Synthesis of Diverse Green Carbon Nanomaterials through Fully

Jun 18, 2019 - The ultrahigh-surface-area carbon along with novel carbon materials ... 1583 cm–1, which were identified as the typical D-band (vibra...
0 downloads 0 Views 3MB Size
Research Article Cite This: ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

www.acsami.org

Synthesis of Diverse Green Carbon Nanomaterials through Fully Utilizing Biomass Carbon Source Assisted by KOH Zehao Zhu,† Yujing Liu,*,† Zhijin Ju,† Jianmin Luo,†,§ Ouwei Sheng,† Jianwei Nai,† Tiefeng Liu,† Yangxin Zhou,‡ Yao Wang,*,† and Xinyong Tao*,† †

College of Materials Science and Engineering, Zhejiang University of Technology, Hangzhou 310014, P. R. China Zhejiang Energy Group Research Institute, Hangzhou 310007, P. R. China



Downloaded via BUFFALO STATE on July 21, 2019 at 00:29:04 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: With multiple properties, green carbon nanomaterials with high specific surface area have become extensively attractive as energy storage devices with environmental-friendly features. The primary synthesis attempts were based on alkalis activation, which, however, faced the dilemma of low utilization rate of carbon sources. Herein, the green carbon with ultrahigh surface area (up to 3560 m2/g) was prepared by the KOHassisted biomass carbonization. Moreover, the redundant K2O steam and CxHy flow were further utilized; as a result, the carbon materials with a wide range of morphological diversity were collected on the Cu foam. Accordingly, we carried out density functional theory simulations to reveal the mechanism of Oadatom-promoted CH4 dissociation over the Cu surface for carbon formation. The electrodes of electrochemical capacitor fabricated by carbon synthesis possess a 170% higher specific capacitance compared with commercial carbon electrodes. As such, this strategy might be promising in developing hierarchical carbons along with sufficient carbon sources for broadening their potential applications. KEYWORDS: ultrahigh-surface-area green carbon, diversely morphological carbon, KOH, first-principle calculation, supercapacitor



INTRODUCTION Carbon nanomaterials with adjustable hierarchical nano/ microhole structures have been widely demonstrated to be applied in gas absorption,1−3 sewage treatment,4,5 supercapacitors,6−9 and lithium-ion batteries.10−13 Carbon nanomaterials with ultrahigh surface area were prepared through various ways.14−17 The three-dimensional (3D) porous nanonetwork composed of poly(styrene-co-divinylbenzene) was carbonized at 1000 °C, resulting in the powdery carbon aerogel with a high specific surface area of 2354 m2/g.18 Similarly, a porous matrix with a 5-fold interpenetrated topological network matrix was employed to prepare carbon with an ultrahigh surface area and the specific surface area of about 2066 m2/g.19 Apart from the carbonization of the polymer matrix, the activation from CO2 and H2O steam also contributed to the formation of the porous carbon. In a typical way, the CO 2 and H2 O steam can diffuse into the microstructures of carbon source, build, and enlarge multiple micropores by physical hole drilling and chemical reaction with carbon. Zeng et al. reacted hydroquinone and formaldehyde to prepare a carbon nanobelt through the one-pot method, achieving an ultrahigh surface area up to 3445 m2/g.11 The hierarchical porous carbon with polyacrylonitrile was regulated by the CO2 flow; as a result, the hierarchically porous graphitic carbon obtained a specific surface area of 2500 m2/g.20 Also, © 2019 American Chemical Society

the chemical activation by inorganic salts such as ZnCl2, KHCO3, and K2CO3 proved to be an efficient way for porous carbon synthesis.21 Carbon materials with a specific surface area of 2280 m2/g were prepared using sugarcane as carbon source and FeCl3 and ZnCl2 as activation agents.22 The nutshell doping with the ZnCl2 salt exhibited a specific surface area of 1718 m2/g after carbonization.23 In addition, the chemical activation by alkalis such as KOH was generally considered as the most effective approach to prepare nanocarbons including graphene, 24 carbon nanotubes (CNTs),25 and porous carbon.26 Compared with other methods, KOH activation exhibited an advantage in coordinating most of the carbon sources, the artificial polymer,27,28 fossil fuel,29,30 graphene hydrogel,31 biomass,32−35 and even organic waste.36,37 For instance, polyaniline,38 anthracite,39 papaya,40 and exfoliated graphite oxide41 were all successfully carbonized by KOH-induced activation resulting in carbon materials with an ultrahigh surface area. Especially, the biomass, as a low-cost and renewable carbon resource, was widely employed to obtain green carbon nanomaterials,42−50 particularly assisted by accessory alkaline ingredients.51−54 However, this method Received: May 14, 2019 Accepted: June 18, 2019 Published: June 18, 2019 24205

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

Research Article

ACS Applied Materials & Interfaces still met the problems of low carbon productive rate, as less than 10% of carbon sources could be utilized, since the hydrocarbon and alkyne flow emissions during carbonization would generate inevitable carbon loss. Meanwhile, the carbon emissions produced by pyrolyzation of carbon source also lead to excessive air pollution. Herein, we improved the synthetic technology to solve this problem. The ultrahigh-surface-area carbon along with novel carbon materials such as carbon horns, carbon tubes, and carbon black nanoparticles (NPs) were simultaneously obtained (Figure 1). The reaction was accomplished in a

Figure 2. (a) N2 adsorption−desorption isotherms of CFL under different active temperatures and times. (b) N2 adsorption− desorption isotherms of CFL and CFS under 900 °C for 1 h. The inset exhibits the distribution of pore diameter. (c) X-ray diffraction (XRD) patterns of CFL and CFS. (d) Raman spectra of CFL and CFS, and the scanning electron microscopy (SEM) images of (e) CFL and (f) CFS. Figure 1. Schematic illustration of the ultrahigh-surface-area carbon and diverse carbon nanomaterial synthesis.

after 1 h. However, while the active time further increased to 2 h, the typical type IV sorption isotherm suggested a lower BET specific surface area of 2274 m2/g, which was possibly ascribed to overactivation. Hence, increasing the active temperature and time could induce the structural transformation from microporous to mesoporous, and the highest specific surface area of carbon might stand at the initial status of the transformation, which was also verified in the previous work of high-surfacearea biomass carbon.58 Apart from lotus leaf, pig skin was also employed to prepare carbon nanomaterials. The carbon from the pig skin (CFS) exhibited an even higher specific surface area of 3560 m2/g compared with that from CFL (Figure 2b). Both carbons prepared from KOH blending biomass sources achieved a higher specific surface area compared with commercial carbon (CC) (1629 m2/g). The distribution of the pore diameter is also shown in Figure 2b, and the majority of pores possessed a diameter in the range of 2−3 nm. The X-ray diffraction (XRD) analysis of CFL and CFS exhibited two broad peaks, which corresponded to the (002) and (100) planes of graphite with low crystallinity (Figure 2c). Raman spectroscopy exhibited two characteristic peaks at 1350 and 1583 cm−1, which were identified as the typical D-band (vibrations of dangling bonded carbon atoms within amorphous graphite phase as in-plane terminations) and the G-band (vibration of sp2-bonded carbon atoms crystallized in the graphite hexagonal lattice), respectively. The morphologies of the CFL and CFS were examined by scanning electron microscopy (SEM) (Figure 2e,f). Both CFL and CFS turned out to be uniform-flakeshaped fragments with an average diameter of 15 μm. In addition, the surfaces of the as-grown carbons were fairly rough, which may originate from KOH etching. Particularly, the rough surfaces indicated a uniform porous structure, which resulted from the hole-drilling process. After preparing an ultrahigh-specific-surface-area carbon, we proceeded to focus on the abundant carbon source and heat wasted by the CxHy flow. The Cu foam was placed at the end

chemical vapor deposition (CVD) tube furnace and the hightemperature region provided the site where KOH and biomass reacted for ultrahigh-surface-area carbon formation. The lowtemperature region at the end of the tube furnace with a Cu collector was designed for collecting K2O- and CxHycontaining gas and waste heat, resulting in novel carbon material synthesis. The mechanism of O-adatom-promoted CH4 dissociation over Cu surface was further identified by density functional theory (DFT) calculations. The specific capacitance of the electrochemical capacitor electrodes fabricated by synthesized carbon was improved to as high as 170% compared with that of commercial carbon electrodes. In this case, we might shed light on the synthesis of green hierarchical carbons and enhance the utilization of carbon sources by reusing the CxHy flow.



RESULTS AND DISCUSSION The lotus leaves were selected as the biomass carbon source, which was further carbonized in the presence of KOH. According to the previous literature, the hydrocarbon produced by the biomass pyrolysis, such as CH4, C2H4, C2H6, and C3H6, could act as the carbon source for further carbonization assisted by the alkalis and/or inorganic salts.55−57 The permanent porosity of the as-grown carbon from lotus leaf (CFL) was examined using the Brunauer− Emmett−Teller (BET) model. The N2 adsorption−desorption analysis of the CFL samples (activated under 800 °C for 1 or 2 h) exhibited a typical type I sorption isotherm, indicating a microporous structure (Figure 2a). As the active temperature increased to 850 or 900 °C, the reversible type IV isotherm appeared, which was the primary characteristics of mesoporous materials. Typically, the calculated specific surface area of CFL increased from 3106 to 3450 m2/g as the active temperature and time increased. In this case, the highest specific surface area of 3450 m2/g was achieved at the condition of 900 °C 24206

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

Research Article

ACS Applied Materials & Interfaces

According to our previous research, K and/or O doping was necessary for carbon deposition.60,61 We further reconfirmed that in this case, our controlled experiments showed that no carbon materials can be obtained on the Cu collector without introducing K2O. To exemplify the mechanism of K2Oinduced carbon deposition on Cu, first-principles calculations were carried out to investigate the dissociation of the most common CxHy, i.e., CH4, over the Cu surface. As is known, the graphene growth generally occurs in three steps: dissociation of methane (CH4), migration of carbon atoms on the Cu surface, and nucleation of graphene nanoislands.62,63 It is generally considered that graphene growth conditions are controlled by methane dehydrogenation process, which is similar in this research. Previous experimental study64 has revealed that Cu(111) surface played a key role in the growth of chemical vapor deposition (CVD) graphene. Thus, we employed Cu(111) as a substrate to simulate the revolution of carbon in this study. The CH4 dissociation reaction on pristine Cu(111) is as follows

of the tube furnace. Surprisingly, we noticed that various carbon materials were obtained on the Cu collector. According to the SEM images of the carbons on Cu, a large amount of carbon horns, carbon tubes, and carbon black NPs were observed at the Cu surface (Figures 3 and S1). The carbon

CHx(ad) → CHx − 1(ad) + H(ad)

(1)

In our study, we mainly focus on the first dissociation step, i.e. CH4(ad) → CH3(ad) + H(ad)

Figure 3. SEM images of (a) carbon horns, (b) carbon tubes, and (c) carbon black NPs. The corresponding energy-dispersive X-ray spectroscopy maps of (d) carbon horns, (e) carbon tubes, and (f) carbon black NPs in which C, O, and K were identified. The scale bar represents 5 μm in (d) and 500 nm in (e) and (f). (g) TEM image of a carbon horn. (h) The high-resolution TEM image of zone arrowed in (g) showing the graphite lattice of (002).

(2)

and neglect the reaction of the hydrogen molecule formation process because CH4 dissociation reaction prefers direct dehydrogenation (H atom) to H abstraction (H2).65 Before discussing the effects of O and K adatoms on CH4 dissociation, the adsorption behaviors of O and K adatoms on the Cu(111) surface were investigated. As shown in Figure S5, the Cu(111) surface has four possible adsorption sites: top site (top), bridge site (bri), face-centered cubic site (fcc), and hexagonal close-packed site (hcp). We calculated the adsorption energy of O adatom at all of the sites and found that the fcc site exhibited the largest adsorption energy of −5.17 eV. Three Cu atoms coordinated with O adatom and had an average Cu−O bond length of 1.90 Å in the most stable configuration. However, the preabsorbed O atom did not affect the adsorption behavior of CH4 and the adsorption energy, also the adsorption site was almost the same as that on the pristine Cu(111) surface. The only difference was the significant charge transfer between O adatom and Cu substrate as shown in Figure S6, which led to the perturbation of the Cu−O bonds and the induction of a possible surface dipole.66 Nevertheless, there was no charge transfer between the CH4 molecule and pristine Cu substrate. On the contrary, the Cu(111) surface hardly adsorbed K atoms, suggesting that the effect of K atoms could be ignored. Our results showed that CHx radicals always had larger adsorption energies on the Cu(111) surface with O adatoms when compared with Ead on pristine Cu(111), as shown in Table S1. This suggests that O adatoms enhance CHx adsorption on Cu(111). Before the CH4 dissociation reaction, the distance between the CH4 molecule and pristine Cu(111) surface was 3.30 Å and the carbon atom was directly above the surface copper atom without chemical bonding, implying that the molecule was physically absorbed at the top site. The adsorption energy was −0.28 eV, close to the previous calculation results.67−69 The Cu−C distance reduced to 3.25 Å, while the adsorption energy slightly increased to −0.30 eV on the Cu(111) surface with preadsorbed O. The most stable adsorption sites of CHx radicals on the pristine and Cu(111) surfaces with preadsorbed O are shown in Table S1. The CHx

horn grew uniformly with a micrometer-sized hollow structure (Figure 3a). The carbon tubes exhibited an average diameter of 200 nm (Figure 3b), and the carbon black assembled as nanoparticles (120 nm) or even nanoparticle clusters (Figure 3c). Further, the transmission electron microscopy (TEM) clearly revealed the walls of horn and the characteristic (002) plane of the graphitized carbon (Figure 3g,h). Combined with XRD and Raman analysis, the TEM imaging demonstrated that the synthesized carbon possessed localized graphitization. It is noticed that the morphologies of the as-grown carbon nanomaterials were quite different from the widely reported CNTs catalyzed by KOH-decorated metal.25 The carbon deposition originating from the pyrolyzation of carbonenriched flow was particularly regulated by the catalyst nanocrystals (e.g., transition metals or their alloys),25,59 especially the interface between catalyst surface and solid carbon.60 Hence, the carbon growth was dominated and limited on certain crystal facets, resulting in various carbon morphologies induced by different crystallization orientations of the uncontrollable metallic nanoparticles.60,61 The carbon collected on the Cu foam was quantified, and the weight ratio of carbon nanomaterials to biomass source was calculated as 0.201, indicative of the 20.1% improvement of carbon sources utilization. Furthermore, in addition to carbon element, K and O were found to be uniformly distributed among carbon horns (Figure 3d). Similarly, the carbon tubes and carbon black NPs also exhibited typical crystalline features of graphite and the strong elemental signals of K and O (Figures 3e,f and S3, S4). Therefore, we speculated that K- and/or O-induced Cu catalyst might contribute to the basic growth of the diverse carbon. 24207

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

Research Article

ACS Applied Materials & Interfaces

mechanism. During the transition of pristine Cu(111), the active Cu center acts as an electron donor, which can transfer charge carriers from Cu to the bonding CH3 and unfilled antibonding C−H. Therefore, both CH3 and H adatoms act as acceptors and gain electrons from the intercalation Cu atom, leading to electron repulsion. This results in high energy barrier as well as low reaction activity for the CH4 dissociation reaction. As we discussed above, the high electronegativity of O tends to introduce significant electron transfer. The O adatom could attract the broken H atom, thus the hydrogen atom loses charge, while CH3 gains charge from Cu, as shown in Figure. S6. Therefore, there is a strong attractive interaction between CH3 and H with the assistance of the O adatom. According to the activation strain model, the reaction energy barrier can be divided into two parts: the activation strain energy with a positive value leading to the increase in the reaction energy barrier and the dipole−dipole interaction energy with a negative value, giving rise to reduced reaction energy barrier.70,71 After O is absorbed on Cu (111), the C−H bond becomes shorter, impling lower strain energy, which contributes in lowering the reaction energy barrier. Meanwhile, the stronger interaction between CH3 and H causes a greater negative value of the interaction energy and further reduces the reaction energy barrier. Consequently, the O adatom is conducive in reducing the CH4 dissociation energy barrier and enhancing the reaction activity. Subsequently, the CFL with an ultrahigh surface area was fabricated into the electrodes of the electrochemical capacitor to explore their potential applications. The galvanostatic charge/discharge curves of the electrochemical capacitor built by CFL and CC are presented in Figure 5a. Importantly,

molecules would rather be absorbed at the hollow sites than at the top and bridge sites; these results show good consistency with prior studies.68,69 These adsorption configurations arise from the Cu−C bonding characteristics, and hollow sites would generate more Cu−C bonds at the hollows and facilitate CHx adsorption. Thus, the most stable hollow sites were selected as the start point and the end point in the calculation of the CH4 dissociation reaction path. Next, we performed nudged elastic band calculations along the possible CH4 dissociation reaction path between the initial states and final states to search the transition states on the pristine Cu(111) surface and that with preabsorbed O. Figure 4a illustrates the geometries of the transition states and energy

Figure 4. (a) Reaction pathways and energy barrier for the CH4 dissociation reaction path on the pristine Cu(111) surface and that with preabsorbed O. Transition states of (b) the pristine Cu(111) surface and (c) that with preabsorbed O, where Hbro denotes the broken H atom. The net Bader charge values are highlighted in red.

barriers. After CH4 was adsorbed on the pristine Cu(111), a C−H bond of CH4 ruptured while methyl molecule and hydrogen atoms were obtained. The CH3 and H atom were absorbed on the same Cu atom, indicating that the C−H bond perturbed with Cu intercalation. The C atom and broken H atom are 2.13 and 1.62 Å away from the Cu atom, respectively, as shown in Figure 4b. CH3 moved from the top site to the fcc site, while the broken H atom diffused to the fcc site after CH4 dissociation reaction. The energy barrier is 1.51 eV, and this reaction is endothermic. When O was absorbed on the Cu(111) surface, the CH4 molecule was absorbed on the top site, similar to results on pristine Cu(111). Figure 4c illustrates that O adatom activated the C−H bond because the adsorption site changed from a fcc hollow site to a bridge site. During the transition, the C−H bond was ruptured and the displaced hydrogen atom bonded with C atom and O atom, with the bond lengths of 1.42 and 1.25 Å, respectively. In the final state, CH3 moved to the hcp site and OH was adsorbed on the fcc site. The energy barrier of 1.15 eV was lower than that on the pristine Cu(111) surface (1.51 eV), accompanied by the release of 0.03 eV of energy. Comparing Figures S7 and S8, we observe that the occurrence of O adatom leads to a lower reaction energy barrier in every CHx dissociation reaction step, indicating the catalytic activity of O atom. As such, the simulation suggests that the CH 4 dissociation reaction activity on the pristine Cu(111) was enhanced by O adatoms. We investigated the progress of electron transfer in these reactions to better understand the mechanism of the high activity of the Cu(111) system with preabsorbed O. The net Bader charge was calculated to detect the activation

Figure 5. (a) Galvanostatic charge/discharge curves of the electrochemical capacitor fabricated by CFL and CC in the first, second, and third cycle. (b) Corresponding cycling performance at a current density of 0.2 A/g.

the electrode with CFL exhibited much higher specific capacitances (up to 106 F/g in the first cycle) than CC (up to 39 F/g in the first cycle). Indeed, the specific capacitances of the electrode with CFL still remained as high as 79 F/g after 100 cycles under the current density of 0.2 A/g, while the electrode with CC only possessed a specific capacitance of 26 F/g. This enhancement can be attributed to the large specific surface area and the hierarchical porous structure of the CFL. The optimization of the CFL-based electrodes together with the mechanism of how the carbon structure influences the Coulombic efficiency of the electrochemical capacitor still needs to be further investigated. The fairly high specific capacitance and the stable cyclic performance of the electrochemical capacitor built by CFL compared with the CC indicated that the synthesized hierarchical ultrahigh-surfacearea carbon would have possible potential in advanced electrochemical capacitors. 24208

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

Research Article

ACS Applied Materials & Interfaces



Molecular Engineering of High-Performance Crystalline Porous Materials. Nat. Commun. 2016, 7, No. 13645. (3) Jalilov, A. S.; Li, Y.; Tian, J.; Tour, J. M. Ultra-High Surface Area Activated Porous Asphalt for CO2 Capture through Competitive Adsorption at High Pressures. Adv. Energy Mater. 2017, 7, No. 1600693. (4) Bolisetty, S.; Mezzenga, R. Amyloid-Carbon Hybrid Membranes for Universal Water Purification. Nat. Nanotechnol. 2016, 11, 365− 371. (5) Zhu, Z. G.; Ji, C. H.; Zhong, L. L.; Liu, S.; Cui, F.; Sun, H. L.; Wang, W. Magnetic Fe−Co Crystal Doped Hierarchical Porous Carbon Fibers for Removal of Organic Pollutants. J. Mater. Chem. A 2017, 5, 18071−18080. (6) Lin, T.; Chen, I. W.; Liu, F. X.; Yang, C. Y.; Bi, H.; Xu, F. F.; Huang, F. Q. Nitrogen-Doped Mesoporous Carbon of Extraordinary Capacitance for Electrochemical Energy Storage. Science 2015, 350, 1508−1513. (7) Chen, L.-F.; Lu, Y.; Yu, L.; Lou, X. W. Designed Formation of Hollow Particle-Based Nitrogen-Doped Carbon Nanofibers for HighPerformance Supercapacitors. Energy Environ. Sci. 2017, 10, 1777− 1783. (8) Chen, S. R.; Liang, J. Y.; Zhao, J. X.; Li, H. F.; Liu, Y.; Deng, H.; Liu, C.; Tang, S.; Cao, Y. C. Preparation of Nitrogen Doped ClewsLike Carbon Materials and Their Application As the Electrode in Supercapacitor. Chin. Chem. Lett. 2019, 30, 787−791. (9) Guan, T. T.; Li, K. X.; Zhao, J. H.; Zhao, R. J.; Zhang, G. L.; Zhang, D. D.; Wang, J. L. Template-Free Preparation of LayerStacked Hierarchical Porous Carbons from Coal Tar Pitch for High Performance All-Solid-State Supercapacitors. J. Mater. Chem. A 2017, 5, 15869−15878. (10) Li, Z.; Zhang, J.; Guan, B. Y.; Lou, X. W. D. Mesoporous Carbon@Titanium Nitride Hollow Spheres as an Efficient SeS2 Host for Advanced Li-SeS2 Batteries. Angew. Chem., Int. Ed. 2017, 56, 16003−16007. (11) Zeng, S.-Z.; Zeng, X. R.; Tu, W. X.; Yao, Y. C.; Yu, L.; Wu, H. L.; Jin, W. W.; Huang, H. T.; Zou, J. Z. Facile and Tailored Synthesis of Ultrahigh-Surface-Area Clews of Carbon Nanobelts for High-Rate Lithium−Sulfur Batteries. J. Mater. Chem. A 2017, 5, 23209−23220. (12) Mo, R.; Rooney, D.; Sun, K.; Yang, H. Y. 3D Nitrogen-Doped Graphene Foam with Encapsulated Germanium/Nitrogen-Doped Graphene Yolk-Shell Nanoarchitecture for High-Performance Flexible Li-ion Battery. Nat. Commun. 2017, 8, No. 13949. (13) Zhou, X. S.; Yu, L.; Yu, X.-Y.; Lou, X. W. D. Encapsulating Sn Nanoparticles in Amorphous Carbon Nanotubes for Enhanced Lithium Storage Properties. Adv. Energy Mater. 2016, 6, No. 1601177. (14) Kong, B.; Zu, L. H.; Peng, C. X.; Zhang, Y.; Zhang, W.; Tang, J.; Cordelia, S.; Zhang, L. D.; Chen, H. X.; Wang, Y.; Liu, Y.; He, H. L.; Wei, J.; Lin, X. C.; Luo, W.; Yang, J. P.; Zhao, Z. W.; Liu, Y.; Yang, J. H.; Zhao, D. Y. Direct Superassemblies of Freestanding MetalCarbon Frameworks Featuring Reversible Crystalline-Phase Transformation for Electrochemical Sodium Storage. J. Am. Chem. Soc. 2016, 138, 16533−16541. (15) Kong, B.; Tang, J.; Zhang, Y. Y.; Jiang, T.; Gong, X. G.; Peng, C. X.; Wei, J.; Yang, J. P.; Wang, Y. C.; Wang, X. B.; Zheng, G. F.; Cordelia, S.; Zhao, D. Y. Incorporation of Well-Dispersed Sub-5-nm Graphitic Pencil Nanodots into Ordered Mesoporous Frameworks. Nat. Chem. 2016, 8, 171−178. (16) Kong, B.; Tang, J.; Cordelia, S.; Li, W.; Wei, J.; Fang, Y.; Wang, Y. C.; Zheng, G. F.; Zhao, D. Y. Oriented Mesoporous Nanopyramids as Versatile Plasmon-Enhanced Interfaces. J. Am. Chem. Soc. 2014, 136, 6822−6825. (17) Kong, B.; Tang, J.; Zhang, Y. Y.; Cordelia, S.; Gong, X. G.; Liu, Y.; Zhang, W.; Yang, J. P.; Wang, W. S.; Sun, X. T.; Wang, Y. F.; Zheng, G. F.; Zhao, D. Y. Branched Artificial Nanofinger Arrays by Mesoporous Interfacial Atomic Rearrangement. J. Am. Chem. Soc. 2015, 137, 4260−4266. (18) Zheng, J.; Huang, J. L.; Xu, F.; Zhu, F.; Wu, D. C.; Ouyang, G. F. Powdery Polymer and Carbon Aerogels with High Surface Areas

CONCLUSIONS This work offered a green and efficient method to synthesize ultrahigh-surface-area (as high as 3560 m2/g) carbon and novel carbon materials with the aim to fully utilize the redundant CxHy flow and the waste heat. SEM, TEM, Raman, and XRD analysis were employed to evaluate the structures and morphologies of diverse carbon materials. Moreover, the mechanism of O adatoms’ promotion of carbon formation over the Cu surface was demonstrated by the DFT calculations. Finally, the obtained ultrahigh-surface-area carbon was preliminarily applied in electrochemical capacitor and exhibited a higher specific capacitance and cyclic stability compared with that of CC. As such, these novel carbon materials might realize fascinating potential in energy storage, especially for electrochemical capacitors.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.9b08420.



Experimental sections describing the synthesis of carbon nanomaterials, their characterization, electrochemical measurements, and computational details (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (Y.L.). *E-mail: [email protected] (Y.W.). *E-mail: [email protected] (X.T.). ORCID

Jianwei Nai: 0000-0001-9681-5498 Xinyong Tao: 0000-0002-6233-4140 Present Address §

Thayer School of Engineering, Dartmouth College, 14 Engineering Drive, Hanover, New Hampshire 03755, United States (J.L.).

Author Contributions

The manuscript was written through contributions of all of the authors. All of the authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge the financial support by the National Natural Science Foundation of China (Grant nos. 51722210, 51677170, 51777194, 51572240, U1802254, and 51871201), Natural Science Foundation of Zhejiang Province (Grant nos. LD18E020003, LY16E070004, LY18B030008, and LY17E020010), Zhejiang Provincial Research and Development Program (Grant no. 2018C01G6081111), and Zhejiang Energy Group R&D (Grant no. ZNKJ-2017-069).



REFERENCES

(1) Belmabkhout, Y.; Bhatt, P. M.; Adil, K.; Pillai, R. S.; Cadiau, A.; Shkurenko, A.; Maurin, G.; Liu, G.; Koros, W. J.; Eddaoudi, M. Natural Gas Upgrading Using a Fluorinated MOF with Tuned H2S and CO2 Adsorption Selectivity. Nat. Energy 2018, 3, 1059−1066. (2) Zhai, Q. G.; Bu, X.; Mao, C.; Zhao, X.; Daemen, L.; Cheng, Y.; Ramirez-Cuesta, A. J.; Feng, P. An Ultra-Tunable Platform for 24209

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

Research Article

ACS Applied Materials & Interfaces for High-Performance Solid Phase Microextraction Coatings. Nanoscale 2017, 9, 5545−5550. (19) Hu, F.; Liu, C.; Wu, M.; Pang, J.; Jiang, F.; Yuan, D.; Hong, M. An Ultrastable and Easily Regenerated Hydrogen-Bonded Organic Molecular Framework with Permanent Porosity. Angew. Chem., Int. Ed. 2017, 56, 2101−2104. (20) Estevez, L.; Prabhakaran, V.; Garcia, A. L.; Shin, Y.; Tao, J.; Schwarz, A. M.; Darsell, J.; Bhattacharya, P.; Shutthanandan, V.; Zhang, J. G. Hierarchically Porous Graphitic Carbon with Simultaneously High Surface Area and Colossal Pore Volume Engineered via Ice Templating. ACS Nano 2017, 11, 11047−11055. (21) Sevilla, M.; Fuertes, A. B. A Green Approach to HighPerformance Supercapacitor Electrodes: The Chemical Activation of Hydrochar with Potassium Bicarbonate. ChemSusChem 2016, 9, 1880−1888. (22) Xiao, X.; Chen, B.; Zhu, L.; Schnoor, J. L. Sugar CaneConverted Graphene-Like Material for the Superhigh Adsorption of Organic Pollutants from Water via Coassembly Mechanisms. Environ. Sci. Technol. 2017, 51, 12644−12652. (23) Ahmadpour, A.; Do, D. D. The Preparation of Activated Carbon from Macadamia Nutshell by Chemical Activation. Carbon 1997, 35, 1723−1732. (24) Zhao, Y. F.; Huang, S. F.; Xia, M. R.; Sarish, R.; Mu, S. C.; Kou, Z. K.; Zhang, Z.; Chen, Z. Y.; Gao, F. M.; Hou, Y. L. N-P-O CoDoped High Performance 3D Graphene Prepared through Red Phosphorous-Assisted “Cutting-Thin” Technique: A Universal Synthesis and Multifunctional Applications. Nano Energy 2016, 28, 346− 355. (25) Zhang, Z.; Mu, S. C.; Zhang, B. W.; Tao, L.; Huang, S. F.; Huang, Y. Z.; Gao, F. M.; Zhao, Y. F. A Novel Synthesis of Carbon Nanotubes Directly from an Indecomposable Solid Carbon Source for Electrochemical Applications. J. Mater. Chem. A 2016, 4, 2137−2146. (26) Zhao, Y.; Ran, W.; He, J.; Song, Y. F.; Zhang, C. M.; Xiong, D.B.; Gao, F. M.; Wu, J. S.; Xia, Y. Y. Oxygen-Rich Hierarchical Porous Carbon Derived from Artemia Cyst Shells with Superior Electrochemical Performance. ACS Appl. Mater. Interfaces 2015, 7, 1132− 1139. (27) Jiang, J. M.; Nie, P.; Ding, B.; Zhang, Y. D.; Xu, G. Y.; Wu, L. Y.; Dou, H.; Zhang, X. G. Highly Stable Lithium ion Capacitor Enabled by Hierarchical Polyimide Derived Carbon Microspheres Combined with 3D Current Collectors. J. Mater. Chem. A 2017, 5, 23283−23291. (28) Zhou, X.; Wang, P. L.; Zhang, Y. G.; Wang, L. L.; Zhang, L. T.; Zhang, L.; Xu, L.; Liu, L. Biomass Based Nitrogen-Doped StructureTunable Versatile Porous Carbon Materials. J. Mater. Chem. A 2017, 5, 12958−12968. (29) Lozano-Castelló, D.; Calo, J. M.; Cazorla-Amorós, D.; LinaresSolano, A. Carbon Activation with KOH as Explored by Temperature Programmed Techniques, and The Effects of Hydrogen. Carbon 2007, 45, 2529−2536. (30) Zhai, D. Y.; Li, B. H.; Kang, F. Y.; Du, H. D.; Xu, C. J. Preparation of Mesophase-Pitch-Based Activated Carbons for Electric Double Layer Capacitors with High Energy Density. Microporous Mesoporous Mater. 2010, 130, 224−228. (31) Li, H.; Tao, Y.; Zheng, X. Y.; Li, Z. J.; Liu, D. H.; Xu, Z.; Luo, C.; Luo, J. Y.; Kang, F. Y.; Yang, Q.-H. Compressed Porous Graphene Particles for Use as Supercapacitor Electrodes with Excellent Volumetric Performance. Nanoscale 2015, 7, 18459−18463. (32) Zeng, J.; Wei, L.; Guo, X. Bio-Inspired High-Performance Solid-State Supercapacitors with The Electrolyte, Separator, Binder and Electrodes Entirely from Kelp. J. Mater. Chem. A 2017, 5, 25282− 25292. (33) Genovese, M.; Lian, K. Polyoxometalate Modified Pine Cone Biochar Carbon for Supercapacitor Electrodes. J. Mater. Chem. A 2017, 5, 3939−3947. (34) Cui, J. F.; Xi, Y. L.; Chen, S.; Li, D. H.; She, X. L.; Sun, J.; Han, W.; Yang, D. J.; Guo, S. J. Prolifera-Green-Tide as Sustainable Source for Carbonaceous Aerogels with Hierarchical Pore to Achieve Multiple Energy Storage. Adv. Funct. Mater. 2016, 26, 8487−8495.

(35) Li, X.; Xing, W.; Zhuo, S.; Zhou, J.; Li, F.; Qiao, S. Z.; Lu, G. Q. Preparation of Capacitor’s Electrode from Sunflower Seed Shell. Bioresour. Technol. 2011, 102, 1118−1123. (36) Blankenship, T. S.; Mokaya, R. Cigarette Butt-Derived Carbons Have Ultra-High Surface Area and Unprecedented Hydrogen Storage Capacity. Energy Environ. Sci. 2017, 10, 2552−2562. (37) Sun, H. X.; La, P. Q.; Yang, R. X.; Zhu, Z. Q.; Liang, W. D.; Yang, B. P.; Li, A.; Deng, W. Q. Innovative Nanoporous Carbons with Ultrahigh Uptakes for Capture and Reversible Storage of CO2 and Volatile Iodine. J. Hazard. Mater. 2017, 321, 210−217. (38) He, J. J.; To, J. W. F.; Psarras, P. C.; Yan, H. P.; Atkinson, T.; Holmes, R. T.; Nordlund, D.; Bao, Z.; Wilcox, J. Tunable PolyanilineBased Porous Carbon with Ultrahigh Surface Area for CO2 Capture at Elevated Pressure. Adv. Energy Mater. 2016, 6, No. 1502491. (39) Lozano-Castelló, D.; Cazorla-Amorós, D.; Linares-Solano, A.; Shiraishi, S.; Kurihara, H.; Oya, A. Influence of Pore Structure and Surface Chemistry on Electric Double Layer Capacitance in NonAqueous Electrolyte. Carbon 2003, 41, 1765−1775. (40) Wang, S. J.; Wang, R. T.; Zhang, Y. B.; Zhang, L. Highly Porous Carbon with Large Electrochemical Ion Absorption Capability for High-Performance Supercapacitors and Ion Capacitors. Nanotechnology 2017, 28, No. 445406. (41) 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. (42) Xia, Y.; Xiao, Z.; Dou, X.; Huang, H.; Lu, X. H.; Yan, R. J.; Gan, Y. P.; Zhu, W. J.; Tu, J. P.; Zhang, W. K.; Tao, X. Y. Green and Facile Fabrication of Hollow Porous MnO/C Microspheres from Microalgaes for Lithium-Ion Batteries. ACS Nano 2013, 7, 7083−7092. (43) Xia, Y.; Zhang, W. K.; Xiao, Z.; Huang, H.; Zeng, H. J.; Chen, X. R.; Chen, F.; Gan, Y. P.; Tao, X. Y. Biotemplated Fabrication of Hierarchically Porous NiO/C Composite from Lotus Pollen Grains for Lithium-Ion Batteries. J. Mater. Chem. 2012, 22, 9209−9215. (44) Xia, Y.; Zhang, W. K.; Huang, H.; Gan, Y. P.; Xiao, Z.; Qian, L. C.; Tao, X. Y. Biotemplating of Phosphate Hierarchical Rechargeable LiFePO4/C Spirulina Microstructures. J. Mater. Chem. 2011, 21, 6498−6501. (45) Tao, X. Y.; Zhang, J. T.; Xia, Y.; Huang, H.; Du, J.; Xiao, H.; Zhang, W. K.; Gan, Y. P. Bio-Inspired Fabrication of Carbon Nanotiles for High Performance Cathode of Li-S Batteries. J. Mater. Chem. A 2014, 2, 2290−2296. (46) Tao, X. Y.; Chai, W. C.; Xu, F. Q.; Luo, J. M.; Xiao, H.; Liang, C.; Gan, Y. P.; Huang, H.; Xia, Y.; Zhang, W. K. Bio-Templated Fabrication of Highly Defective Carbon Anchored MnO Anode Materials with High Reversible Capacity. Electrochim. Acta 2015, 169, 159−167. (47) Jin, C. B.; Sheng, O. W.; Luo, J. M.; Yuan, H. D.; Fang, C.; Zhang, W. K.; Huang, H.; Gan, Y. P.; Xia, Y.; Liang, C.; Zhang, J.; Tao, X. Y. 3D Lithium Metal Embedded Within Lithiophilic Porous Matrix for Stable Lithium Metal Batteries. Nano Energy 2017, 37, 177−186. (48) Sheng, O. W.; Jin, C. B.; Luo, J. M.; Yuan, H. D.; Fang, C.; Huang, H.; Gan, Y. P.; Zhang, J.; Xia, Y.; Liang, C.; Zhang, W. K.; Tao, X. Y. Ionic Conductivity Promotion of Polymer Electrolyte with Ionic Liquid Grafted Oxides for All-Solid-State Lithium−Sulfur Batteries. J. Mater. Chem. A 2017, 5, 12934−12942. (49) Jin, C. B.; Sheng, O. W.; Lu, Y.; Luo, J. M.; Yuan, H. D.; Zhang, W. K.; Huang, H.; Gan, Y. P.; Xia, Y.; Liang, C.; Zhang, J.; Tao, X. Y. Metal Oxide Nanoparticles Induced Step-Edge Nucleation of Stable Li Metal Anode Working Under an Ultrahigh Current Density of 15 mA cm−2. Nano Energy 2018, 45, 203−209. (50) Jin, C. B.; Sheng, O. W.; Zhang, W. K.; Luo, J. M.; Yuan, H. D.; Yang, T.; Huang, H.; Gan, Y. P.; Xia, Y.; Liang, C.; Zhang, J.; Tao, X. Y. Sustainable, Inexpensive, Naturally Multi-Functionalized Biomass Carbon for Both Li Metal Anode and Sulfur Cathode. Energy Storage Mater. 2018, 15, 218−225. (51) Wei, L.; Sevilla, M.; Fuertes, A. B.; Mokaya, R.; Yushin, G. Hydrothermal Carbonization of Abundant Renewable Natural 24210

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211

Research Article

ACS Applied Materials & Interfaces Organic Chemicals for High-Performance Supercapacitor Electrodes. Adv. Energy Mater. 2011, 1, 356−361. (52) Chen, W. X.; Zhang, H.; Huang, Y. Q.; Wang, W. K. A Fish Scale Based Hierarchical Lamellar Porous Carbon Material Obtained Using A Natural Template for High Performance Electrochemical Capacitors. J. Mater. Chem. 2010, 20, 4773−4775. (53) Wang, P.; Ye, H.; Yin, Y. X.; Chen, H.; Bian, Y. B.; Wang, Z. R.; Cao, F. F.; Guo, Y. G. Fungi-Enabled Synthesis of Ultrahigh-SurfaceArea Porous Carbon. Adv. Mater. 2019, 31, No. 1805134. (54) Li, D. H.; Chang, G. J.; Zong, L.; Xue, P.; Wang, Y. Z.; Xia, Y. Z.; Lai, C.; Yang, D. J. From Double-Helix Structured Seaweed to SDoped Carbon Aerogel with Ultra-High Surface Area for Energy Storage. Energy Storage Mater. 2019, 17, 22−30. (55) Tao, X. Y.; Du, J.; Yang, Y. C.; Li, Y. P.; Xia, Y.; Gan, Y. P.; Huang, H.; Zhang, W. K.; Li, X. D. TiC Nanorods Derived from Cotton Fibers: Chloride-Assisted VLS Growth, Structure, and Mechanical Properties. Cryst. Growth Des. 2011, 11, 4422−4426. (56) Wang, J.; Kaskel, S. KOH Activation of Carbon-Based Materials for Energy Storage. J. Mater. Chem. 2012, 22, 23710. (57) Lozano-Castelló, D.; Lillo-Rodenas, M. A.; Cazorla-Amoros, D.; Linares-Solano, A. Preparation of Activated Carbons from Spanish Anthracite I. Activation by KOH. Carbon 2001, 39, 741−749. (58) Xia, X. H.; Liu, H. B.; Shi, L.; He, Y. D. Tobacco Stem-Based Activated Carbons for High Performance Supercapacitors. J. Mater. Eng. Perform. 2012, 21, 1956−1961. (59) Rodriguez, N. M.; Chambers, A.; Baker, R. T. K. Catalytic Engineering of Carbon Nanostructures. Langmuir 1995, 11, 3862− 3866. (60) Tao, X. Y.; Zhang, X. B.; Zhang, L.; Cheng, J. P.; Liu, F.; Luo, J. H.; Luo, Z. Q.; Geise, H. J. Synthesis of Multi-Branched Porous Carbon Nanofibers and Their Application in Electrochemical DoubleLayer Capacitors. Carbon 2006, 44, 1425−1428. (61) Tao, X. Y.; Zhang, X. B.; Cheng, J. P.; Wang, Y. W.; Liu, F.; Luo, Z. Q. Synthesis of Novel Multi-Branched Carbon Nanotubes with Alkali-Element Modified Cu/MgO Catalyst. Chem. Phys. Lett. 2005, 409, 89−92. (62) Ago, H.; Ogawa, Y.; Tsuji, M.; Mizuno, S.; Hibino, H. Catalytic Growth of Graphene: Toward Large-Area Single-Crystalline Graphene. J. Phys. Chem. Lett. 2012, 3, 2228−2236. (63) Zhang, X. Y.; Xu, Z. W.; Hui, L.; Xin, J.; Ding, F. How the Orientation of Graphene Is Determined during Chemical Vapor Deposition Growth. J. Phys. Chem. Lett. 2012, 3, 2822−2827. (64) Wood, J. D.; Schmucker, S. W.; Lyons, A. S.; Pop, E.; Lyding, J. W. Effects of Polycrystalline Cu Substrate on Graphene Growth by Chemical Vapor Deposition. Nano Lett. 2011, 11, 4547−4554. (65) Li, K.; He, C. Z.; Jiao, M. G.; Wang, Y.; Wu, Z. J. A FirstPrinciples Study on The Role of Hydrogen in Early Stage of Graphene Growth During the CH4 Dissociation on Cu(111) and Ni(111) Surfaces. Carbon 2014, 74, 255−265. (66) Soon, A.; Todorova, M.; Delley, B.; Stampfl, C. Oxygen Adsorption and Stability of Surface Oxides on Cu(111): A FirstPrinciples Investigation. Phys. Rev. B 2006, 73, No. 165424. (67) Niu, T. C.; Jiang, Z.; Zhu, Y. G.; Zhou, G. W.; van Spronsen, M. A.; Tenney, S. A.; Boscoboinik, J. A.; Stacchiola, D. OxygenPromoted Methane Activation on Copper. J. Phys. Chem. B 2018, 122, 855−863. (68) Wang, J.; Wang, G. C. Promotion effect of methane activation on Cu (111) by the surface-active oxygen species: a combination of DFT and ReaxFF study. J. Phys. Chem. C 2018, 122, 17338−17346. (69) Gajewski, G.; Pao, C. W. Ab Initio Calculations of The Reaction Pathways for Methane Decomposition Over The Cu (111) Surface. J. Chem. Phys. 2011, 135, No. 064707. (70) de Jong, G. T.; Bickelhaupt, F. M. Transition-State Energy and Position Along The Reaction Coordinate in An Extended Activation Strain Model. ChemPhysChem 2007, 8, 1170−1181. (71) Fernández, I.; Bickelhaupt, F. M. The Activation Strain Model and Molecular Orbital Theory: Understanding and Designing Chemical Reactions. Chem. Soc. Rev. 2014, 43, 4953−4967.

24211

DOI: 10.1021/acsami.9b08420 ACS Appl. Mater. Interfaces 2019, 11, 24205−24211