Synthesis of a Carbonaceous Two-Dimensional (2D) Material - ACS

5 hours ago - Despite tremendous accomplishment achieved in 2D materials, little progress has been made in carbonaceous 2D material beyond graphene ...
0 downloads 0 Views 666KB Size
Subscriber access provided by UNIV OF LOUISIANA

Letter

Synthesis of a Carbonaceous Two-Dimensional (2D) Material Taewoo Kim, Jun Ho Lee, Geonhui Lee, Jaewoo Lee, Hyelynn Song, Jae Young Jho, Hong H. Lee, and Yong Hyup Kim ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.9b01808 • Publication Date (Web): 23 Apr 2019 Downloaded from http://pubs.acs.org on April 23, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Synthesis of a Carbonaceous Two-Dimensional (2D) Material Taewoo Kim,†,‡ Junho Lee,,‡ Geonhui Lee, ,‡ Jaewoo Lee,  Hyelynn Song,  Jae Young Jho,‖ Hong H. Lee,*,‖ and Yong Hyup Kim*, †Department School ‖School

of Mechanical Engineering, Incheon National University, Incheon 22012, South Korea

of Mechanical and Aerospace Engineering, Seoul National University, Seoul 08826, South Korea of Chemical and Biological Engineering, Seoul National University, Seoul 08826, South Korea

ABSTRACT: Despite tremendous accomplishment achieved in 2D materials, little progress has been made in carbonaceous 2D material beyond graphene and graphene oxide. Here we report a 2D material of carbonaceous nanoplates (CANP). The bottom-up synthesis of CANP is green, separation-free, and massive. The nanoplates are two to three monolayers thick with an average interlayer spacing of 0.57 nm. The synthesis involves viscosity-aided two-dimensional growth of fragmented glucose derivatives, and leads to complete conversion of glucose to the 2D nanoplates. Application tests demonstrate usefulness of the affordable 2D material. KEYWORDS: two-dimensional materials, two-dimensional growth, carbonaceous plates, furan-arene oxide, hydrothermal process solution for hydrothermal conversion. In this way, the water in the glucose solution in the petri dish can be evaporated during the hydrothermal treatment so that the concentration of glucose solution can still be increased for still higher viscosity. To distinguish from the usual liquid-filled hydrothermal process, it is to be called ‘vapor-filled’ (VF) hydrothermal process. Figure 1b shows the damp solid mat of CANP that was taken out of the device at the end of the reaction. The synthesis is massive in the sense that not just a part of glucose but the whole of it is converted to the 2D material. When the device was cooled to room temperature, it was still pressurized, indicating that gaseous byproducts formed. They were identified as carbon monoxide, carbon dioxide, methane, and ethane (Fig. S2). No trace of glucose could be identified by Fourier transform infrared spectroscopy (FT-IR), indicating that the whole glucose was completely converted to the carbonaceous nanoplates, water, and gaseous byproducts. When 16 g of glucose was used, 7.76 g of CANP was synthesized, revealing that the yield to CANP is 48.5% by weight. The nanoplates of CANP mat could be well dispersed in deionized water and organic solvent of N-Methyl-2pyrrolidone (NMP) by ultra-sonication, as shown in Fig. 1c.

The advent of graphene has generated a great deal of interest in two-dimensional (2D) materials, eventually leading to the discovery of various 2D materials such as transition metal dichalcogenides, layered oxides, monoelemental semiconductors, and other 2D compounds.1-2 Great stride achieved in these non-carbon 2D materials contrasts little progress made in carbonaceous 2D materials beyond graphene and graphene oxide.3-9 Motivated by the outstanding properties that the carbon-based 2D materials of graphene and, in particular, graphene oxide (GO) possesses,9-10 we began exploring possible ways of synthesizing a carbonaceous 2D material. A promising candidate for synthesizing carbon materials is the hydrothermal carbonization process, but it mainly fabricates zero- and one-dimensional structures.4 More recently, several novel methods have been reported for the synthesis of 2D carbon materials from glucose using moltensalt5-6 or temporary layered graphic carbon nitride.7-8 In this work, we report the synthesis of a 2D material of carbonaceous nanoplates (CANP) that is green, separationfree, and massive. Unlike the previous papers synthesizing 2D carbon materials,5-8 CANP is synthesized from only glucose and water without the aid of any other materials. Glucose is completely converted to CANP in a specially devised hydrothermal reactor through a two dimensional growth that is induced by high viscosity, induced in the sense that no 2D growth occurs at low viscosity. We characterize CANP physically and chemically. The amorphous planar structure of CANP consists of mainly furan and benzene rings with considerable oxygen content (~30%). We then test and demonstrate the usefulness of CANP as a 2D material for applications involving polymers and water treatment membrane. The device we used for the synthesis is illustrated in Fig. 1a. It is a custom-made reactor that is equivalent to an autoclave (refer to Supporting Information for fabrication of CANP). The distinguishing feature of the hydrothermal device is that it is devoid of liquid except for the glucose solution in the petri dish, unlike the typical practice of filling an autoclave with a

1 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1. Synthesis method and characteristics of synthesized carbonaceous nanoplates (CANP). (a) Schematic of the device for synthesizing CANP. (b) Optical images of as-synthesized solid mat of CANP. (c) Colloidal solution of CANP dispersed in deionized water and N-methyl-2-pyrrolidone (NMP). (d) Transmission electron microscopy (TEM) image of CANP on carbon gird. (e) Atomic force microscopy (AFM) image of CANP. (f) Height versus distance graph obtained along the white line in the AFM image of e, The length of CANP indicated by white line is 725 nm, the height is 1.3 nm, and the aspect ratio is 560. reaches a critical level, burst nucleation takes place.13-14 A stable nucleus forms when the number of adspecies exceeds a certain level. These stable nuclei can coalesce to grow in size, typically through Oswald ripening.15 The effects the viscosity has on the formation of stable nuclei and the nuclei growth can be examined from the viewpoint of the adspecies and nuclei movement. It is natural that the adspecies, for one, would find themselves more difficult to move in a more viscous solution. The same applies to stable nuclei, all the more so because of their larger size. The number of stable nuclei formed in the burst nucleation, therefore, would be smaller for higher viscosity solution of glucose because of the lower mobility of adspecies that lowers the chance for them to get together. The coalescence of stable

The geometric features of the CANP were investigated by transmission electron microscopy (TEM) and atomic force microscopy (AFM). The high resolution TEM image in Fig. 1d reveals the plate structure of CANP. It is thin enough that the carbon grid below the CANP can be identified. The AFM image in Fig. 1e shows that the length of the CANPs ranges from 200 nm to 700 nm. The AFM scanning result along the white line in Fig. 1e is given in Fig. 1f. The image shows that the thickness of CANP is relatively uniform at approximately 1.3 nm over the whole length of approximately 700 nm. The formation of nanoplates as verified by TEM and AFM indicates a two-dimensional growth of nuclei of converted glucose. Fragmented glucose derivatives are known to form in the hydrothermal conversion of glucose to carbon spheres.11-12 When the concentration of these derivatives, or adspecies,

Figure 2. Effect of glucose concentration on CANP dimensions. (a) Dependence of average aspect ratio on glucose concentration. (b-e) AFM images of CANP made from different glucose solutions of (b) 20, (c) 40, (d) 60, and (e) 80 wt%.

2 ACS Paragon Plus Environment

Page 2 of 7

Page 3 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 3. Characterization of CANP. (a) Distribution histograms of length, (b) height, and (c) aspect ratio of CANPs. Average values of length, height, and aspect ratio are 380 nm, 1.3 nm, and 310, respectively. (d) Fourier transform infrared spectra of glucose powder, CANP, and GO. (e) X-ray photoelectron spectroscopy spectrum of CANP. (f) 13C nuclear magnetic resonance spectrum of CANP. While a monosaccharide of glucose was chosen for its conversion to CANP material here, polysaccharides such as sucrose can also be used for a raw material for CANP synthesis (Fig. S5). For the new 2D material of CANP, its physical and chemical nature was probed by various instrumental methods. The size distributions of the CANP as determined by AFM with more than 100 nanoplates are given in Fig. 3a, b, and c, respectively, for the plate length, the height, and the aspect ratio. The average length of the CANP is 380 nm (Fig. 3a) and the average thickness is 1.3 nm (2 or 3 layers) (Fig. 3b), which agrees well with the TEM images, yielding an average aspect ratio of 310 (Fig. 3c). The CANP was synthesized under the typical synthesis conditions with 20 g of the 80 wt% glucose solution. The FT-IR results in Fig. 3d show that the spectra of CANP and GO have the same peaks. Furthermore, the peaks observed in the CANP spectrum at 1613 cm-1 (sp2 bonding, C=C bond) and 1705 cm-1 (carbonyl and carboxylic groups) cannot be found in the FT-IR spectrum of glucose powder, which implies that the carbon double bond and oxygen functional groups were generated by the hydrothermal process. The XPS spectra in Fig. 3e confirm that the CANP is decorated with abundant oxygen functional groups. The XPS spectra represent hydroxyl group (285.6 eV), epoxy group (286.7 eV), carbonyl group (288.2 eV), and carboxyl group (289.4 eV). The same oxygen functional groups were observed in the XPS spectrum of GO (Fig. S6). The difference between CANP and GO in functional groups is the atomic concentration of oxygen. CANP has a lower concentration of oxygen (30%) than GO (40%). The XRD spectrum in Fig. S7b indicates that the CANP is layered, and its structure is not crystalline. In contrast to the sharp band of GO, the CANP band is broad, implying that the layered structure is turbostratic rather than graphitic.17 The average interlayer spacing of CANP obtained from the XRD data is 0.573 nm, less than that of GO (0.835 nm), but larger than that of reduced graphene oxide (rGO) and graphite (0.362 nm and 0.334 nm, respectively).18

nuclei would be impeded by the high viscosity that makes the large nuclei practically immobile. The adspecies formed after the burst nucleation can either get together to create stable nuclei or migrate to be incorporated into the existing nuclei. The induction time16, which is needed for many bodies to configure themselves into a stable nucleus, would be longer for a more viscous solution. The adspecies would then favor the simpler and quicker process of migration and incorporation rather than nucleation when the solution is more viscous. These adspeices are more likely to be attached to the more energetically favorable edge sites rather than the sites on the basal plane, leading to a directional growth or two-dimensional growth. Therefore, the resulting 2D material would have a larger aspect ratio when the glucose solution is more viscous since then the adspecies are less mobile and thus less number of stable nuclei would form in the burst nucleation. The actual outcome of the nuclei growth as affected by viscosity is given in Figure 2. The glucose concentration of the solution in weight percent is equivalent to viscosity since the viscosity increases exponentially with increasing concentration (see Fig. S3). The highest glucose concentration attainable at room temperature was 80 wt% in water, for which the viscosity is approximately 1,800 centipoise (cP). In contrast, the concentration typically used in the carbon sphere synthesis is around 1 mole per liter (16 wt%). Figure 2a shows that the aspect ratio of plate length to thickness increases with increasing concentration or viscosity, indicating that a more viscous solution leads to a more two dimensional growth. More than 100 nanoplates were examined by AFM for each of the various concentrations of glucose solution, as shown in the figure (Fig. 2b-e and Fig. S4). Figure 2, taken as a whole, suggests that the two dimensional growth is induced by viscosity in the sense that the 2D growth does take place and the 2D material forms only when the viscosity is sufficiently high.

3 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The 13C NMR spectrum of CANP shown in Fig. 3f reveals that the CANP has C=C-O (151 ppm) and C=C-O bonds (111 ppm) besides carbonyl (C=O, 203 ppm), carboxyl (C(O)O, 175 ppm), and hydroxyl groups (C-OH, 70 ppm). The C=C-O and C=C-O bonds come from furanic group,19 which is a fivemembered aromatic ring with four carbon atoms and one oxygen, suggesting that the amorphous planar structure of CANP consists mainly of interconnected furan and benzene rings. Furanic groups have not been observed in NMR spectrum of GO and rGO.20 The Raman spectra in Fig. S8 show that both CANP and GO have the same D (1360 cm-1) and G (1580 cm-1) peaks.21 These characterization results indicate that the CANP has the same functional groups as GO does, but its basal plane is much more sparsely populated by the oxygen functional groups, which partly accounts for the smaller oxygen content of CANP (30%) than that of GO (40%). The CANP also has an aspect ratio (310) that is smaller than GO (1000). The oxygen content of CANP can be modulated between 25% and 38% by manipulating temperature and time, as provided in Supporting Information (Fig. S9). With the characterization results on hand, we set out to test the usefulness of the CANP material. The first was its utility as an additive for reinforcing the mechanical properties of polymer. For the purpose, polyurethane composites were prepared by a hot-press method to investigate the mechanical properties by tensile tests. The CANP contents in the polyurethane nanocomposite were 0.25, 0.5, 0.75, 1, and 2 wt%. The pristine polyurethane with no CANP was included for comparison. Figure. S10a is the optical image of the polyurethane nanocomposite film with 1 wt% CANP, showing uniform dispersion of CANP in the polyurethane

Page 4 of 7

matrix without any noticeable aggregation. Both ultimate tensile strength and fracture strain increased with the addition of CANP in the range of 0.25 to 1.0 wt% (Fig. S10b). Figures 4a and 4b show the tensile strength and the toughness obtained by the tests, respectively, for the polymer nanocomposites containing CANP, GO and rGO. The tensile strength of the nanocomposite was the highest at 0.75 wt% of CANP, which is 49% higher than that of pristine polyurethane. In the case of the toughness, it was the highest at 0.5 wt% of CANP, which is more than 600% greater than that of pristine polyurethane. High surface area and abundant oxygen functional groups of CANP enhanced the mechanical properties of polyurethane by promoting the interactions between CANP and polyurethane. CANP distributed in polyurethane prevents internal crack propagation and at the same time induces a lot of microcracks which could absorb fracture energy and relieve stress concentrations.22-24 The tensile strength, strain, and toughness all decrease for the CANP content larger than 1 wt% because of aggregation of CANP. The polyurethane nanocomposites with rGO and GO were also tested for comparison. The results in the figures show that the tensile strength of the nanocomposite rather decreases with increasing content of GO in polyurethane but the toughness increases slightly. An incompatibility of GO with the solvent used for polyurethane, which is NMP, could explain the nanocomposite performance.25-26 Aggregates of GO were visible in the test film. For better compatibility with the solvent, rGO was tested as an additive. As shown in the figures, both the tensile strength and the toughness increased

Figure 4. Mechanical and anti-biofouling properties of CANP composites. (a) Dependence of tensile strength of polyurethane nanocomposite on weight percent of various fillers. (b) Dependence of toughness of the nanocomposite on weight percent of various fillers. (c) Contact angle of PSf/30-CANP nanocomposite membrane for various CANP contents. (d) Flux decline with time in crossflow system down to approximately 25% of the initial flux for pristine PSf membrane, and composite membranes of PSf/30-CANP (CANP with oxygen content of 30%), PSf/38-CANP, and PSf/GO.

4 ACS Paragon Plus Environment

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

with increasing rGO content up to 1 wt%, indicating that the solvent compatibility is better for rGO. Given that the length of CANP is shorter than that of rGO, we believe that the superior properties of CANP composite are due to the strong interaction between CANP and polyurethane. Abundant oxygen functional groups in CANP have a role in inducing covalent bond formation between CANP and polyurethane matrix.27 Furthermore, the broad length and thickness distributions of CANP (Fig. 3a and b) would help create a hierarchical structure that significantly improves the mechanical properties.22, 28 In view of a prior study on poly(methylmethacrylate) (PMMA) nanocomposite29 for which the solvent was tetrahydrofuran (THF), a comparison was made between functionalized graphene (FG)-added PMMA and CANPadded PMMA (See Supporting Information S3 and Fig. S11). Both composites show better mechanical properties than pristine PMMA but functionalized graphene-PMMA is more effective than CANP-PMMA. To further test the usefulness of the CANP material, it was utilized as an additive in forming a composite membrane with polysulfone (PSf) for water treatment. The motivation stemmed from the facts that GO is known to mitigate membrane biofouling during water treatment30-32 and CANP has the same oxygen functional groups as GO. For the fabrication, polymer solutions of PSf in NMP were prepared by varying the CANP content in the solution. The contents were 0, 0.28, 0.56, 0.84, and 1.12 wt%, with respect to PSf weight. These contents were adopted on the basis of the previous studies that reported superior antifouling capability at approximately 1 wt% GO.30 Figure 4c shows that the contact angle decreases with increasing CANP content in the composite membrane, indicating that the surface becomes more hydrophilic. The hydrophilic surface with low surfacewater interfacial energy is known to resist protein adsorption and cell adhesion.33 The solid–liquid interfacial free energy, −ΔGSL, was found to be the largest at ~70 mJ m-2 for the PSf/CANP membrane with 1.12 wt% CANP. The value without CANP was ~63 mJ m-2. This 11% increase is comparable to those (7~17%) of the membranes with similar GO content.30 On the strength of the finding, ultrafiltration (UF) experiment was conducted using Pseudomonas aeruginosa as a foulant in a cross-flow system to test CANP’s ability to mitigate biofouling. Composite membranes of PSf/CANP and PSf/GO, both containing 1.12 wt% additive, and pristine PSf membrane were prepared for comparison. Two different PSf/CANP membranes of 30-CANP (30% oxygen content) and 38-CANP (38% oxygen content) were used to assess the effect of the oxygen content. Figure 4d shows the UF experimental results. Because of biofouling, the flux in all cases decline with time in the cross-flow system. The usual practice is to run the system until the decline reaches a certain level, backwash the system, and then run the system again. In the experiment, this level was taken as 25% of the initial flux. The membrane operation time before backwashing is only 5.5 hours for the pristine PSf membrane as shown in the figure. It is extended to 8.5 hours for 30-CANP composite membrane, 10 hours for 38-CANP membrane, and 12 hours for PSf/GO composite membrane. Although PSF/CANP membrane is not as effective as PSf/GO, it almost doubles the operation time before backwashing from 5.5 hrs to 10 hrs, demonstrating its utility as an additive for water treatment membrane. In summary, we have synthesized a carbonaceous 2D material with oxygen functional groups. The 2D material of

CANP is two to three monolayers thick, on the average, with an interlayer spacing of approximately 0.57 nm. The aspect ratio can be manipulated up to 310 by adjusting the glucose concentration. High viscosity of glucose solution has been the key to the two-dimensional growth and thus the synthesis of the 2D material. To provide a high viscosity environment, a vapor-filled hydrothermal condition has been conceived, and a reactor suitable for the synthesis devised. The clean synthesis, involving only water and glucose, leads to complete conversion of glucose to a damp solid mat of CANP, which can be used as such, requiring no separation, or simply dried if so desired. Application tests have demonstrated its usefulness for polymers and water treatment membrane. The availability of the inexpensive CANP material bodes well for practical applications.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI:

AUTHOR INFORMATION Corresponding Author *[email protected] *[email protected]

Author Contributions ‡These authors contributed equally. T.K., J.L., and G.L. contributed to experiment design, measurements, data analysis, and manuscript preparation. J.L. and H.S. contributed to experimental measurements and data analysis. J.Y.J., H.H.L., and Y.H.K. contributed to planning experiments, data analysis, and manuscript preparation.

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This research was supported by the National Research Foundation (NRF) of Korea (NRF-2019R1C1C1007926 and NRF-2016R1E1A1A01942110), Tetrels Technology Corporation, and the Program of Development of Space Core Technology through NRF funded by the Ministry of Science, ICT and Future Planning (NRF-2015M1A3A3A05027630). The authors also acknowledge support from the Institute of Advanced Aerospace Technology at Seoul National University.

REFERENCES (1) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.-J.; Loh, K. P.; Zhang, H. The Chemistry of Two-Dimensional Layered Transition Metal Dichalcogenide Nanosheets. Nat Chem 2013, 5, 263-275. (2) Butler, S. Z.; Hollen, S. M.; Cao, L.; Cui, Y.; Gupta, J. A.; Gutiérrez, H. R.; Heinz, T. F.; Hong, S. S.; Huang, J.; Ismach, A. F.; Johnston-Halperin, E.; Kuno, M.; Plashnitsa, V. V.; Robinson, R. D.; Ruoff, R. S.; Salahuddin, S.; Shan, J.; Shi, L.; Spencer, M. G.; Terrones, M.; Windl, W.; Goldberger, J. E. Progress, Challenges, and Opportunities in Two-Dimensional Materials Beyond Graphene. ACS Nano 2013, 7, 2898-2926. (3) Hernandez, Y.; Nicolosi, V.; Lotya, M.; Blighe, F. M.; Sun, Z.; De, S.; McGovern, I. T.; Holland, B.; Byrne, M.; Gun'Ko, Y. K.; Boland, J. J.; Niraj, P.; Duesberg, G.; Krishnamurthy, S.; Goodhue, R.; Hutchison, J.; Scardaci, V.; Ferrari, A. C.; Coleman, J. N. High-Yield Production of Graphene by

5 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Liquid-Phase Exfoliation of Graphite. Nat Nano 2008, 3, 563568. (4) Hu, B.; Wang, K.; Wu, L. H.; Yu, S. H.; Antonietti, M.; Titirici, M. M. Engineering Carbon Materials from the Hydrothermal Carbonization Process of Biomass. Advanced Materials 2010, 22, 813-828. (5) Liu, X. F.; Fechler, N.; Antonietti, M.; Willinger, M. G.; Schlogl, R. Synthesis of Novel 2-D Carbon Materials: Sp(2) Carbon Nanoribbon Packing to Form Well-Defined Nanosheets. Materials Horizons 2016, 3, 214-219. (6) Liu, X. F.; Giordano, C.; Antonietti, M. A Facile MoltenSalt Route to Graphene Synthesis. Small 2014, 10, 193-200. (7) Li, X. H.; Antonietti, M. Polycondensation of Boron- and Nitrogen-Codoped Holey Graphene Monoliths from Molecules: Carbocatalysts for Selective Oxidation. Angew. Chem.-Int. Edit. 2013, 52, 4572-4576. (8) Li, X.-H.; Kurasch, S.; Kaiser, U.; Antonietti, M. Synthesis of Monolayer-Patched Graphene from Glucose. Angewandte Chemie International Edition 2012, 51, 96899692. (9) Dreyer, D. R.; Park, S.; Bielawski, C. W.; Ruoff, R. S. The Chemistry of Graphene Oxide. Chemical Society Reviews 2010, 39, 228-240. (10) Novoselov, K. S.; Falko, V. I.; Colombo, L.; Gellert, P. R.; Schwab, M. G.; Kim, K. A Roadmap for Graphene. Nature 2012, 490, 192-200. (11) Sevilla, M.; Fuertes, A. B. Chemical and Structural Properties of Carbonaceous Products Obtained by Hydrothermal Carbonization of Saccharides. Chemistry – A European Journal 2009, 15, 4195-4203. (12) Li, M.; Li, W.; Liu, S. Control of the Morphology and Chemical Properties of Carbon Spheres Prepared from Glucose by a Hydrothermal Method. Journal of Materials Research 2012, 27, 1117-1123. (13) LaMer, V. K.; Dinegar, R. H. Theory, Production and Mechanism of Formation of Monodispersed Hydrosols. Journal of the American Chemical Society 1950, 72, 48474854. (14) Thanh, N. T. K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of Nanoparticles in Solution. Chemical Reviews 2014, 114, 7610-7630. (15) Lifshitz, I. M.; Slyozov, V. V. The Kinetics of Precipitation from Supersaturated Solid Solutions. Journal of Physics and Chemistry of Solids 1961, 19, 35-50. (16) Kashchiev, D.; Verdoes, D.; van Rosmalen, G. M. Induction Time and Metastability Limit in New Phase Formation. Journal of Crystal Growth 1991, 110, 373-380. (17) Kim, P.; Johnson, A.; Edmunds, C. W.; Radosevich, M.; Vogt, F.; Rials, T. G.; Labbe, N. Surface Functionality and Carbon Structures in Lignocellulosic-Derived Biochars Produced by Fast Pyrolysis. Energy & Fuels 2011, 25, 46934703. (18) Moon, I. K.; Lee, J.; Ruoff, R. S.; Lee, H. Reduced Graphene Oxide by Chemical Graphitization. Nat. Commun. 2010, 1, 73. (19) Falco, C.; Perez Caballero, F.; Babonneau, F.; Gervais, C.; Laurent, G.; Titirici, M.-M.; Baccile, N. Hydrothermal Carbon from Biomass: Structural Differences between Hydrothermal and Pyrolyzed Carbons Via 13c Solid State Nmr. Langmuir 2011, 27, 14460-14471. (20) Park, S.; Hu, Y.; Hwang, J. O.; Lee, E.-S.; Casabianca, L. B.; Cai, W.; Potts, J. R.; Ha, H.-W.; Chen, S.; Oh, J.; Kim, S. O.; Kim, Y.-H.; Ishii, Y.; Ruoff, R. S. Chemical Structures of Hydrazine-Treated Graphene Oxide and Generation of Aromatic Nitrogen Doping. Nat. Commun. 2012, 3, 638.

Page 6 of 7

(21) Thomsen, C.; Reich, S. Double Resonant Raman Scattering in Graphite. Physical Review Letters 2000, 85, 5214-5217. (22) Zhao, Z. B.; Teng, K. Y.; Li, N.; Li, X. J.; Xu, Z. W.; Chen, L.; Niu, J. R.; Fu, H. J.; Zhao, L. H.; Liu, Y. Mechanical, Thermal and Interfacial Performances of Carbon Fiber Reinforced Composites Flavored by Carbon Nanotube in Matrix/Interface. Composite Structures 2017, 159, 761-772. (23) Sui, X. H.; Shi, J.; Yao, H. W.; Xu, Z. W.; Chen, L.; Li, X. J.; Ma, M. J.; Kuang, L. Y.; Fu, H. J.; Deng, H. Interfacial and Fatigue-Resistant Synergetic Enhancement of Carbon Fiber/Epoxy Hierarchical Composites Via an Electrophoresis Deposited Carbon Nanotube-Toughened Transition Layer. Composites Part a-Applied Science and Manufacturing 2017, 92, 134-144. (24) Teng, K. Y.; Ni, Y.; Wang, W.; Wang, H. B.; Xu, Z. W.; Chen, L.; Kuang, L. Y.; Ma, M. J.; Fu, H. J.; Li, J. Adjustable Micro -Structure, Higher-Level Mechanical Behavior and Conductivities of Preformed Graphene Architecture/Epoxy Composites Via Rtm Route. Composites Part a-Applied Science and Manufacturing 2017, 94, 178-188. (25) Li, W. X.; Xu, Z. W.; Chen, L.; Shan, M. J.; Tian, X.; Yang, C. Y.; Lv, H. M.; Qian, X. M. A Facile Method to Produce Graphene Oxide-G-Poly(L-Lactic Acid) as an Promising Reinforcement for Plla Nanocomposites. Chem. Eng. J. 2014, 237, 291-299. (26) Zhao, X. M.; Li, N.; Jing, M. L.; Zhang, Y. F.; Wang, W.; Liu, L. S.; Xu, Z. W.; Liu, L. Y.; Li, F. Y.; Wu, N. Monodispersed and Spherical Silver Nanoparticles/Graphene Nanocomposites from Gamma-Ray Assisted in-Situ Synthesis for Nitrite Electrochemical Sensing. Electrochimica Acta 2019, 295, 434-443. (27) Pokharel, P.; Pant, B.; Pokhrel, K.; Pant, H. R.; Lim, J. G.; Lee, D. S.; Kim, H. Y.; Choi, S. Effects of Functional Groups on the Graphene Sheet for Improving the Thermomechanical Properties of Polyurethane Nanocomposites. Compos. Pt. B-Eng. 2015, 78, 192-201. (28) Qian, H.; Greenhalgh, E. S.; Shaffer, M. S. P.; Bismarck, A. Carbon Nanotube-Based Hierarchical Composites: A Review. Journal of Materials Chemistry 2010, 20, 4751-4762. (29) RamanathanT; Abdala, A. A.; StankovichS; Dikin, D. A.; Herrera Alonso, M.; Piner, R. D.; Adamson, D. H.; Schniepp, H. C.; ChenX; Ruoff, R. S.; Nguyen, S. T.; Aksay, I. A.; Prud'Homme, R. K.; Brinson, L. C. Functionalized Graphene Sheets for Polymer Nanocomposites. Nat. Nanotechnol. 2008, 3, 327-331. (30) Lee, J.; Chae, H.-R.; Won, Y. J.; Lee, K.; Lee, C.-H.; Lee, H. H.; Kim, I.-C.; Lee, J.-m. Graphene Oxide Nanoplatelets Composite Membrane with Hydrophilic and Antifouling Properties for Wastewater Treatment. Journal of Membrane Science 2013, 448, 223-230. (31) Xu, Z. W.; Wu, T. F.; Shi, J.; Teng, K. Y.; Wang, W.; Ma, M. J.; Li, J.; Qian, X. M.; Li, C. Y.; Fan, J. T. Photocatalytic Antifouling Pvdf Ultrafiltration Membranes Based on Synergy of Graphene Oxide and Tio2 for Water Treatment. Journal of Membrane Science 2016, 520, 281-293. (32) Zhang, J. G.; Xu, Z. W.; Shan, M. J.; Zhou, B. M.; Li, Y. L.; Li, B. D.; Niu, J. R.; Qian, X. M. Synergetic Effects of Oxidized Carbon Nanotubes and Graphene Oxide on Fouling Control and Anti-Fouling Mechanism of Polyvinylidene Fluoride Ultrafiltration Membranes (Vol 448, Pg 81, 2013). Journal of Membrane Science 2014, 451, 319-319. (33) Krishnan, S.; Weinman, C. J.; Ober, C. K. Advances in Polymers for Anti-Biofouling Surfaces. Journal of Materials Chemistry 2008, 18, 3405-3413.

6 ACS Paragon Plus Environment

Page 7 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Table of Contents

7 ACS Paragon Plus Environment