Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Biomedical Applications of Graphene and 2D Nanomaterials
Biomedical Applications of Graphene and 2D Nanomaterials
Biomedical Applications of Graphene and 2D Nanomaterials
Ebook866 pages9 hours

Biomedical Applications of Graphene and 2D Nanomaterials

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Biomedical Applications of Graphene and 2D Nanomaterials provides a much-needed reference on the biomedical applications of 2D nanomaterials, as well as theoretical knowledge on their structure, physicochemical properties and biomedical applications. Chapters are dedicated to growth areas, such as size and shape-dependent chemical and physical properties and applications, such as in diagnostic and therapeutic products. The book also discusses the concept, development and preclinical studies of 2D nanomaterials-based biomedical tools, such as biosensors, artificial organs and photomedicine. Case studies and reports form the core of the book, making it an ideal resource on potential applications in biomedical science and engineering.

This timely resource for scientists and engineers in this rapidly advancing field features contributions from over 30 leaders who address advanced methods and strategies for controlling the physical-chemical properties of 2D nanomaterials, along with expert opinions on a range of 2D nanomaterials that have therapeutic and diagnostic applications.

  • Presents advanced methods and strategies for controlling the physical-chemical properties of 2D nanomaterials
  • Provides state-of-the-art biomedical applications for 2D nanomaterials, including graphene and boron nitride
  • Includes key information from a broad selection of subject areas for researchers in both materials, engineering and medicine
LanguageEnglish
Release dateMar 31, 2019
ISBN9780128162699
Biomedical Applications of Graphene and 2D Nanomaterials

Related to Biomedical Applications of Graphene and 2D Nanomaterials

Titles in the series (97)

View More

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Biomedical Applications of Graphene and 2D Nanomaterials

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Biomedical Applications of Graphene and 2D Nanomaterials - Md Nurunnabi

    Australia

    Chapter 1

    Two-Dimensional Nanomaterials: Crystal Structure and Synthesis

    Khaled Parvez    School of Chemistry, University of Manchester, Manchester, United Kingdom

    Abstract

    Since the discovery of graphene in 2004, research on two-dimensional (2-D) nanomaterials has grown exponentially in the fields of material science, condense matter physics, chemistry, and nanotechnology. The 2-D nanomaterials such as graphene, transition metal dichalcogenides, silicate clays, and hexagonal boron nitride provide enhanced physical, chemical, and biological functionality owing to their uniform shapes, high surface-to-volume ratios, and surface charge. However, research on 2-D nanomaterials is still its infancy, with majority of research focusing on elucidating unique material characteristics. To characterize the layer-dependent changes in properties and to provide pathways for their integration into a multitude of applications, it is essential to explore the reliable synthesis of single- and few-layer 2-D nanomaterials. Therefore, many synthetic strategies such as micromechanical exfoliation, liquid-phase exfoliation, and chemical vapor deposition have been developed to synthesize high-quality and ultrathin nanosheets showing their own merits and demerits in preparing 2-D nanomaterials. In this chapter, we summarize the state-of-the-art progress of this dynamically developed material family with a particular focus on their crystal structure and synthetic methods. Eventually, the potential trends and future direction for synthesizing technology for 2-D nanomaterials are proposed.

    Keywords

    2-D nanomaterials; Graphene

    Chapter Outline

    1Introduction

    2Crystal Structures of 2-D Materials

    2.1Graphene

    2.2Hexagonal Boron Nitride

    2.3.Transition Metal Dichalcogenides

    2.4Black Phosphorus

    2.5Graphitic Carbon Nitride

    2.6MXenes

    2.7Silicene and Germanene

    3.Synthetic Methods

    3.1Micromechanical Cleavage

    3.2Liquid-Phase Exfoliation

    3.3Shear Exfoliation

    3.4Electrochemical Exfoliation

    3.5Oxidation-Assisted Exfoliation

    3.6Hydro/Solvothermal Synthesis

    3.7Chemical Vapor Deposition

    4.Conclusion

    References

    1 Introduction

    Two-dimensional (2-D) materials have become a central topic of research interests since the exfoliation of graphene in 2004. The 2-D feature is unique and indispensable to access unprecedented physical, electronic, and chemical properties due to electron confinement in two dimensions. Graphene, a one-atom-thick and crystalline carbon film, is an exemplary model due to its unexpected properties including high carrier mobility; quantum hall effect; high specific surface area; and excellent optical, electronic, and thermal properties. Because of its remarkable properties, applications using graphene in a wide range of areas including high-speed electronics (1) and optical devices (2), energy generation and storage (3–5), hybrid materials (6), chemical sensors (7), and even DNA sequencing (8,9) have all been explored. The prerequisite for such applications is the mass production of graphene in a controlled manner because the numbers of graphene layers and the defects in these graphene layers significantly influence the subsequent properties. Methods such as mechanical exfoliation (10), liquid-phase exfoliation (11,12), electrochemical exfoliation (13–15), oxidation-assisted exfoliation (16–18), chemical vapor deposition (CVD) (19,20), and shear exfoliation (21,22) have been developed in order to make suitable graphene layers. Despite these efforts, the fine control of the number of layers and structure of graphene sheets over an entire substrate remains a major challenge.

    Since the discovery of the exotic properties of graphene, the explorations of other graphene-analogous 2-D nanomaterials are also growing. To name a few, transition metal dichalcogenides (TMDCs), layered metal oxides, hexagonal boron nitride (h-BN), graphitic carbon nitride (g-C3N4), layered double hydroxides (LDHs), MXenes, and black phosphorus (BP) are typical graphene-like 2-D nanomaterials that exhibit versatile properties due to their similar structure features but different compositions from graphene. The common feature of these layered materials is that the bulk 3-D crystals are stacked structures. They involve van der Waals interactions between adjacent sheets with strong covalent bonding within each sheet. Such materials span the entire range of electronic structures, from insulator, to semiconductor, to metal, and display interesting properties. Because of their distinct properties and high specific surface areas, these 2-D materials are important in various applications such as optoelectronics, spintronics, catalysts, chemical and biological sensors, supercapacitors, solar cells, and lithium-ion batteries.

    2 Crystal Structures of 2-D Materials

    Until now, large quantities of 2-D nanomaterials have been prepared by various synthetic methods. Even though the composition and crystal structures vary in different materials, they all can be categorized into two types: layered and nonlayer-structured materials. In layered materials, the in-plane atoms connect to each other by strong chemical bonding in each layer, while these layers stack together through the van der Waals interaction. Graphite is a typical example of the layered materials in which graphene layers are stacked together forming the bulk graphite. Besides graphite, there are many layered materials, such as h-BN, TMDCs, g-C3N4, BP, and transition metal oxides (TMOs). In contrast, other materials crystallize in three dimensions via atomic or chemical bonding, forming bulk crystals, such as metal oxides and metal chalcogenides. In this section, we will introduce the crystal structures of these widely explored 2-D nanomaterials based on the composition.

    2.1 Graphene

    Graphene is single-atom-thick graphite, an allotrope of carbon in the form of 2-D structure. It is composed of a hexagonal close-packed carbon network, in which each atom covalently bonds to three neighboring ones through the σ-bond. The distance between two carbon atoms is about 1.42 Å. Individual layers stack together through the van der Waals force to form the graphite in which the distance between the adjacent layer is about 3.35 Å (Fig. 1) (23).

    Fig. 1 Schematic illustration of (A) graphene and (B) hexagonal boron nitride ( h -BN). Reproduced with permission from Ba, K.; Jiang, W.; Cheng, J. X.; Bao, J. X.; Xuan, N. N.; Sun, Y. Y.; Liu, B.; Xie, A. Z.; Wu, S. W.; Sun, Z. Z. Sci. Rep. 2017, 7, 45584, Copyright 2017 Macmillan Publishers Limited.

    2.2 Hexagonal Boron Nitride

    The bulk h-BN has a layered structure similar to that of graphite and therefore is also known as white graphite. It consists of equal number of boron and nitrogen atoms arranged in a hexagonal structure (24). Within each layer, the boron and nitrogen atoms are covalently bonded, and the layers stack together by the van der Waals force to form the bulk crystal. Compared with graphite, the bulk h-BN exhibits similar lattice constant and interlayer distances (3.30–3.33 Å) (Fig. 1B). Similar to graphite, bulk h-BN can also be exfoliated into single or few atomic layers. Furthermore, carbon, nitrogen, and boron can all be intermixed to create doped or alloyed layers with tunable electronic properties (i.e., p- and n-type doing and bandgap engineering) (25). Exfoliated thin-layer h-BN has been widely explored as ultraflat insulating substrates (26), encapsulation materials (27), and dielectrics (28).

    2.3 Transition Metal Dichalcogenides

    TMDCs consists of hexagonal layers of metal atoms (M) sandwiched between two layers of chalcogen atoms (X) with a MX2 stoichiometry. Depending on the combination of chalcogen (e.g., S, Se, or Te) and transition metal (typically Mo, W, Nb, Re, Ni, or V), TMDCs occur in more than 40 different categories. In layered structures, each layer typically has a thickness of 6–7 Å, which consists of a hexagonally packed layer of metal atoms sandwiched between two layers of chalcogen atoms (X units in the unit cell, that is, the number of layers in the stacking sequence (Fig. 2D). Natural MoS2 is commonly found in the 2H phase. Synthetic MoS2 often contains the 3R phase. In both cases, the metal coordination is trigonal prismatic. The most frequently studied TMDCs (e.g., MoS2, MoSe2, WS2, and WSe2) typically have a tunable bandgap that transition from indirect in the bulk to direct in the monolayers (30). Unusual properties are also observed such as lattice symmetry-induced valley Hall effect (31), valley polarization (32), and superconductivity (33). These TMDCs typically have bandgaps in the range of 1–2 eV, with the measured carrier mobility generally in the order of 100 cm² V− 1 s− 1 at room temperature (34). A class of closely related layered chalcogenide materials are Bi2Se3 and Bi2Te3 that attracted considerable interests as topological insulators due to their unique electronic properties and are potential for quantum electronic devices and thermoelectric applications. Recent studies demonstrate a substantially reduced thermal conductivity and enhanced power factor with reducing sheet thickness (35,36).

    Fig. 2 (A) Three-dimensional representation of the structure of transition metal dichalcogenides (TMDCs). (B) Archetypal coordination in TMDCs. Octahedral coordination (left) often forms 1T structures, whereas triangular prismatic coordination (right) forms 2H or 3R structures. (C) Top view of monolayers constructed from octahedral (left) and triangular prismatic (right) coordination. (D) Side views of 1T, 2H, and 3R polytype structures. (A) Reproduced with permission from Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Nat. Nanotechnol. 2011, 6, 147, Copyright 2011 Nature Publishing Group; (D) Reproduced with permission from Zhang, Y. J.; Yoshida, M.; Suzuki, R.; Iwasa, Y. 2D Mater 2015, 2, 0044004, Copyright 2015 IOP publishing.

    2.4 Black Phosphorus

    BP is another 2-D material that has received increasing attention. The bulk BP crystallizes into a layered orthorhombic crystal structure. The distance between adjacent layers is 5.3 Å (Fig. 3A), and individual layers stack together through the van der Waals force (37). A BP monolayer, also known as phosphorene, is composed of an armchair vertically staggered hexagonal lattice structure, in which one P atom bonds with other three. Among the four P atoms, three of them are located in the same plane with a bond length of 0.2224 nm, while the fourth one is located at the parallel adjacent plane (bond length 0.2244 nm) (38). The top view of BP along the z direction shows a hexagonal structure with bond angles of 96.3 degree and 102.1 degree (Fig. 3B) (37). Depending on the number of layers, BP displays tunable bandgap from 0.33 eV for bulk to 1.5 eV for monolayer with a carrier mobility up to ~ 1000 cm² V− 1 s− 1 (39,40). With an intrinsic direct bandgap and decent carrier mobility, BP may bridge the gap between graphene and TMDCs.

    Fig. 3 (A) Side view of the BP crystal lattice. The interlayer spacing is 5.3 Å. (B) Top view of the lattice of single-layer BP. The bond angles are shown. The corresponding x, y, and z directions are indicated in both (A) and (B). x and y correspond to the armchair and zigzag directions of BP, respectively. (C) and (D) Scheme of s -triazine and tri- s -triazine-based connection in graphitic carbon nitride ( g -C 3 N 4 ), respectively. Blue and gray spheres represent nitrogen and carbon atoms, respectively. (A) and (B) Reproduced with permission from Ling, X.; Wang, H.; Huang, S. X.; Xia, F. N.; Dresselhaus, M. S. Proc. Natl. Acad. Sci. U.S.A. 2015, 112, 4523, Copyright 2015 National Academy of Sciences; (C) and (D) Reproduced with permission from Zheng, Y.; Liu, J.; Liang, J.; Jaroniec, M.; Qiao, S. Z. Energ. Environ. Sci. 2012, 5, 6717, Copyright 2012, the Royal Society of Chemistry.

    2.5 Graphitic Carbon Nitride

    The g-C3N4 is another graphite analogue with a van der Waals layered structure. The g-C3N4 crystal structure can be regarded as N-substitute graphite framework consisting of π-conjugated graphitic planes formed via sp² hybridization of carbon and nitrogen atoms. The stacking distance of two layers in g-C3N4 is d = 3.26 Å and is 3% more dense than the packing in crystalline graphite (d = 3.35 Å). This smaller interlayer distance can be explained by altering the localization of electrons and strengthening the binding between layers due to nitrogen atom substitution. There are two different structural models for the g-C3N4: One is s-triazine constructed by the condensed s-triazine units with a periodic array of single carbon vacancies, and the second one is condensed tri-s-triazine units connected through the planar tertiary amino groups with larger periodic vacancies in the lattice (Fig. 3C and D) (41). The density function theory (DFT) calculations further indicated that the g-C3N4 networks composed of the melon-based segments are thermodynamically more stable than the melamine-based arrangements because the tri-s-triazine unit is energetically more stable than s-triazine.

    2.6 MXenes

    MXenes are a class of 2-D layered transition metal carbides and/or nitride produced by selective etching of the raw MAX phase with a general formula of Mn + 1AXn (n = 1, 2, or 3), where M is the transition metal (e.g., Ti, V, Nb, and Cr) (Fig. 4A), A is an element from group IIIA or IVA (i.e., Al, Si, Sn, and In), and X is carbon and/or nitrogen (42). MAX phases have a layered hexagonal structure with P63/mmc symmetry in which M layers are nearly hexagonally close packed together and X atoms fill the octahedral sites (43). The element A is metallically bonded to the M element and interleaved in Mn + 1Xn layers. The A layers can be selectively etched from the MAX phase using strong etching solutions, for example, HF, forming MXenes with three different structures, that is, M2X, M3X2, or M4X3 (43).

    Fig. 4 (A) Crystal structure of MAX phases M 2 AX, M 3 AX 2 , and M 4 AX 3 (i.e., with n  = 1, 2, and 3). (B) and (C) Crystal structures of silicene and germanene, where the atoms belonging to the inequivalent sublattices are separated by distances (buckling) of 0.44 and 0.65 Å for silicine and germanene, respectively. (A) Reproduced with permission from Dhakal, C.; Aryal, S.; Sakidja, R.; Ching, W. Y. J. Eur. Ceram. Soc. 2015, 35, 3203, Copyright 2015 Elsevier; (B) and (C) Reproduced with permission from Dimoulas, A. Microelectron. Eng. 2015, 131, 68, Copyright 2014 Elsevier.

    2.7 Silicene and Germanene

    Silicene and germanene are considered to be single layers of sp² hybridized silicon and germanium forming 2-D honeycomb lattice. The structure of germanene was first theoretically proposed together with that of silicone Ge bond distance of 2.25 and 2.38 Å for silicone and germanene, respectively, smaller than the corresponding distances in the bulk materials (Fig. 4B and C) (46). The buckling is a result of their tendency to make sp³ hybridization and gives them stability since it brings the pz orbitals closer, thus facilitating pz overlapping and π-bond formation. Although silicone does not exists in freestanding form, it is of immense scientific interest to find whether silicone can be engineered on appropriate substrates that could stabilize a 2-D honeycomb lattice, in order to obtain a new Si allotrope with properties similar to that of graphene.

    3 Synthetic Methods

    The capabilities for the preparation of thin 2-D nanomaterials with desired composition, size, thickness, crystal phase, and surface property are of particular importance for the further study of their physical, chemical, and electronic properties and variety of potential applications. The methods for preparing 2-D nanomaterials include micromechanical cleavage, liquid-phase exfoliation, shear exfoliation, CVD, and wet-chemical synthesis. In this section, we focus on the various synthetic methods to prepare 2-D nanomaterials.

    3.1 Micromechanical Cleavage

    Micromechanical cleavage is a traditional but efficient method to obtain atomically thin 2-D nanomaterials (47). In 2004, Novoselov and Geim for the first time successfully cleaved a single-layer graphene from small crystals of highly oriented pyrolytic graphite (HOPG) by the Scotch tape-based micromechanical cleavage (10). Fig. 5 illustrates the micromechanical exfoliation process for graphene. The exfoliation mechanics of this method are that the Scotch tape is applied to the HOPG surface and thus exerts a normal force. By repeating this normal force numerous times, the graphitic layer becomes thinner and thinner, and finally, it will become a single-layer graphene (47). Later, the same group demonstrated the extension of this technique for the exfoliation of ultrathin 2-D nanomaterials including h-BN, MoS2, and NbSe2 from their parent layered bulk crystals (48). Since then, this method has been widely used to cleave various kinds of 2-D nanomaterials ranging from TMDCs (e.g., TiS2, TaS2, MoSe2, WS2, WSe2, and MoTe2) (49–51), to topological insulator (e.g., Bi2Te3, Bi2Se3, and Sb2Te3) (52,53), to h-BN (26,48,54). Nevertheless, this approach can be categorized as a nondestructive technique because no chemicals are required during the fabrication process. Therefore, the exfoliated single- or few-layer nanosheets kept the high crystal quality, usually defined as pristine, from their bulk crystals. The size of the produced 2-D crystals can be up to a few to tens of micrometers, and the surfaces are very clean because no chemicals are introduced during the exfoliation process. The excellent crystal quality with minimum defects makes the mechanically cleaved thin 2-D nanomaterials compelling candidates for the fundamental study of the intrinsic physical, optical, and electronic properties as well as high-performance electronic and/or optoelectronic devices such as transistors and phototransistors.

    Fig. 5 An illustrative procedure of Scotch tape-based micromechanical cleavage of graphene. Reproduced with permission from Yi, M.; Shen, Z. G. J. Mater. Chem. A 2015, 3, 11700, Copyright 2015 the Royal Society of Chemistry.

    Although micromechanical cleavage has many advantages, such as high crystal quality, clean surface, and wide applicability, there are several disadvantages that restrict its practical application. The production yield is quite low, and thick flakes always coexist on the substrate along with the single- or few-layer nanosheets. Also the production rate is very slow and is not competitive with CVD growth and solution-based methods making it difficult to realize the demands for various practical applications. Moreover, the size, thickness, and shape of the exfoliated 2-D nanomaterials are difficult to control because the exfoliation process is operated manually by hands and, therefore, lacks precision, controllability, or reproducibility. Finally, a substrate is a prerequisite to support the produced 2-D sheets in the exfoliation process, thus eliminating the possibility of producing freestanding or solution-dispersed nanosheets.

    3.2 Liquid-Phase Exfoliation

    Two-dimensional materials can be exfoliated in liquid environments exploiting ultrasounds to extract individual layers. The liquid-phase exfoliation generally involves three steps: (i) dispersion in a solvent, (ii) exfoliation, and (iii) purification. The third step is necessary to separate exfoliated from unexfoliated flakes and usually requires ultracentrifugation. Exfoliation can be done via chemical wet dispersion followed by ultrasonication both in aqueous and nonaqueous solvents. The basic idea is that sonication can induce liquid cavitation, which in turn induces bubbles in the solution. Microjets and shockwaves passed through the layered bulk crystals dispersed in solution when these bubbles collapse. In this case, an intensive tensile stress will be generated on the layered bulk crystals, thus leading to the exfoliation of layered bulk crystals into thin layers of sheets. The key factor for achieving efficient exfoliation of layered bulk crystals is matching the surface energy between the layered bulk crystal and the solvent system. Graphene was first exfoliated in liquid phase by Coleman’s group in 2008 by ultrasonicating the graphite in N-methylpyrrolidone (NMP) (11). The obtained graphene sheets are considered to be pristine because no chemical functionalization is involved. After centrifugation to remove unexfoliated bulk crystals, the remaining dispersion consists of 28% single-layer graphene. Unfortunately, the yield of single-layer graphene was very low, approximately 1 wt% with a concentration of 0.01 mg/mL that is relatively low for further application. Further optimization of solvents led to some improvements in the dispersability of graphene. For instance, concentrations as high as 0.03 mg/mL, 0.1 mg/mL, and 0.5 mg/mL using ortho-dichlorobenzene (ODCB) (55), perfluorinated aromatic solvents such as pentafluorobenzonitrile (56) and benzylamine (57). However, due to the long ultrasonication process, the lateral dimension of the LPE graphene is relatively small (< 3 μm). Despite the significant progress achieved in the exfoliation of graphite, the aforementioned solvents have a high-boiling point and are relatively toxic. Therefore, efforts have been devoted to the exploration of low-boiling-point solvent for the sonication-assisted liquid exfoliation. For example, Coleman and coworkers have demonstrated the sonication of graphite in some low-boiling solvents, including isopropanol and chloroform to obtain a relatively high concentration of graphene suspensions (58). Some other low-boiling solvents such as propanol and acetonitrile were also used for LPE of graphite (59,60).

    Besides graphene, Coleman’s group further extended LPE method for the exfoliation of other layered bulk crystals into 2-D nanosheets including MoS2, MoSe2, WS2, h-BN, NbSe2, TaSe2, NiTe2, MoTe2, and Bi2Te3 (12). Both the experimental and theoretical results suggested that the good matching of surface tension between the layered crystals and the solvents is a key factor for the efficient exfoliation (12). The solvent is also important in stabilizing the exfoliated nanosheets and prohibiting their restacking and/or aggregation. Zhou et al. also demonstrated that the mixer of water and ethanol is effective for exfoliating and dispersing TMDC nanosheets (61).

    Water is a natural solvent choice because of its nontoxicity, which can provide possibilities for manipulation, in particular thin film fabrication. However, pure water is inefficient to exfoliate 2-D crystals due to large surface energy. However, a recent study has demonstrated that pure water can be a promising solvent in the sonication-assisted exfoliation method by simply heating the water at elevated temperature (e.g., 60°C) (62). At this point, water becomes both effective for the exfoliation of layered bulk crystals of graphite, h-BN, MoS2, WS2, and MoSe2 into 2-D nanosheets in the sonication process and the stabilization of the exfoliated nanosheets after exfoliation because of the presence of the platelet surface charges induced by edge functionalization or intrinsic polarity. The successful exfoliation of layered bulk crystals in pure water makes this method promising for the preparation of thin 2-D nanomaterials for practical applications.

    In the LPE method, the effective exfoliation can only be achieved in solvents with matching surface energy to the layered bulk crystals, thereby making it difficult to find a suitable solvent for each system for each layered bulk crystal. Alternatively, sonication of layered bulk crystals in aqueous solution with additional polymers or surfactants is another promising way for exfoliating them into thin 2-D nanosheets (Fig. 6) (63). The surface tension of the aqueous solution can be easily tuned by addition of polymers or surfactants, thus matching it to the surface energy of the layered bulk crystals and achieving efficient exfoliation of layered materials (64). In a rigorous investigation on surfactant-assisted graphite exfoliation in water, a variety of ionic and nonionic surfactants were explored. For example, anionic surfactants such as 4-dodecylbenzenesulfonic acid (SDBS) and sodium cholate (SC) have been reported to successfully exfoliate graphene from graphite with a concentration of 0.1 mg/mL and 0.3 mg/mL in water, respectively (65,66). Some other ionic surfactants, such as sodium deoxycholate (SDC), cetyltrimethylammonium bromide (CTAB), and acetic acid, have also been used to assist the exfoliation of graphite by sonication (67,68). Several nonionic surfactants for exfoliating 2-D nanosheets in aqueous solutions have been employed such as Pluronic P-123, polyoxyethylene sorbitan monooleate (Tween 80), polyoxyethylene sorbitan trioleate (Tween 85), polyoxyethylene(4)dodecyl ether (Brij 30), and polyoxyethylene octyl phenyl ether (Triton X-100) (69). In addition to these, several polymers such as polystyrene (PS), polyvinyl chloride (PVC), polymethyl methacrylate (PMMA), poly[styrene-b-(2-vinylpyridine)] (PS-b-P2VP), and poly(isoprene-b-acrylic acid) (PI-b-PAA) have also been investigated for sonication-assisted exfoliation of bulk crystals (69–71).

    Fig. 6 Schematic illustration of sonication-assisted liquid exfoliation of graphite into graphene. Reproduced with permission from Ciesielski, A.; Samori, P. Chem. Soc. Rev. 2014, 43, 381, Copyright 2014 the Royal Society of Chemistry.

    So far, the sonication-assisted liquid exfoliation method has been widely utilized for exfoliation of numerous layered compounds into thin 2-D nanomaterials, such as graphene (63), TMDCs (e.g., MoS2, WS2, MoSe2, WSe2, MoTe2, and TaSe2) (12,72–74), h-BN (12,72), topological insulators (e.g., Bi2Te3, Bi2Se3, and Sb2Te3) (75,76), black phosphorous (BP) (77,78), and g-C3N4 (79). In addition to thickness, the concentration and lateral size of the obtained nanosheets can be roughly tuned by controlling the solvent system, sonication time, ultrasonic power, temperature, surfactants, and/or additives. Currently, the sonication-assisted liquid-phase exfoliation might be the most widely used approach for the preparation of solution-dispersed 2-D nanomaterials. However, there are several disadvantages for the sonication-assisted LPE method. Among them, the yield of the single-layer nanosheets is low. As known, some of the extraordinary properties of 2-D nanomaterials only can be observed in its single-layer form. Also, the lateral size of the produced nanosheets is relatively small because the sonication force breaks down the big nanosheets into small fragments. Moreover, for the sonication in aqueous polymer and/or surfactant solution, the residual surfactants adsorbed on the exfoliated nanosheets are undesirable for their applications in electronics, optoelectronics, and energy storage devices. The sonication process also may induce some defects on the exfoliated 2-D nanosheets, which will affect their properties.

    3.3 Shear Exfoliation

    Another recently emerging method for scaling up the production of 2-D nanomaterials uses mixer-driven fluid dynamics also known as shear force-assisted exfoliation. Based on a high-shear rotor-stator mixer, Paton et al. (21) demonstrated a shear-assisted large-scale exfoliation for producing dispersions of graphene flakes (Fig. 7). It was believed that under high-speed rotation, the mixer can generate high shear rates in liquid containing layered bulk crystals. The exfoliation of layered bulk crystals can be triggered by the shear force induced by the high shear rates. The shear force apparatus has a mixing head constituting a rotor and a stator. By using this device, graphite can be exfoliated into graphene with lateral size of 300–800 nm in NMP using high-speed rotation of the mixing head (21). This method was also used to exfoliate BP bulk crystals into few-layer nanosheets (77,80).

    Fig. 7 (A) A Silverson model L5M high-shear mixer with mixing head in a 5 L beaker of graphene dispersion. (B) Closeup view of a DD32 mm mixing head and (C) a DD16 mm mixing head with the rotor (left) separated from its stator. (D) Photograph of graphene-NMP dispersions. (E) Wide-field TEM image of shear exfoliated graphene nanosheets (after centrifugation). Reproduced with permission from Paton, K. R.; Varrla, E.; Backes, C.; Smith, R. J.; Khan, U.; O'Neill, A.; Boland, C.; Lotya, M.; Istrate, O. M.; King, P.; Higgins, T.; Barwich, S.; May, P.; Puczkarski, P.; Ahmed, I.; Moebius, M.; Pettersson, H.; Long, E.; Coelho, J.; O'Brien, S. E.; McGuire, E. K.; Sanchez, B. M.; Duesberg, G. S.; McEvoy, N.; Pennycook, T. J.; Downing, C.; Crossley, A.; Nicolosi, V.; Coleman, J. N. Nat. Mater. 2014, 13, 624, Copyright 2014 Nature Publishing Group.

    The shear exfoliation mechanism was further revealed in terms of rotor diameter and characteristics of mixer-induced fluid dynamics (21). It was found that even when Reynolds number ReMix of the flow field is less than 10⁴, which corresponds to a not fully developed turbulent flow, well-exfoliated graphene can still be obtained. But when the shear rate is lower than 10⁴ s− 1, graphene flakes are poorly exfoliated. In the case of ReMix < 10⁴, that is, laminar flow, graphene can still be well produced if shear rate is higher than 10⁴ s− 1. Also, in case of shear mixing at a number of different combinations of rotating speed N and rotor diameter D, the minimum shear rate is also ~ 10⁴ s− 1. This suggests than any mixer that can achieve a shear rate > 10⁴ s− 1 can be used to produce graphene. The exfoliation mechanism that occurs in both laminar and turbulent regimes should be the same, and good exfoliation can happen without turbulence. Currently, the exfoliation of graphite can be achieved in liquid volumes up to hundreds of liters with a production rate of 1.44 g/h. The shear exfoliation mechanism also explained in terms of fluid dynamics where the cavitation and collision effects also favor efficient exfoliation (81). However, in the rotor-stator mixer, high shear rates are mainly localized in the gap between the rotor and stator and in the holes of the stator. This implies a well-defined localized region of high shear rate, indicating that most of the exfoliation events are localized in the vicinity of the rotor stator. In order to overcome the shortcomings of the rotor-stator mixer, a fully developed turbulence with all regions of high shear rate is necessary. To generate high shear rates in all regions, two different groups independently reported the exfoliation of graphite into graphene in aqueous surfactant solution by using a kitchen blender (22,82). A kitchen blender is a simple and commercially available rotating blade mixer, which can generate turbulence in the whole container. By optimizing the mixing parameters including the nanosheet concentration, liquid volume, rotor speed, and mixing time, the production rate cloud reaches the order of milligram per minute scale. The obtained nanosheet dimensions are in the range of 2–12 layers for thickness and 40–200 nm of lateral size. Recently, h-BN, MoS2, and WS2 nanosheets are produced by using the kitchen blender (22). These results imply that industrial rotating blade stirred tank reactors are promising for the large-scale graphene production.

    3.4 Electrochemical Exfoliation

    Electrochemical exfoliation is an efficient method to exfoliate and/or intercalate layered materials into a few-layer 2-D nanosheets (13,15,83,84). Electrochemical reactions that occurred on the layer-structured electrode would result in the intercalation and/or exfoliation of the electrode (85). There are a number of appealing features for electrochemical exfoliation such as simplicity, short processing time, easy operation, and control and potential for scale up. The quality of the exfoliated nanosheets is strongly affected by the potential applied and the electrolyte used. Depending on the type of potential applied, the electrochemical processes can be divided into either (i) cathodic exfoliation, mostly in organic solvents (propylene carbonate, DMSO, etc.) containing lithium or alkylammonium salts as electrolyte (86–88), or (ii) anodic exfoliation in ionic liquid/water mixtures or aqueous solutions of acids (e.g., H2SO4, HClO4, H3PO4, or H2C2O4) and/or inorganic salts (13–15,89–92).

    Anodic exfoliation is of particular interest since water and/or ionic liquids are used rather than organic solvents. The process also affords high and rapid production rate. A two-electrode cell is usually used for electrochemical exfoliation, with a working electrode consisting of bulk layered material, a counter electrode such as Pt, an electrolyte, and a power supply. Liu et al. (90) reported the electrochemical synthesis of graphene using aqueous solutions of imidazolium-based ILs as electrolyte. By applying 15 V potential between two graphite rods, IL stabilized graphene was obtained. Following this work, exfoliation of graphene was reported in a mixture of H2SO4 and KOH electrolyte, yielding graphene nanosheets with lateral sizes of several to 30 μm (92). More than 60% of these sheets are bilayer with A–B stacking. The exfoliated nanosheets underwent oxidation at the high potential (+ 10 V) used. This oxidation issue can be mitigated to some extent by utilizing aqueous solutions of inorganic salts in place of acids as electrolytes. For example, Parvez et al. (13) investigated a series of aqueous inorganic salts (i.e., (NH4)2SO4, Na2SO4, and K2SO4) as electrolytes for the exfoliation process (Fig. 8A and B). Exfoliation of graphite in these electrolytes produced over 85% of 1–3 layers of graphene sheets with lateral sizes up to 44 μm. More importantly, as-prepared graphene has an oxygen content of only 5.5 at.%, which is significantly lower than the graphene produced in acid electrolytes. It was suggested that electrochemical exfoliation of graphite in acid electrolytes can result in overoxidation of graphene during the exfoliation process; however, this overoxidation can be avoided by using inorganic salt electrolytes instead of acid electrolytes. The mechanism of the anodic exfoliation of graphite was also studied that occurs in three individual steps: (1) the initial oxidation of the anodes by nucleophilic hydroxyl ions (OH−) or radicals (OH•) that were electrolyzed from water, (2) intercalation of the oxidized anode by negatively charged ions and water molecules, and (3) exfoliation of the expanded anode into nanosheets by the formed gases from the electrooxidation of the intercalated ions or molecules (Fig. 8C) (13). Note that the hydroxyl radicals or ions resulting from water electrolysis tend to compromise the graphitic structure under anodic conditions. With this in mind, Yang et al. (93) introduced a series of reducing agents (i.e., (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO), ascorbic acid, and sodium borohydride) to scavenge these radicals. Notably, TEMPO-assisted exfoliation gave large graphene nanosheets with an oxygen content of only 4.5 at.%. Chen et al. (94) demonstrated the enhancement of 25 wt% in graphene yield by the addition of melamine in the electrolyte. The melamine-induced hydrophilic force from the basal plane not only facilitated exfoliation but also provided protection for graphene against oxidation.

    Fig. 8 (A) Schematic illustration of electrochemical exfoliation of graphene. (B) AFM image of electrochemically exfoliated graphene on Si/SiO 2 substrate. (C) Schematic illustration of the mechanism of electrochemical exfoliation. (D) Schematic illustration of the electrochemical Li-intercalation-assisted exfoliation method for the preparation of single- or few-layer TMD nanosheets. (E) Photographs of 2-D nanosheet (MoS 2 , WS 2 , TiS 2 , and TaS 2 ) dispersions produced by Li-intercalation-assisted exfoliation. (C) Reproduced with permission from Parvez, K.; Wu, Z. S.; Li, R.; Liu, X.; Graf, R.; Feng, X.; Mullen, K. J. Am. Chem. Soc. 2014, 136, 6083, Copyright 2014 the American Chemical Society; (E) Reproduced with permission from Zeng, Z. Y.; Yin, Z. Y.; Huang, X.; Li, H.; He, Q. Y.; Lu, G.; Boey, F.; Zhang, H. Angew. Chem. Int. Ed. 2011, 50, 11093, Copyright 2011 Wiley-VCH.

    Apart from graphene, TMDCs such as MoS2 and topological insulators (e.g., Bi2Se3 and Bi2Te3) were also electrochemically exfoliated into single- and few-layer nanosheets. Liu et al. (83) reported the electrochemical exfoliation of bulk MoS2 by applying a DC bias between MoS2 and Pt electrodes in Na2SO4 electrolyte, starting with a low-positive bias (i.e., + 2 V) to wet the bulk MoS2 followed by a larger bias (+ 10 V) to exfoliate the bulk crystal. This resulted in producing single- and few-layer 2H-phase MoS2 nanosheets with lateral sizes of 5–50 μm. In a similar approach, Ambrosi et al. (95) demonstrated the electrochemical exfoliation of bulk Bi2Te3 and Bi2Se3 crystals into single- and few-layer nanosheets.

    Unlike anodic exfoliation, which employs positive potentials, cathodic exfoliation process applied negative potential without involving oxidizing conditions, thus producing high-quality nanosheets. As shown in Fig. 8D, a Li-ion battery-like setup is generally utilized to facilitate the cathodic exfoliation. In the cell, a bulk layered material serves as the cathode, and Li foil is employed as anode to provide Li ions. With the insertion of Li+ ions into the spaces between bulk layers, the interlayer distance is expanded, hence weakening van der Waals interactions. Meanwhile, the inserted Li+ ions are reduced by electrons during the discharge process to yield metallic Li, which reacts with water to form Li(OH) and H2 gas promoting exfoliation (96). Zeng et al. (97) demonstrated the high-yield production of single-layer materials such as MoS2, WS2, TiS2, TaS2, ZrS2, and graphene, through a controllable lithiation process at room temperature (Fig. 8E). The benefit of this method is the ability to tune the level of intercalated Li+ ions by controlling the cutoff voltage at different stages. Furthermore, nanosheets of h-BN, NbSe2, WSe2, Sb2Se3, and Bi2Te3 and ternary chalcogenides including Ta2NiS5 and Ta2NiSe5 were successfully fabricated by optimizing cutoff voltage and discharge current (97,98). However, in most cases, the exfoliated nanosheets are in the form of a few layers, and further processing like extensive ultrasonication (> 10 h) or other multistep treatments (e.g., microwave irradiation) is required to get single-layer nanosheets (88,99,100).

    3.5 Oxidation-Assisted Exfoliation

    One of the low-cost methods for large-scale production of graphene is the oxidation-assisted exfoliation of graphite (101). This method is currently the most popular method for producing graphene nanosheets because of its potential scalability, high yield, and excellent dispersability in various solvents, which facilitates processability toward many applications (17,102). The basic idea underlying this method is to use strong oxidizing agents to oxidize graphite to form graphene oxide (GO). The Hummers method is the most frequently used protocol for preparing graphite oxide and its single layers, that is, GO, which are obtained after delamination in a suitable solvent by ultrasonication (103). This method consists of treating graphite with a mixture of sodium nitrate, potassium permanganate, and sulfuric acid. Over the years, efforts have been made to improve the Hummers method by eliminating the use of sodium nitrate, thus preventing the production of toxic nitrous gas (16). The addition of phosphoric acid led to a higher yield of GO with a higher degree of oxidation. The mechanism for the formation of GO was recently defined in three distinct, independent steps (104). The active oxidizing species is dimanganese heptoxide (Mn2O7), which is formed when potassium permanganate reacts with sulfuric acid (105). However, Mn2O7 is known to detonate when heated to temperatures above 55°C when in contact with organic compounds. Therefore, controlling the temperature in such a high-risk reaction is crucial particularly at the industrial scale. Recently, the preparation of GO (named as "GO-n") with an almost intact σ framework of carbon atoms was achieved by controlling the reaction temperature at 5–10°C during the oxidation step, thereby preventing the large extent of overoxidation of graphene layers (106). Under such conditions, the formation of CO2 during the oxidation step can be minimized. Moreover, cold water was continuously added to the mixture over a period of 1 day to avoid heating the concentrated sulfuric acid, which is an essential step to preserve the carbon framework in GO. However, the yield of GO-n was lower compared with that of conventionally prepared GO. The flake size of GO-n was determined to be 2–10 μm. Nevertheless, the oxidation-assisted exfoliation is very effective for the preparation of GO nanosheets; it is hard to further extend this method to other layered

    Enjoying the preview?
    Page 1 of 1