Facile Synthesis of Highly Porous Carbon from Rice Husk - ACS

Jun 24, 2017 - Highly porous carbon materials have attracted great interest for a wide range of important applications. Many examples for their synthe...
1 downloads 32 Views 3MB Size
Research Article pubs.acs.org/journal/ascecg

Facile Synthesis of Highly Porous Carbon from Rice Husk Yeru Liang,* Chen Yang, Hanwu Dong, Wenqi Li, Hang Hu, Yong Xiao, Mingtao Zheng, and Yingliang Liu* College of Materials and Energy, South China Agricultural University, Guangzhou 510642, P. R. China S Supporting Information *

ABSTRACT: Highly porous carbon materials have attracted great interest for a wide range of important applications. Many examples for their synthesis exist, but these synthetic processes can be quite complex and also very time-consuming. There is still a major challenge to develop a facile yet versatile conceptual approach to produce them. Here, we present an efficient, activation-free, post-treatment-free strategy for the synthesis of highly porous carbon by a simple carbonization of a mixture of rice husk and polytetrafluoroethylene (PTFE) powder. PTFE employed here can in situ generate HF to etch out natural silica during the carbonization treatment of rice husk. This strategy not only reduces the synthesis procedure by combining carbonization and post-removal of silica into a single step but also eliminates completely the usage of hazardous HF or corrosive NaOH or KOH. The as-synthesized carbon materials exhibit a BET surface area as high as 2051 m2/g without any activation treatment, which is about 20 times enhanced in porosity compared to that of the traditional carbon material from rice husk. With the combination of the high porosity and the valuable hierarchical porous structure, the as-prepared porous carbon materials serve well as electrodes for supercapacitive energy storage, including a large capacitance of 317 F/g, good rate performance, and high capacitances per surface area. These findings could provide a new avenue for the facile production of high-performance porous carbon materials with promising applications in various areas. KEYWORDS: Highly porous carbon, Rice husk, Activation-free, No more post-treatment, Energy storage



INTRODUCTION Porous carbons have attracted significant scientific and technological interest because of their unique features including large specific surface area, extraordinary electrical and thermal properties.1−5 These intriguing characteristics lead to a range of important applications such as energy storage,6−9 adsorption,10,11 separation,12,13 drug delivery,14,15 and catalysis.16−19 Because of the decisive role of the surface area in determining porous materials’ performance in many applications, development of highly porous carbons has been one of the most active areas in the carbonaceous material science.20,21 Many methods for the synthesis of high-surface-area carbon exist, including activation,22−26 direct carbonization of porous organic materials,21,27−32 halogenation of porous carbides,33,34 and templating strategies (Figure S1).35−39 Among the various methods, chemical activation is a powerful and well-established technique to construct a highly porous structure with large surface area (e.g., above 2000 m2/ g).40−43 However, it usually involves the use of a high-amount corrosive activator such as KOH, which is inconsistent with the urgent need of energy conservation and environmental protection. Furthermore, in order to remove impurities derived from activators, repeated post-purification treatments are often required after activation. For example, plenty of water is often © 2017 American Chemical Society

required to wash off the impurities from the product, inevitably leading to the generation of tremendous waste liquid pollution. Despite remarkable advances of templating methodologies, the most common hard templates, i.e., silica-based materials, are facing a significant bottleneck.44,45 The removal of the silica template after carbonization always requires harsh repeated chemical post-treatments, such as hydrofluoric acid (HF) or NaOH etching. HF is hazardous, and NaOH post-treatment is usually time-consuming and not-cost efficient because of it being handled at elevated temperatures.44,46−48 Though direct carbonization and high-temperature chlorination to prepare high-surface-area carbon have been reported as alternate methods, they involve the use of costly carbon sources or expensive catalysts. Besides, some of these starting porous materials are prepared with complicated multiple synthetic steps or under rigorous conditions (e.g., anhydrous and anaerobic), leading to scale-up challenges.49−51 As a cheap and environmentally friendly sustainable resource, many biomasses have already been explored as attractive raw materials for fabricating carbon materials.43,52−56 Particularly, Received: April 27, 2017 Revised: June 20, 2017 Published: June 24, 2017 7111

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117

Research Article

ACS Sustainable Chemistry & Engineering

Figure 1. Schematic illustration of the comparison of the conventional and our synthetic route developed here to HPCs. The gray arrow shows the fabrication of conventional carbon materials by carbonization of RHs. This method is complicated and usually leads to carbon products with low surface areas. Thus, chemical activation treatments are required to increase their surface areas. The red arrow shows fabrication of HPCs via in situ removal of natural silica during carbonization treatment of RHs. This facile route can combine the carbonization and post-removal of silica in a single step and lead to the generation of a well-developed hierarchical porous structure upon carbonization.

supply in the world, and thus, this is energy-efficient and easy to scale-up. With the combination of the high porosity and the valuable hierarchical porous structure, the as-prepared HPCs serve well when used as electrode materials for supercapacitors. These findings will provide a new avenue for the facile preparation of high-surface-area carbon materials from biomass for challenging environmental and energy issues.

porous carbons fabricated from waste rice husks (RHs) have shown promising applications as electrodes, adsorbents, catalyst supports, and others. Nevertheless, the problem is that the activation method is still a general even essential way to obtain those highly porous carbon materials (HPCs) from biomass including RHs. There is very limited simple and satisfactory processes for the production of biomass-based HPCs to date. Hence, establishing a facile and efficient synthesis protocol using low-cost precursors for the preparation of high-surfacearea porous carbons is still a great challenge but very desirable for their facile preparation and broad applications. Herein, we report the development of a cost-effective, activation-free, and post-treatment-free way to prepare a class of HPCs by in situ etching natural silica during the carbonization treatment of RHs (Figure 1). The key to this preparation strategy is utilization of polytetrafluoroethylene (PTFE), which can in situ generate HF to etch out the natural nanosilica during carbonization treatment of RHs. As illustrated in Figure 1, the overall strategy is very straightforward. The HPCs can be obtained by simply mixing the PTFE powder and the RHs, followed by direct carbonization. Compared with the reported methods of producing porous carbon, this strategy demonstrates several dominant characteristics. First of all, the carbonization and silica removal are strategically accomplished simultaneously, reducing the synthetic steps when compared to normal templating procedures to porous carbon. Moreover, the post-treatment for the removal of silica, i.e., the use of hazardous HF or repeated NaOH washes, is eliminated. In addition, the surface area of the as-obtained carbon materials can be up to 2051 m2/g without any activation treatment, which is 20 times enhanced in porosity compared to that of the traditional RH-based carbon materials. Finally, RHs employed here are abundant and renewable material sources with a huge



RESULTS AND DISCUSSION As a class of naturally occurring biomass, RHs are composed of nanosilica, cellulose, lignin, and very little other materials including fats and protein (Figure 2a).57,58 As the major constituent, small and large silica nanoparticles are aggregated and formed on the intercellular layers and surface of cell walls.59 Thermogravimetric analysis (TGA) of RHs shows that the content of the main inorganic material, i.e., nanosilica, is about 14 wt % (Figure S2). In order to remove these silica compounds, post-etching with corrosive HF or chemical activation is normally needed after the carbonization of RH (Figure 1). In contrast to the conventional synthesis methods with a two-step process, our route combines the carbonization and post-removal of silica in a single step with complete avoidance of post-treatment. The RHs are mixed with PTFE powder and then directly carbonized through heating under N2 atmosphere, readily leading to the formation of target sample HPC-RH-1. Scanning electron microscope (SEM) images at different magnifications are displayed in Figure 2b and c. It is found that after a simple carbonization treatment at 900 °C, the obtained HPC-RH-1 reveals a typical hierarchical porous morphology. The magnified SEM image in Figure 2c shows that after the high temperature carbonization treatment, the resulting carbon frameworks are fractured on the nanometer scale to form 7112

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117

Research Article

ACS Sustainable Chemistry & Engineering

the HPC-RH-1 displays a high adsorption uptake at low relative pressure (P/P0), indicating the presence of plentiful micropores. After that, the adsorption amount continuously increases but still does not reach a plateau near the P/P0 of 1.0, implying the existence of mesopores and macropores. The density functional theory (DFT) pore size distribution in Figure 3b shows the micropores center at 0.8 and 1.3 nm, and the meso-/ macropores range from 2 to 120 nm with a maximum at 17 nm. Brunauer−Emmett−Teller (BET) specific surface area (SBET) and pore volume of HPC-RH-1 are calculated to be as high as 2051 m2/g and 1.36 cm3/g, respectively, and their micropore surface area and meso-/macropore surface area are measured to be 1274 and 777 m2/g, respectively. It is noteworthy that this high-surface-area HPC-RH-1 with good reproducibility outperforms the majority of porous carbon without any extra activation techniques and is even comparable to a lot of activated carbons which are produced through complicated chemical activation procedures (Figure S9 and Table S1). In sharp contrast, the control carbon sample has much smaller SBET of 101 m2/g and pore volume of 0.16 cm3/g (Figure 3a), which are in good agreement with the observation from SEM and TEM images. The maximum pore size distribution peaks for the control carbon sample are fixed at 4 and 37 nm (Figure S10). Such a difference in pore structure mainly resulted from the formation of additional pores by the removal of volatile oxygen-containing substances upon further pyrolysis. It is demonstrated that when a silica-containing hybrid is mixed with PTFE and heated in a N2 atmosphere with a slow rate (e.g., 5 °C/min), simultaneous carbonization and silica removal occur within 650 °C.48 That is to say, the silica component of RH can be completely removed below 650 °C by releasing SiF4, CO2, and other oxygenate gases (Figure S11), leading to the creation of porous carbonaceous materials. When further increasing the temperature until 900 °C and maintaining 900 °C for several hours, more mass loss occurs because the porous structure provides much more room for large quantities of volatile noncarbon elements and carbon-containing compounds to effectively release. Therefore, more nanopores are formed, leading to a high-surface-area carbonaceous structure in a single step. On the contrary, for the control carbon sample, the presence of a tight silica−carbon interface during the whole carbonization process inhibits the formation of such additional room for the release of volatile products (Figures S7 and S8), resulting in a relatively low surface area. The carbon framework structure of HPC-RH-1 is studied by wide-angle X-ray diffraction (XRD) patterns, Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS). The XRD pattern exhibits two broad peaks centered at 2θ ≈ 22° and 44°, demonstrating the formation of the low crystallinity degree of the carbon framework (Figure S12). The Raman spectrum shows two characteristic peaks centered at about 1350 and 1600 cm−1, which are attributed to the D and G bands, respectively (Figure S13). A broader D-band further indicates that HPC-RH-1 possesses low graphite crystallinity of the carbon framework, whose carbon nanosheets and clusters possess a microcrystalline plane crystal size (i.e., La) of 2.4 nm and then form abundant micropores (the inset of Figure 2d). An integrated illustration of the chemical composition of HPCRH-1 from XPS is shown in Figure S14. It is shown that the HPC-RH-1 is mainly composed of carbon, oxygen, and hydrogen elements. No other elements like silicon are detected, further indicative of the complete removal of the silica component for the final HPC-RH-1 sample.

Figure 2. SEM images of (a) RH, (b,c) HPC-RH-1, and (e) control carbon sample; the inset digital photo in panel a shows the macroscopical form of RH. TEM images of (d) HPC-RH-1 and (f) control carbon sample; the inset in panel d shows the graphite-like microcrystalline structure in HPC-RH-1.

abundant carbon nanoparticles with diameters in the range of 10−40 nm. The close and loose aggregation of these carbon nanoparticles lead to the formation of mesopores and macropores, respectively. Transmission electron microscope (TEM) observation confirms the highly porous structure in HPC-RH-1 (Figure 2d). The absence of a Si peak in energydispersive X-ray spectroscopy (EDS) of HPC-RH-1 carbon demonstrates the complete removal of the nanosilica component (Figure S3). For comparison, a control carbon sample is prepared through a traditional two-step strategy. EDS analyses demonstrate the chemical compositions of HPC-RH-1, and the control carbon sample reveals few differences. The silica of the control sample is also removed by HF post-etching (Figures S3−S8). However, as shown in Figure 2e and f, the control carbon sample displays a significantly poorer porous structure when compared with that of HPC-RH-1. N2 adsorption measurement is employed to quantitatively illuminate the pore characteristics of the as-prepared samples. As shown in N2 adsorption−desorption isotherms of Figure 3a,

Figure 3. Pore structures of typical samples. (a) N2 adsorption− desorption isotherms of HPC-RH-1 (red) and control carbon sample (gray). (b) DFT pore size distribution curve of HPC-RH-1. (c) SBET and (d) pore volume of HPCs obtained at various carbonization times. 7113

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117

Research Article

ACS Sustainable Chemistry & Engineering The nanostructure of HPCs can be readily adjusted through tuning the fabrication conditions (Table S2 and Figures S15 and S16). For example, it is found that carbonization time plays an important role in altering BET surface area. When extending the carbonization time from 4 to 20 h, both BET surface area and pore volume decrease (Figure 3c and d). In addition, the BET surface area increases from 1095 to 2051 m2/g with increasing the silica/PTFE weight ratio from 0.03 to 0.055. While further increasing the weight ratio to 0.09, the BET surface area decreases to 1383 m2/g (Figures S17 and S18). Furthermore, calculated from the Raman spectra of the samples obtained at various carbonization times, the values of La of HPCs gradually increase when increasing carbonization time (Table S3 and Figures S13 and S19). Owing to their well-developed porosity and robust hierarchical porous morphology, HPCs developed in this study hold great promise for their application in many fields. Here, we focus on the electrochemical properties of HPCs used as the electrodes of supercapacitors. Supercapacitor, an emerging class of energy storage device with high power density, long cycle life, and short charging time, is highly attractive as a power source for portable electronic devices.60−63 Currently, the core technology of high-performance supercapacitors is the development of carbon-based electrodes with both high capacitance and fast ion/electron transport properties. To analyze the usability of HPCs for promising supercapacitor electrodes, the HPC-RH-1 is evaluated as a proof-of-concept demonstration in a two-electrode symmetrical cell by using 6 mol/L KOH aqueous solution as an electrolyte. Typical quasi-rectangular shapes in cyclic voltammetry (CV) curves and triangular profiles in galvanostatic charge−discharge curves both demonstrate the ideal electrochemical capacitive characteristics (Figures S20 and S21). The capacitances calculated from galvanostatic charge−discharge measurements at various densities are summarized in Figure 4a. It is clearly seen that HPC-RH-1 exhibits excellent charge storage capacity. For example, HPC-RH-1 possesses specific capacitances as high as 317 F/g at a current density of 50 mA/g, significantly higher than 202 F/g for the commercially benchmark activated carbon YP-50 with a BET surface area of 1713 m2/g. Even under the high current density of 10000 mA/g, high capacitance of 245 F/g could still be retained for HPC-RH-1, giving good rate capability (Figure 4b). These values are also clearly higher than those of many other reported high-performance carbon-based electrodes (Table S4), demonstrating that the HPC-RH-1 developed here is very attractive in electrochemical energy storage. The significant superiority of HPC-RH-1 in terms of the electrochemical performances is repeatable (Figure S22), which is mainly attributed to its more advanced porous structure. For activated carbon YP-50, most of its pores are micropores and/or small-sized mesopores (Figure S23), a large proportion of which are located in the micron/millimeterscaled carbon particles. Because of the absence of efficient ion diffusion pathways, many of its inner micropores are hard to immerse in electrolytes, leading to poor utilization of the surface area and a decrease in capacitance. On the contrary, HPC-RH-1 simultaneously possesses a high surface area and a valuable hierarchical porous morphology. The more robust porosity in an HPC-RH-1-based electrode, especially its microporosity, enhances the formation of an electric double layer for higher charge storage when compared with that of the YP-50-based electrode (Figure S24). Meanwhile, the hierarchally meso- and macroporous structure offers minimized

Figure 4. Comparison of supercapacitive performance for typical samples. (a) Specific capacitances of HPC-RH-1 and commercially benchmark activated carbon YP-50. (b) Capacitance retention ratios (solid) and specific capacitances per surface area (hollow) of HPCRH-1.

diffusive resistance to electrolyte ion transport to access the micropores, thus having lower impedance (Figure S25) and better specific capacitance per surface area when compared with those of YP-50 (Figures 4b and S26).



CONCLUSIONS In conclusion, we have developed a versatile and effective approach for the synthesis of highly porous carbons by simple carbonization of a mixture of rice husk and PTFE powder. The PTFE is capable of in situ etching natural silica components of rice husk during the carbonization process, thus eliminating the tedious post-treatment and then decreasing the synthetic steps. The as-obtained carbon material possesses a high surface area of up to 2051 m2/g, which exhibits excellent supercapacitive energy storage performance. It is envisioned that further tuning of pore structures could be achieved by variation of precursor type and size, carbonization temperature, and heating rate. We hope that this work will provide new opportunities for the efficient production of high-performance porous materials with promising applications in various fields such as supercapacitors, lithium−sulfur batteries, adsorption, and catalysis.



METHODS

Materials. RHs were obtained as a byproduct of rice harvested in the suburbs of Xiantao City in P. R. China and directly used as the starting material without any pretreatment. Polytetrafluoroethylene (DuPont Company, 7A), hydrofluoric acid (HF, Guangzhou Chemical Reagent Factory), and other chemicals were used as received. Synthesis of HPCs-RH. In a typical process, RH powder was obtained by smashing about 50 g of RH in an electric grinder (Laifu, LFP-800T) for 20 min with a revolving speed of 29000 r/min. For the preparation of HPC-RH-1, 0.5 g of RH powder was closely mixed with 1.28 g of PTFE powder by grinding in a mortar. The mixture was subsequently transferred to an alumina boat. The boat was then placed inside a tubular furnace and then heated to 900 °C at 5 °C/min for 4 h 7114

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117

Research Article

ACS Sustainable Chemistry & Engineering under N2 flow, leading to the formation of the HPC-RH-1 product. For fabrication of other HPC-RHs, the overall procedures were exactly the same as those of HPC-RH-1 except for the carbonization time and RH/PTFE weight ratio. Detailed fabrication parameters are listed in Table S2. Synthesis of Control Carbon Sample. RH powder was placed inside a tubular furnace and then heated to 900 °C at 5 °C/min for 4 h under N2 flow. After carbonization, the as-obtained control carbon/ silica composites were washed using 40 wt % HF solution to remove the silica, followed by filtration and repeated flushing with distilled water until the filtrate was neutral. After drying at 100 °C, the control carbon sample was obtained. Structural Characterization. The microstructure of the samples was investigated by a Quanta 400F scanning electron microscope (SEM) and a FEI Tecnai G2 Spirit transmission electron microscope (TEM). Raman measurements were carried out with an inVia-Reflex Renishaw Raman system. Thermogravimetric (TG) analysis was performed under flowing air conditions at a heating rate of 20 °C/min. TG coupled with Fourier transform infrared spectrometry (TG-FTIR) measurements were performed under N2 flow at a heating rate of 10 °C/min. The TG and FTIR spectra were recorded on a Netzsch TG209 and a Bruker Vector-22 infrared spectroscope, respectively. N2 adsorption measurements were carried out using a Micromeritics ASAP 2020 analyzer at 77K. The BET surface area (SBET) was estimated by the BET theory. The micropore surface area (Smic) was analyzed by the t-plot method, and then the external surface area (Sext) was obtained by subtracting the Smic from the SBET. The pore size distribution was determined by original DFT combined with nonnegative regularization and medium smoothing. The total pore volume (Vt) was calculated according to the amount adsorbed at a relative pressure P/P0 of about 0.99. XRD patterns were conducted on a DMAX 2200 VPC diffractometer using Cu K radiation (40 kV, 30 mA). Raman measurements were carried out with inVia-Reflex Renishaw Raman system. An ESCALAB250 instrument was used to record the XPS data. Fabrication and Measurements of Supercapacitors. The electrodes in the form of round sheets with a diameter of 1.2 cm were obtained by pressing a mixture film of 87 wt % carbon sample, 5 wt % carbon black with a BET surface area of 62 m2/g, and 8 wt % PTFE. The average mass of the carbon sample for an electrode was controlled at about 3 mg. The specific surface area of the electrode (Selectrode) was calculated according to the equation Selectrode = Scarbon−sample × 87% + Scarbon−black × 5% + SPTFE × 8%, where Scarbon‑sample, Scarbon‑black, and SPTFE represent the BET surface area of the carbon sample, carbon black, and PTFE, respectively. Nickel foam was used as the current collector. 6 mol/L KOH aqueous solution was used as the electrolyte. The electrochemical measurements were characterized with the assembled coin-type supercapacitor. CV, galvanostatic charge− discharge, and electrochemical impedance spectroscopy (EIS) measurements were carried out using a Chenhua electrochemical workstation. The signal amplitude in the EIS measurement was 5 mV.



Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully acknowledge financial support from the projects of the National Natural Science Foundation of China (51602107 and U1501242), Program for Pearl River New Star of Science and Technology in Guangzhou (201710010104), and the Cultivation Fund of Scientific and Technical Youth Talents of South China Agricultural University.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.7b01315. Additional information about material characterization (PDF)



REFERENCES

(1) Liu, J.; Wickramaratne, N. P.; Qiao, S. Z.; Jaroniec, M. Molecularbased design and emerging applications of nanoporous carbon spheres. Nat. Mater. 2015, 14, 763−774. (2) Zhai, Y.; Dou, Y.; Zhao, D.; Fulvio, P. F.; Mayes, R. T.; Dai, S. Carbon Materials for Chemical Capacitive Energy Storage. Adv. Mater. 2011, 23, 4828−4850. (3) Wan, Y.; Shi, Y.; Zhao, D. Supramolecular aggregates as templates: Ordered mesoporous polymers and carbons. Chem. Mater. 2008, 20, 932−945. (4) Su, D. S.; Schlogl, R. Nanostructured Carbon and Carbon Nanocomposites for Electrochemical Energy Storage Applications. ChemSusChem 2010, 3, 136−168. (5) Su, F.; Poh, C. K.; Chen, J. S.; Xu, G.; Wang, D.; Li, Q.; Lin, J.; Lou, X. W. Nitrogen-containing microporous carbon nanospheres with improved capacitive properties. Energy Environ. Sci. 2011, 4, 717−724. (6) Zheng, G.; Lee, S. W.; Liang, Z.; Lee, H.-W.; Yan, K.; Yao, H.; Wang, H.; Li, W.; Chu, S.; Cui, Y. Interconnected hollow carbon nanospheres for stable lithium metal anodes. Nat. Nanotechnol. 2014, 9, 618−623. (7) Wang, D. W.; Li, F.; Liu, M.; Lu, G. Q.; Cheng, H. M. 3D Aperiodic Hierarchical Porous Graphitic Carbon Material for HighRate Electrochemical Capacitive Energy Storage. Angew. Chem., Int. Ed. 2008, 47, 373−376. (8) Zhang, C.; Lv, W.; Tao, Y.; Yang, Q.-H. Towards superior volumetric performance: design and preparation of novel carbon materials for energy storage. Energy Environ. Sci. 2015, 8, 1390−1403. (9) Zhang, C.; Wu, H. B.; Yuan, C.; Guo, Z.; Lou, X. W. Confining Sulfur in Double-Shelled Hollow Carbon Spheres for Lithium−Sulfur Batteries. Angew. Chem., Int. Ed. 2012, 51, 9592−9595. (10) Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.; et al. Marinas, B. J.; Mayes, A. M., Science and technology for water purification in the coming decades. Nature 2008, 452, 301−310. (11) Sevilla, M.; Fuertes, A. B. Sustainable porous carbons with a superior performance for CO2 capture. Energy Environ. Sci. 2011, 4, 1765−1771. (12) Kim, H. W.; Yoon, H. W.; Yoon, S.-M.; Yoo, B. M.; Ahn, B. K.; Cho, Y. H.; Shin, H. J.; Yang, H.; Paik, U.; Kwon, S.; Choi, J.-Y.; Park, H. B. Selective Gas Transport Through Few-Layered Graphene and Graphene Oxide Membranes. Science 2013, 342, 91−95. (13) Yu, Z.-L.; Li, G.-C.; Fechler, N.; Yang, N.; Ma, Z.-Y.; Wang, X.; Antonietti, M.; Yu, S.-H. Polymerization under Hypersaline Conditions: A Robust Route to Phenolic Polymer-Derived Carbon Aerogels. Angew. Chem., Int. Ed. 2016, 55, 14623−14627. (14) Fang, Y.; Zheng, G.; Yang, J.; Tang, H.; Zhang, Y.; Kong, B.; Lv, Y.; Xu, C.; Asiri, A. M.; Zi, J.; Zhang, F.; Zhao, D. Dual-Pore Mesoporous Carbon@Silica Composite Core−Shell Nanospheres for Multidrug Delivery. Angew. Chem., Int. Ed. 2014, 53, 5366−5370. (15) Zhang, L.; Li, Y. C.; Jin, Z. X.; Chan, K. M.; Yu, J. C. Mesoporous carbon/CuS nanocomposites for pH-dependent drug delivery and near-infrared chemo-photothermal therapy. RSC Adv. 2015, 5, 93226−93233. (16) Bhosale, M. E.; Illathvalappil, R.; Kurungot, S.; Krishnamoorthy, K. Conjugated porous polymers as precursors for electrocatalysts and storage electrode materials. Chem. Commun. 2016, 52, 316−318.

AUTHOR INFORMATION

Corresponding Authors

*(Y.R. Liang) E-mail: [email protected]. *(Y.L. Liu) E-mail: [email protected]. ORCID

Yeru Liang: 0000-0002-6169-9981 7115

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117

Research Article

ACS Sustainable Chemistry & Engineering (17) Fang, B.; Kim, J. H.; Kim, M.; Yu, J.-S. Ordered Hierarchical Nanostructured Carbon as a Highly Efficient Cathode Catalyst Support in Proton Exchange Membrane Fuel Cell. Chem. Mater. 2009, 21, 789−796. (18) Yu, H. J.; Shang, L.; Bian, T.; Shi, R.; Waterhouse, G. I. N.; Zhao, Y. F.; Zhou, C.; Wu, L. Z.; Tung, C. H.; Zhang, T. R. NitrogenDoped Porous Carbon Nanosheets Templated from g-C3N4 as MetalFree Electrocatalysts for Efficient Oxygen Reduction Reaction. Adv. Mater. 2016, 28, 5080−5086. (19) Chen, P.; Wang, L.-K.; Wang, G.; Gao, M.-R.; Ge, J.; Yuan, W.J.; Shen, Y.-H.; Xie, A.-J.; Yu, S.-H. Nitrogen-doped nanoporous carbon nanosheets derived from plant biomass: an efficient catalyst for oxygen reduction reaction. Energy Environ. Sci. 2014, 7, 4095−4103. (20) Chae, H. K.; Siberio-Perez, D. Y.; Kim, J.; Go, Y.; Eddaoudi, M.; Matzger, A. J.; O’Keeffe, M.; Yaghi, O. M. A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004, 427, 523−527. (21) Xu, F.; Tang, Z. W.; Huang, S. Q.; Chen, L. Y.; Liang, Y. R.; Mai, W. C.; Zhong, H.; Fu, R. W.; Wu, D. C. Facile synthesis of ultrahigh-surface-area hollow carbon nanospheres for enhanced adsorption and energy storage. Nat. Commun. 2015, 6, 7221. (22) 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. (23) Niu, S.; Zhou, G.; Lv, W.; Shi, H.; Luo, C.; He, Y.; Li, B.; Yang, Q.-H.; Kang, F. Sulfur confined in nitrogen-doped microporous carbon used in a carbonate-based electrolyte for long-life, safe lithiumsulfur batteries. Carbon 2016, 109, 1−6. (24) Wang, X.; Kong, D.; Zhang, Y.; Wang, B.; Li, X.; Qiu, T.; Song, Q.; Ning, J.; Song, Y.; Zhi, L. All-biomaterial supercapacitor derived from bacterial cellulose. Nanoscale 2016, 8, 9146−9150. (25) Long, C.; Chen, X.; Jiang, L.; Zhi, L.; Fan, Z. Porous layerstacking carbon derived from in-built template in biomass for high volumetric performance supercapacitors. Nano Energy 2015, 12, 141− 151. (26) Wu, D. C.; Liang, Y. R.; Yang, X. Q.; Zou, C.; Li, Z. H.; Lv, G. F.; Zeng, X. H.; Fu, R. W. Preparation of activated ordered mesoporous carbons with a channel structure. Langmuir 2008, 24, 2967−2969. (27) Yang, S. J.; Kim, T.; Im, J. H.; Kim, Y. S.; Lee, K.; Jung, H.; Park, C. R. MOF-Derived Hierarchically Porous Carbon with Exceptional Porosity and Hydrogen Storage Capacity. Chem. Mater. 2012, 24, 464−470. (28) Xu, Z. X.; Zhuang, X. D.; Yang, C. Q.; Cao, J.; Yao, Z. Q.; Tang, Y. P.; Jiang, J. Z.; Wu, D. Q.; Feng, X. L. Nitrogen-Doped Porous Carbon Superstructures Derived from Hierarchical Assembly of Polyimide Nanosheets. Adv. Mater. 2016, 28, 1981−1987. (29) Zheng, F. C.; Yang, Y.; Chen, Q. W. High lithium anodic performance of highly nitrogen-doped porous carbon prepared from a metal-organic framework. Nat. Commun. 2014, 5, 5261. (30) Zhang, W.; Wu, Z.-Y.; Jiang, H.-L.; Yu, S.-H. Nanowire-Directed Templating Synthesis of Metal−Organic Framework Nanofibers and Their Derived Porous Doped Carbon Nanofibers for Enhanced Electrocatalysis. J. Am. Chem. Soc. 2014, 136, 14385−14388. (31) Zou, C.; Wu, D.; Li, M.; Zeng, Q.; Xu, F.; Huang, Z.; Fu, R. Template-free fabrication of hierarchical porous carbon by constructing carbonyl crosslinking bridges between polystyrene chains. J. Mater. Chem. 2010, 20, 731−735. (32) Li, Z.; Li, L.; Li, Z.; Liao, H.; Zhang, H. Ultrathin carbon gauze for high-rate supercapacitor. Electrochim. Acta 2016, 222, 990−998. (33) Korenblit, Y.; Rose, M.; Kockrick, E.; Borchardt, L.; Kvit, A.; Kaskel, S.; Yushin, G. High-Rate Electrochemical Capacitors Based on Ordered Mesoporous Silicon Carbide-Derived Carbon. ACS Nano 2010, 4, 1337−1344. (34) Gogotsi, Y.; Nikitin, A.; Ye, H.; Zhou, W.; Fischer, J. E.; Yi, B.; Foley, H. C.; Barsoum, M. W. Nanoporous carbide-derived carbon with tunable pore size. Nat. Mater. 2003, 2, 591−594. (35) Lee, J.; Kim, J.; Hyeon, T. Recent Progress in the Synthesis of Porous Carbon Materials. Adv. Mater. 2006, 18, 2073−2094.

(36) Jun, S.; Joo, S. H.; Ryoo, R.; Kruk, M.; Jaroniec, M.; Liu, Z.; Ohsuna, T.; Terasaki, O. Synthesis of New, Nanoporous Carbon with Hexagonally Ordered Mesostructure. J. Am. Chem. Soc. 2000, 122, 10712−10713. (37) Fan, Z.; Liu, Y.; Yan, J.; Ning, G.; Wang, Q.; Wei, T.; Zhi, L.; Wei, F. Template-Directed Synthesis of Pillared-Porous Carbon Nanosheet Architectures: High-Performance Electrode Materials for Supercapacitors. Adv. Energy Mater. 2012, 2, 419−424. (38) Ding, S. J.; Wang, Z. Y.; Madhavi, S.; Lou, X. W. SBA-15 derived carbon-supported SnO2 nanowire arrays with improved lithium storage capabilities. J. Mater. Chem. 2011, 21, 13860−13864. (39) Liang, Y.; Liu, H.; Li, Z.; Fu, R.; Wu, D. In situ polydopamine coating-directed synthesis of nitrogen-doped ordered nanoporous carbons with superior performance in supercapacitors. J. Mater. Chem. A 2013, 1, 15207−15211. (40) Zhu, Y.; Murali, S.; Stoller, M. D.; Ganesh, K. J.; 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. (41) Ding, J.; Wang, H. L.; Li, Z.; Cui, K.; Karpuzov, D.; Tan, X. H.; Kohandehghan, A.; Mitlin, D. Peanut shell hybrid sodium ion capacitor with extreme energy-power rivals lithium ion capacitors. Energy Environ. Sci. 2015, 8, 941−955. (42) 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. (43) 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. (44) Liu, R. L.; Shi, Y. F.; Wan, Y.; Meng, Y.; Zhang, F. Q.; Gu, D.; Chen, Z. X.; Tu, B.; Zhao, D. Y. Triconstituent Co-assembly to ordered mesostructured polymer-silica and carbon-silica nanocomposites and large-pore mesoporous carbons with high surface areas. J. Am. Chem. Soc. 2006, 128, 11652−11662. (45) Li, S.; Liang, Y.; Wu, D.; Fu, R. Fabrication of bimodal mesoporous carbons from petroleum pitch by a one-step nanocasting method. Carbon 2010, 48, 839−843. (46) Wu, D.; Li, Z.; Zhong, M.; Kowalewski, T.; Matyjaszewski, K. Templated Synthesis of Nitrogen-Enriched Nanoporous Carbon Materials from Porogenic Organic Precursors Prepared by ATRP. Angew. Chem., Int. Ed. 2014, 53, 3957−3960. (47) Li, J.; Liang, Y.; Dou, B.; Ma, C.; Lu, R.; Hao, Z.; Xie, Q.; Luan, Z.; Li, K. Nanocasting synthesis of graphitized ordered mesoporous carbon using Fe-coated SBA-15 template. Mater. Chem. Phys. 2013, 138, 484−489. (48) Singh, D. K.; Krishna, K. S.; Harish, S.; Sampath, S.; Eswaramoorthy, M.; No More, H. F. Teflon-Assisted Ultrafast Removal of Silica to Generate High-Surface-Area Mesostructured Carbon for Enhanced CO2 Capture and Supercapacitor Performance. Angew. Chem., Int. Ed. 2016, 55, 2032−2036. (49) Lee, J.-S. M.; Briggs, M. E.; Hasell, T.; Cooper, A. I. Hyperporous Carbons from Hypercrosslinked Polymers. Adv. Mater. 2016, 28, 9804−9810. (50) Chen, Y.-Z.; Wang, C.; Wu, Z.-Y.; Xiong, Y.; Xu, Q.; Yu, S.-H.; Jiang, H.-L. From Bimetallic Metal-Organic Framework to Porous Carbon: High Surface Area and Multicomponent Active Dopants for Excellent Electrocatalysis. Adv. Mater. 2015, 27, 5010−5016. (51) Hu, H.; Han, L.; Yu, M. Z.; Wang, Z. Y.; Lou, X. W. Metalorganic-framework-engaged formation of Co nanoparticle-embedded carbon@Co9S8 double-shelled nanocages for efficient oxygen reduction. Energy Environ. Sci. 2016, 9, 107−111. (52) Dutta, S.; Bhaumik, A.; Wu, K. C. W. Hierarchically porous carbon derived from polymers and biomass: effect of interconnected pores on energy applications. Energy Environ. Sci. 2014, 7, 3574−3592. (53) Zhang, Y.; Liu, X.; Wang, S.; Dou, S.; Li, L. Interconnected honeycomb-like porous carbon derived from plane tree fluff for high performance supercapacitors. J. Mater. Chem. A 2016, 4, 10869− 10877. 7116

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117

Research Article

ACS Sustainable Chemistry & Engineering (54) Wang, R. T.; Wang, P. Y.; Yan, X. B.; Lang, J. W.; Peng, C.; Xue, Q. J. Promising Porous Carbon Derived from Celtuce Leaves with Outstanding Supercapacitance and CO2 Capture Performance. ACS Appl. Mater. Interfaces 2012, 4, 5800−5806. (55) Cai, Y.; Luo, Y.; Xiao, Y.; Zhao, X.; Liang, Y.; Hu, H.; Dong, H.; Sun, L.; Liu, Y.; Zheng, M. Facile Synthesis of Three-Dimensional Heteroatom-Doped and Hierarchical Egg-Box-Like Carbons Derived from Moringa oleifera Branches for High-Performance Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 33060−33071. (56) Wu, C.; Yang, S.; Cai, J.; Zhang, Q.; Zhu, Y.; Zhang, K. Activated Microporous Carbon Derived from Almond Shells for High Energy Density Asymmetric Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 15288−15296. (57) Guo, Y.; Yang, S.; Yu, K.; Zhao, J.; Wang, Z.; Xu, H. The preparation and mechanism studies of rice husk based porous carbon. Mater. Chem. Phys. 2002, 74, 320−323. (58) Chen, H. B.; Wang, H. B.; Yang, L. F.; Xiao, Y.; Zheng, M. T.; Liu, Y. L.; Fu, H. G. High Specific Surface Area Rice Hull Based Porous Carbon Prepared for EDLCs. Int. J. Electrochem. Sci. 2012, 7, 4889−4897. (59) Tabata, S.; Iida, H.; Horie, T.; Yamada, S. Hierarchical porous carbon from cell assemblies of rice husk for in vivo applications. MedChemComm 2010, 1, 136−138. (60) Zhang, L. L.; Zhao, X. S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 2520−2531. (61) Zhang, L.; You, T.; Zhou, T.; Zhou, X.; Xu, F. Interconnected Hierarchical Porous Carbon from Lignin-Derived Byproducts of Bioethanol Production for Ultra-High Performance Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 13918−13925. (62) Yu, Z. N.; Tetard, L.; Zhai, L.; Thomas, J. Supercapacitor electrode materials: nanostructures from 0 to 3 dimensions. Energy Environ. Sci. 2015, 8, 702−730. (63) Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845−854.

7117

DOI: 10.1021/acssuschemeng.7b01315 ACS Sustainable Chem. Eng. 2017, 5, 7111−7117