight the CV curve of GPCNF-SC in 1 M TEA BF4/AN and EMI TFSI electrolyte (tested at 60 °C), respectively. The CV curves display nearly rectangular shapes and excellent capacitance retentions over a wide range of scan rates from 100 to 1000 mV s−1 in both electrolyte systems, thus giving a high rate capability. The charge/discharge voltage profile of GPCNF-SC in 1 M TEA BF4/AN and EMI TFSI electrolyte is depicted in Figure 4a and b, respectively. Both electrolyte systems exhibit linearly symmetric triangular profiles with small voltage drops at a wide current density range from 1 to 100 A g−1. The outcome further demonstrates the efficient formation of a capacitive behavior with an excellent electrochemical reversibility and a small equivalent series resistance (ESR) in GPCNF-SC.31,57 The rate performances of GPCNF-SC in both electrolytes are shown in Figure 4c, which were calculated from the upper part of discharge curves after voltage drops. The specific capacitance in 1 M TEA BF4/AN electrolyte is up to 142 F g−1 at a current density of 1 A g−1 (areanormalized capacitance of 7.0 μF cm−2), giving a high rate capability of 74% initial capacitance at a high current density of 100 A g−1. Additionally, the EMI TFSI electrolyte system presents a high capacitance of 166 F g−1 at 1 A g−1 (8.2 μF cm−2), highlighting a 68% rate performance at 100 A g−1. These values are superior to most of the corresponding values of the reported 2D biomass-derived nanocarbons in nonaqueous electrolytes (Table S1). The charging/discharging kinetic analysis (Figure

N2 adsorption−desorption isotherm of GPCNF displays a hybrid I/IV type with a typical Langmuir hysteresis (Figure 2e), indicating a well-defined hierarchical pore structure.48 The NLDFT further confirms the hierarchical pore size distributions with bimodal pore system (Figure 2f), which is mainly centered at 1.3 and 4.2 nm, respectively. The former may be related to the small wormlike pores on the carbon nanoflakes generating from the ZnCl2 activation process.52 The latter could result from the interagglomerate pores between the interlinked carbon nanoflakes, which accounts for 86% of total pore volume (1.93 cm3 g−1). A high ratio of mesopore volume to total pore volume was obtained, which is consistent with the FESEM and FETEM analyses. The specific surface area is up to 2026 m2 g−1 based on the BET method. The integration of hierarchical pore size distribution, high mesopore volume and large specific surface area is beneficial for a high-rate electrode performance in nonaqueous electrolytes.24 The supercapacitive performances of GPCNF-supercapacitor (GPCNF-SC) were evaluated by using the most recommended reliable industry method/standard of two-electrode symmetric systems.53−55 CV, galvanostatic charge/discharge, and EIS technologies were employed. The most popular organic electrolyte of 1 M TEA BF4/AN and ionic liquid of EMI TFSI were applied as electrolyte, respectively. Figure 3a depicts the CV curves of our samples in EMI TFSI electrolyte. The CV curve of GPCNF has an enhanced capacitance response with a rectangular shape compared with that of HR tested as baseline. The outcome manifests a nearly ideal capacitive behavior in GPCNF electrode even without electroactive additives used. The operating temperature of EMI TFSI electrolyte system was evaluated under 0, 25, 60, and 100 °C at different scan rates (Figures 3b and S4). The CV curves reveal good capacitive responses with quasi-rectangular shapes at ambient temperature of 25 °C and even at a low temperature of 0 °C. The good low-temperature performance is noteworthy and is 1302

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering

Figure 6. (a) Ragone plots of GPCNF-SC based on active electrode materials in 1 M TEA BF4/AN and EMI TFSI electrolytes and the performance comparison of GPCNF-SC with those of state-of-the-art 2D biomass-derived carbon electrodes in nonaqueous electrolytes. (b) Ragone plots of packaged GPCNF-SC based on total-device weight and the performance comparison of GPCNF-SC with several standard devices (data used from ref 48).

ion size and viscosity of the electrolytes, but both two systems give the high-level values. Such outcomes further indicate that the GPCNF electrode has excellent ion/electron transport and high charge/discharge rate capabilities. In addition, Figure 5c expresses the cyclic life for GPCNF-SC in both electrolytes, such as 96% capacitance retention in 1 M TEA BF4/AN system and 93% in EMI TFSI system, respectively, after 10 000 cycles at 5 A g−1. In both systems the specific capacitance after cycle slightly faded due to the partial degradation of electrolyte on the oxygencontaining (5.74 wt %) electrode interfaces.24 However, the decay of the specific capacitances is both less than 10% of their initial values, which could be mainly ascribed to the thin interlinked nanoflakes with high ratios of sp2/sp3 and C/O. The results are comparable with state-of-the-art predecessors in organic and ionic liquid systems (Table S1). Figure 6a shows the Ragone plots of GPCNF-SC, which are utilized to assess the energy-power characteristics. The maximum energy density of 70.5 Wh kg−1 was obtained in the EMI TFSI electrolyte system based on the active electrode materials, which is due to the higher operating voltage window of 3.5 V than that of the organic electrolyte system (2.5 V). The high-rate performance leads to an excellent energy maintenance of 37.7 Wh kg−1 at an ultrahigh power density of 77.5 kW kg−1 with a current-drain time of less than 3.6 s in the EMI TFSI electrolyte. We also compare the energy-power densities of GPCNF-SC to those of state-of-the-art 2D biomass-derived carbon electrodes.9,12,14,24,25,28 Comparatively the GPCNF is among the best in the reported 2D biomass-derived carbon electrodes for nonaqueous electrolyte supercapcitors. For practical packaged supercapacitor application, the energy-power densities based on total-device weight can be deduced by a fourth of those based on the electrode materials.18,24,31 These properties of the packaged device are noteworthy in that high energy densities of ∼9−18 Wh kg−1 could be simultaneously obtained along with high power densities in the EMI TFSI system (Figure 6b), thus making our devices potentially competitive against the lead-acid batteries as well as nickel metahydride batteries (Ni-MH).48

4d) further confirms the excellent electrochemical reversibility and high-rate properties of GPCNF-SC. Note that the electrode specific capacitance could be expressed the formula of Celect = k1 + k2t1/2, where k1 denotes a rate-independent component assigned to reversible capacitance, and k2 represents a diffusion-limited component dominated by the charge/discharge time t.48 Where the t equals to 0, the k1 can be deduced from the intercept with the vertical axis. Notably, the k1 dominates in GPCNF electrode, exceeding ∼120 F g−1 in 1 M TEA BF4/AN and ∼130 F g−1 in EMI TFSI electrolyte. The outcome is ascribed to the special thin interlinked nanoflake structures with high mesopore volume that can accelerate ion diffusion kinetics. The EIS technique was employed to explore the ionic/ electronic transport pathways in GPCNF electrode. Figure 5a depicts the Nyquist plots in two electrolyte systems. The almost vertical lines in low-frequency range highlight the pure capacitive behavior.17 The short Warburg regions at high-medium frequencies indicate a good ion diffusion efficiency.58 By extrapolating the vertical line to the real axis, the small resistance of 1.6 and 2.9 Ω was got for the organic and ionic liquid system, respectively. The small values mainly result from the small ESR and charge transfer resistance (Rct).32 The ESR acquired from the first intercept with the real axis in high-frequency region is 0.3 Ω for the organic system and 0.5 Ω for the ionic liquid system, as further demonstrated the observation of the CV and charge/ discharge analyses. The Rct obtained from the diameter of the semicircle in the high- to midfrequency region is of 0.5 and 1.2 Ω for the organic and ionic liquid system, respectively. The Rct could be mainly caused by the interface resistance to the adsorption/desorption of electrolyte ions during the formation of electrical double-layers at the electrode−electrolyte interfaces,14,55,59 and partly stemmed from limited surface Faradaic reaction of oxygen groups.10,60 However, the rather small Rct values indicate an excellent interface accessibility of electrolyte ions into our GPCNF electrode,14,55 which might primarily be attributed to the unique interlinked graphitized porous nanoflake structures with large ion-accessible surface area and high mesopore volume for the efficient ion transport. Figure 5b presents the Bode plots of frequency responses of normalized capacitance. The relaxation time constant τo of 0.90 and 1.15 s was respectively acquired for the organic and ionic liquid system, which are considerably smaller than those of the commercial active carbon electrode (∼10 s),17,61 and activated graphenebased electrodes (1.67 s).31 Noted that the difference of ESR and τo in two electrolyte systems is ascribed to the combined effect of



CONCLUSIONS The GPCNF was successfully created from hydrolysate residue byproducts during lignocellulose bioethanol production through an in situ carbonization−activation process. The obtained GPCNF has an interlinked graphitized porous nanoflake framework with a large surface area of 2026 m2 g−1, bimodal pore system (86% of mesopore volume), and an excellent electric 1303

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering conductivity of 5.4 S cm−1. Due to these special structural merits, the GPCNF-based supercapacitor displays the ultrahigh-rate performances in nonaqueous electrolytes, such as a 74% initial capacitance in 1 M TEA BF4/AN organic electrolyte, and that of 68% in EMI TFSI ionic liquid electrolyte at a high current density of 100 A g−1. Notably, the GPCNF-based supercapacitor in the ionic liquid system exhibits an integrated high energy-power property at the energy density of 37.7 Wh kg−1 corresponding to the high power density of 77.5 kW kg−1. The outcome indicates that our devices can be capable of outputting high energy as the nickel metahydride batteries and, simultaneously, high power as the electrochemical capacitors. Thus, this novel precursor− synthesis route provides a sustainable/low-cost carbon-based electrode material, i.e., GPCNF, for high-performance nonaqueous electrolyte-based supercapacitor with an excellent energy−power combination.



bioethanol production accompanying with high valuable products. Bioresour. Technol. 2016, 199, 68−75. (6) Falco, C.; Sieben, J. M.; Brun, N.; Sevilla, M.; van der Mauelen, T.; Morallon, E.; Cazorla-Amoros, D.; Titirici, M. M. Hydrothermal carbons from hemicellulose-derived aqueous hydrolysis products as electrode materials for supercapacitors. ChemSusChem 2013, 6, 374− 382. (7) Zhang, L.; You, T.; Zhang, L.; Yang, H.; Xu, F. Enhanced fermentability of poplar by combination of alkaline peroxide pretreatment and semi-simultaneous saccharification and fermentation. Bioresour. Technol. 2014, 164, 292−298. (8) Zheng, X.; Luo, J.; Lv, W.; Wang, D.-W.; Yang, Q.-H. Twodimensional porous carbon: synthesis and ion-transport properties. Adv. Mater. 2015, 27, 5388−5395. (9) Fan, X.; Yu, C.; Yang, J.; Ling, Z.; Hu, C.; Zhang, M.; Qiu, J. A layered-nanospace-confinement strategy for the synthesis of twodimensional porous carbon nanosheets for high-rate performance supercapacitors. Adv. Energy Mater. 2015, 5, 1401761. (10) Ling, Z.; Wang, Z.; Zhang, M.; Yu, C.; Wang, G.; Dong, Y.; Liu, S.; Wang, Y.; Qiu, J. Sustainable synthesis and assembly of biomass-derived B/N co-doped carbon nanosheets with ultrahigh aspect ratio for highperformance supercapacitors. Adv. Funct. Mater. 2016, 26, 111−119. (11) Sevilla, M.; Fuertes, A. B. Direct synthesis of highly porous interconnected carbon nanosheets and their application as highperformance supercapacitors. ACS Nano 2014, 8, 5069−5078. (12) Wang, H.; Xu, Z.; Kohandehghan, A.; Li, Z.; Cui, K.; Tan, X.; Stephenson, T. J.; King’ondu, C. K.; Holt, C. M.; Olsen, B. C.; Tak, J. K.; Harfield, D.; Anyia, A. O.; Mitlin, D. Interconnected carbon nanosheets derived from hemp for ultrafast supercapacitors with high energy. ACS Nano 2013, 7, 5131−5141. (13) Zheng, X.; Lv, W.; Tao, Y.; Shao, J.; Zhang, C.; Liu, D.; Luo, J.; Wang, D.-W.; Yang, Q.-H. Oriented and interlinked porous carbon nanosheets with an extraordinary capacitive performance. Chem. Mater. 2014, 26, 6896−6903. (14) Sun, L.; Tian, C.; Li, M.; Meng, X.; Wang, L.; Wang, R.; Yin, J.; Fu, H. From coconut shell to porous graphene-like nanosheets for highpower supercapacitors. J. Mater. Chem. A 2013, 1, 6462−6470. (15) Yan, J.; Wang, Q.; Wei, T.; Fan, Z. Recent advances in design and fabrication of electrochemical supercapacitors with high energy densities. Adv. Energy Mater. 2014, 4, 1300816. (16) Yang, X.; Cheng, C.; Wang, Y.; Qiu, L.; Li, D. Liquid-mediated dense integration of graphene materials for compact capacitive energy storage. Science 2013, 341, 534−537. (17) El-Kady, M. F.; Strong, V.; Dubin, S.; Kaner, R. B. Laser scribing of high-performance and flexible graphene-based electrochemical capacitors. Science 2012, 335, 1326−1330. (18) Zhu, Y.; Murali, S.; Stoller, M. D.; Ganesh, K.; Cai, W.; Ferreira, P. J.; Pirkle, A.; Wallace, R. M.; Cychosz, K. A.; Thommes, M.; Su, D.; Stach, E. A.; Ruoff, R. S. Carbon-based supercapacitors produced by activation of graphene. Science 2011, 332, 1537−1541. (19) Feng, H.; Cheng, R.; Zhao, X.; Duan, X.; Li, J. A low-temperature method to produce highly reduced graphene oxide. Nat. Commun. 2013, 4, 1539−1545. (20) Tian, W.; Gao, Q.; Tan, Y.; Zhang, Y.; Xu, J.; Li, Z.; Yang, K.; Zhu, L.; Liu, Z. Three-dimensional functionalized graphenes with systematical control over the interconnected pores and surface functional groups for high energy performance supercapacitors. Carbon 2015, 85, 351−362. (21) Ghidiu, M.; Lukatskaya, M. R.; Zhao, M.-Q.; Gogotsi, Y.; Barsoum, M. W. Conductive two-dimensional titanium carbide ’clay’ with high volumetric capacitance. Nature 2014, 516, 78−81. (22) Lukatskaya, M. R.; Mashtalir, O.; Ren, C. E.; Dall’Agnese, Y.; Rozier, P.; Taberna, P. L.; Naguib, M.; Simon, P.; Barsoum, M. W.; Gogotsi, Y. Cation intercalation and high volumetric capacitance of twodimensional titanium carbide. Science 2013, 341, 1502−1505. (23) Peng, H.; Ma, G.; Sun, K.; Mu, J.; Lei, Z. One-step preparation of ultrathin nitrogen-doped carbon nanosheets with ultrahigh pore volume for high-performance supercapacitors. J. Mater. Chem. A 2014, 2, 17297−17301.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.6b01390. Calculations of the average layer-to-layer distance, average domain thickness, and average domain width of GPCNF; the process of the cellulosic bioethanol production; the FESEM image of HR; the CV curves of GPCNF-SC in EMI TFSI electrolyte tested at different temperatures; and the comparison of the supercapacitive performance of GPCNF with the state-of-the-art 2D carbons (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected] (Q.G.). ORCID

Qiuming Gao: 0000-0002-1232-4139 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work is supported by National Basic Research Programs of China (973 Program, No. 2014CB931800), Chinese National Science Foundation (No. 21571010 and U0734002), and Chinese Aeronautic Project (No. 2013ZF51069).



REFERENCES

(1) Wargacki, A. J.; Leonard, E.; Win, M. N.; Regitsky, D. D.; Santos, C. N. S.; Kim, P. B.; Cooper, S. R.; Raisner, R. M.; Herman, A.; Sivitz, A. B.; Lakshmanaswamy, A.; Kashiyama, Y.; Baker, D.; Yoshikuni, Y. An engineered microbial platform for direct biofuel production from brown macroalgae. Science 2012, 335, 308−313. (2) Zhang, L.; Zhang, L.; Zhou, T.; Wu, Y.; Xu, F. The dual effects of lignin content on enzymatic hydrolysis using film composed of cellulose and lignin as a structure model. Bioresour. Technol. 2016, 200, 761−769. (3) Zhang, L.; You, T.; Zhang, L.; Li, M.; Xu, F. Comprehensive utilization of waste hemicelluloses during ethanol production to increase lactic acid yield: from pretreatment to fermentation. Biotechnol. Biofuels 2014, 7, 1−7. (4) Gupta, A.; Verma, J. P. Sustainable bio-ethanol production from agro-residues: A review. Renewable Sustainable Energy Rev. 2015, 41, 550−567. (5) Zhuang, X.; Wang, W.; Yu, Q.; Qi, W.; Wang, Q.; Tan, X.; Zhou, G.; Yuan, Z. Liquid hot water pretreatment of lignocellulosic biomass for 1304

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305

Research Article

ACS Sustainable Chemistry & Engineering

with extreme energy-power rivals lithium ion capacitors. Energy Environ. Sci. 2015, 8, 941−955. (44) Ding, J.; Wang, H.; Li, Z.; Kohandehghan, A.; Cui, K.; Xu, Z.; Zahiri, B.; Tan, X.; Lotfabad, E. M.; Olsen, B. C.; Mitlin, D. Carbon nanosheet frameworks derived from peat moss as high performance sodium ion battery anodes. ACS Nano 2013, 7, 11004−11015. (45) Aboutalebi, S. H.; Jalili, R.; Esrafilzadeh, D.; Salari, M.; Gholamvand, Z.; Aminorroaya Yamini, S.; Konstantinov, K.; Shepherd, R. L.; Chen, J.; Moulton, S. E.; Innis, P. C.; Minett, A. I.; Razal, J. M.; Wallace, G. G. High-performance multifunctional graphene yarns: toward wearable all-carbon energy storage textiles. ACS Nano 2014, 8, 2456−2466. (46) Wang, H.; Xu, Z.; Li, Z.; Cui, K.; Ding, J.; Kohandehghan, A.; Tan, X.; Zahiri, B.; Olsen, B. C.; Holt, C. M.; Mitlin, D. Hybrid device employing three-dimensional arrays of MnO in carbon nanosheets bridges battery−supercapacitor divide. Nano Lett. 2014, 14, 1987−1994. (47) Lee, C.-W.; Yoon, S.-B.; Kim, H.-K.; Youn, H.-C.; Han, J.; Roh, K. C.; Kim, K.-B. A two-dimensional highly ordered mesoporous carbon/ graphene nanocomposite for electrochemical double layer capacitors: effects of electrical and ionic conduction pathways. J. Mater. Chem. A 2015, 3, 2314−2322. (48) Lin, T.; Chen, I.-W.; Liu, F.; Yang, C.; Bi, H.; Xu, F.; Huang, F. Nitrogen-doped mesoporous carbon of extraordinary capacitance for electrochemical energy storage. Science 2015, 350, 1508−1513. (49) Zhao, Y.; Ran, W.; He, J.; Song, Y.; Zhang, C.; Xiong, D.-B.; Gao, F.; Wu, J.; Xia, Y. Oxygen-Rich Hierarchical porous carbon derived from artemia cyst shells with superior electrochemical performance. ACS Appl. Mater. Interfaces 2015, 7, 1132−1139. (50) Weingarth, D.; Zeiger, M.; Jäckel, N.; Aslan, M.; Feng, G.; Presser, V. Graphitization as a universal tool to tailor the potential-dependent capacitance of carbon supercapacitors. Adv. Energy Mater. 2014, 4, 1400316. (51) Sevilla, M.; Mokaya, R. Energy storage applications of activated carbons: supercapacitors and hydrogen storage. Energy Environ. Sci. 2014, 7, 1250−1280. (52) Molina-Sabio, M.; Rodríguez-Reinoso, F. Role of chemical activation in the development of carbon porosity. Colloids Surf., A 2004, 241, 15−25. (53) Gogotsi, Y.; Simon, P. True performance metrics in electrochemical energy storage. Science 2011, 334, 917−918. (54) Stoller, M. D.; Ruoff, R. S. Best practice methods for determining an electrode material’s performance for ultracapacitors. Energy Environ. Sci. 2010, 3, 1294−1301. (55) Zhang, L.; Zhang, F.; Yang, X.; Long, G.; Wu, Y.; Zhang, T.; Leng, K.; Huang, Y.; Ma, Y.; Yu, A.; et al. Porous 3D graphene-based bulk materials with exceptional high surface area and excellent conductivity for supercapacitors. Sci. Rep. 2013, 3, 1408−1416. (56) Xu, Z.; Li, Z.; Holt, C. M. B.; Tan, X.; Wang, H.; Amirkhiz, B. S.; Stephenson, T.; Mitlin, D. Electrochemical supercapacitor electrodes from sponge-like graphene nanoarchitectures with ultrahigh power density. J. Phys. Chem. Lett. 2012, 3, 2928−2933. (57) Zhao, J.; Lai, H.; Lyu, Z.; Jiang, Y.; Xie, K.; Wang, X.; Wu, Q.; Yang, L.; Jin, Z.; Ma, Y.; Liu, J.; Hu, Z. Hydrophilic hierarchical nitrogendoped carbon nanocages for ultrahigh supercapacitive performance. Adv. Mater. 2015, 27, 3541−3545. (58) Frackowiak, E.; Béguin, F. Carbon materials for the electrochemical storage of energy in capacitors. Carbon 2001, 39, 937−950. (59) Puthusseri, D.; Aravindan, V.; Madhavi, S.; Ogale, S. 3D microporous conducting carbon beehive by single step polymer carbonization for high performance supercapacitor: The magic of in situ porogen formation. Energy Environ. Sci. 2014, 7, 728−735. (60) Li, Z.; Xu, Z. W.; Wang, H. L.; Ding, J.; Zahiri, B.; Holt, C. M. B.; Tan, X. H.; Mitlin, D. Colossal pseudocapacitance in a high functionality-high surface area carbon anode doubles the energy of an asymmetric supercapacitor. Energy Environ. Sci. 2014, 7, 1708−1718. (61) Wei, L.; Sevilla, M.; Fuertes, A. B.; Mokaya, R.; Yushin, G. Hydrothermal carbonization of abundant renewable natural organic chemicals for high-performance supercapacitor electrodes. Adv. Energy Mater. 2011, 1, 356−361.

(24) Hou, J.; Cao, C.; Idrees, F.; Ma, X. Hierarchical porous nitrogendoped carbon nanosheets derived from silk for ultrahigh capacity battery anodes and supercapacitors. ACS Nano 2015, 9, 2556−2564. (25) Tian, W.; Gao, Q.; Zhang, L.; Yang, C.; Li, Z.; Tan, Y.; Qian, W.; Zhang, H. Renewable graphene-like nitrogen-doped carbon nanosheets as supercapacitor electrodes with integrated high energy−power property. J. Mater. Chem. A 2016, 4, 8690−8699. (26) Genovese, M.; Jiang, J.; Lian, K.; Holm, N. High capacitive performance of exfoliated biochar nanosheets from biomass waste corn cob. J. Mater. Chem. A 2015, 3, 2903−2913. (27) Barin, G. B.; de Fátima Gimenez, I.; da Costa, L. P.; Filho, A. G. S.; Barreto, L. S. Influence of hydrothermal carbonization on formation of curved graphite structures obtained from a lignocellulosic precursor. Carbon 2014, 78, 609−612. (28) Yun, Y. S.; Park, M. H.; Hong, S. J.; Lee, M. E.; Park, Y. W.; Jin, H.J. Hierarchically porous carbon nanosheets from waste coffee grounds for supercapacitors. ACS Appl. Mater. Interfaces 2015, 7, 3684−3690. (29) Long, C.; Chen, X.; Jiang, L.; Zhi, L.; Fan, Z. Porous layer-stacking carbon derived from in-built template in biomass for high volumetric performance supercapacitors. Nano Energy 2015, 12, 141−151. (30) Wang, L.; Mu, G.; Tian, C.; Sun, L.; Zhou, W.; Yu, P.; Yin, J.; Fu, H. Porous graphitic carbon nanosheets derived from cornstalk biomass for advanced supercapacitors. ChemSusChem 2013, 6, 880−889. (31) Kim, T.; Jung, G.; Yoo, S.; Suh, K. S.; Ruoff, R. S. Activated graphene-based carbons as supercapacitor electrodes with macro-and mesopores. ACS Nano 2013, 7, 6899−6905. (32) Ferrero, G.; Fuertes, A.; Sevilla, M. From soybean residue to advanced supercapacitors. Sci. Rep. 2015, 5, 16618. (33) Tian, W.; Gao, Q.; Tan, Y.; Yang, K.; Zhu, L.; Yang, C.; Zhang, H. Bio-inspired beehive-like hierarchical nanoporous carbon derived from bamboo-based industrial by-product as a high performance supercapacitor electrode material. J. Mater. Chem. A 2015, 3, 5656−5664. (34) Li, Z.; Zhang, L.; Amirkhiz, B. S.; Tan, X.; Xu, Z.; Wang, H.; Olsen, B. C.; Holt, C. M. B.; Mitlin, D. Carbonized chicken eggshell membranes with 3D architectures as high-performance electrode materials for supercapacitors. Adv. Energy Mater. 2012, 2, 431−437. (35) Zhang, Z.; Mu, S.; Zhang, B.; Tao, L.; Huang, S.; Huang, Y.; Gao, F.; Zhao, Y. A novel synthesis of carbon nanotubes directly from an indecomposable solid carbon source for electrochemical applications. J. Mater. Chem. A 2016, 4, 2137−2146. (36) Zhao, Y.; Zhang, Z.; Ren, Y.; Ran, W.; Chen, X.; Wu, J.; Gao, F. Vapor deposition polymerization of aniline on 3D hierarchical porous carbon with enhanced cycling stability as supercapacitor electrode. J. Power Sources 2015, 286, 1−9. (37) Wang, L.; Tian, C.; Wang, B.; Wang, R.; Zhou, W.; Fu, H. Controllable synthesis of graphitic carbon nanostructures from ionexchange resin-iron complex via solid-state pyrolysis process. Chem. Commun. 2008, 5411−5413. (38) Peng, H.; Ma, G.; Sun, K.; Mu, J.; Zhang, Z.; Lei, Z. Formation of carbon nanosheets via simultaneous activation and catalytic carbonization of macroporous anion-exchange resin for supercapacitors application. ACS Appl. Mater. Interfaces 2014, 6, 20795−20803. (39) Wang, L.; Tian, C.; Wang, H.; Ma, Y.; Wang, B.; Fu, H. Mass production of graphene via an in situ self-generating template route and its promoted activity as electrocatalytic support for methanol electroxidization. J. Phys. Chem. C 2010, 114, 8727−8733. (40) Lei, H.; Yan, T.; Wang, H.; Shi, L.; Zhang, J.; Zhang, D. Graphenelike carbon nanosheets prepared by a Fe-catalyzed glucose-blowing method for capacitive deionization. J. Mater. Chem. A 2015, 3, 5934− 5941. (41) Zhao, Z.; Hao, S.; Hao, P.; Sang, Y.; Manivannan, A.; Wu, N.; Liu, H. Lignosulphonate-cellulose derived porous activated carbon for supercapacitor electrode. J. Mater. Chem. A 2015, 3, 15049−15056. (42) Jain, A.; Balasubramanian, R.; Srinivasan, M. P. Tuning hydrochar properties for enhanced mesopore development in activated carbon by hydrothermal carbonization. Microporous Mesoporous Mater. 2015, 203, 178−185. (43) Ding, J.; Wang, H.; Li, Z.; Cui, K.; Karpuzov, D.; Tan, X.; Kohandehghan, A.; Mitlin, D. Peanut shell hybrid sodium ion capacitor 1305

DOI: 10.1021/acssuschemeng.6b01390 ACS Sustainable Chem. Eng. 2017, 5, 1297−1305