Găsiți următoarea book favorită
Deveniți un membru astăzi și citiți gratuit pentru 30 zileÎncepeți perioada gratuită de 30 zileInformații despre carte
Graphene: Preparations, Properties, Applications, and Prospects
Până la Kazuyuki Takai, Seiya Tsujimura și Feiyu Kang
Acțiuni carte
Începeți să citiți- Editor:
- Elsevier Science
- Lansat:
- Oct 15, 2019
- ISBN:
- 9780128195772
- Format:
- Carte
Descriere
Graphene: Preparation, Properties, Applications and Prospects provides a comprehensive introduction on the science and engineering of graphene. The book is composed of 9 chapters, including a discussion on what graphene is, detailed descriptions of preparation procedures, applications based on respective properties, including electrical, chemical, mechanical, thermal and biomedical, and reviews on materials derived from graphene (graphene derivatives) and other layered materials.
Provides differentiation on two kinds of graphene, graphene with highly-crystalline layers and reduced graphene oxide with highly-defective layers Thorough reviews a wide variety of preparation procedures of two kinds of graphene, including the formation of graphene foams, films and horns, and the doping of foreign atoms Contains a comprehensive review of electrical, chemical, mechanical, thermal and biomedical properties and applications based on these propertiesInformații despre carte
Graphene: Preparations, Properties, Applications, and Prospects
Până la Kazuyuki Takai, Seiya Tsujimura și Feiyu Kang
Descriere
Graphene: Preparation, Properties, Applications and Prospects provides a comprehensive introduction on the science and engineering of graphene. The book is composed of 9 chapters, including a discussion on what graphene is, detailed descriptions of preparation procedures, applications based on respective properties, including electrical, chemical, mechanical, thermal and biomedical, and reviews on materials derived from graphene (graphene derivatives) and other layered materials.
Provides differentiation on two kinds of graphene, graphene with highly-crystalline layers and reduced graphene oxide with highly-defective layers Thorough reviews a wide variety of preparation procedures of two kinds of graphene, including the formation of graphene foams, films and horns, and the doping of foreign atoms Contains a comprehensive review of electrical, chemical, mechanical, thermal and biomedical properties and applications based on these properties- Editor:
- Elsevier Science
- Lansat:
- Oct 15, 2019
- ISBN:
- 9780128195772
- Format:
- Carte
Despre autor
Legat de Graphene
Mostră carte
Graphene - Kazuyuki Takai
Graphene
Preparations, Properties, Applications and Prospects
Kazuyuki Takai
Seiya Tsujimura
Feiyu Kang
Michio Inagaki
Table of Contents
Cover image
Title page
Copyright
Preface
Acknowledgments
Chapter 1. Introduction
1.1. What is graphene?
1.2. Fundamentals of materials science for carbon materials
1.3. Construction and purposes of the current book
Chapter 2. Preparation of graphene
2.1. Chemical vapor deposition
2.2. Cleavage (peeling)
2.3. Exfoliation via graphene oxide
2.4. Other processes
2.5. Concluding remarks
Chapter 3. Electrical properties and applications
3.1. Fundamental electrical properties
3.2. Applications to information technology
3.3. Applications to social fields
3.4. Concluding remarks
Chapter 4. Chemical properties and applications
4.1. Fundamental chemical properties
4.2. Applications to energy storage and conversion
4.3. Applications to environment remediation
4.4. Concluding remarks
Chapter 5. Mechanical properties and applications
5.1. Fundamental mechanical properties
5.2. Nanolubricants
5.3. Mechanical sensors
5.4. Mechanical reinforcement
5.5. Reduced graphene oxide fibers
5.6. Concluding remarks
Chapter 6. Thermal properties and applications
6.1. Fundamental thermal properties
6.2. Thermal interface materials
6.3. Nanofluids
6.4. Thermoelectric power
6.5. Thermal energy storage
6.6. Concluding remarks
Chapter 7. Biomedical properties and applications
7.1. Biocompatibility
7.2. Cell management
7.3. Drug delivery systems
7.4. Biosensors
7.5. Concluding remarks
Chapter 8. Beyond graphene
8.1. Graphene derivatives
8.2. Single-layer materials
8.3. Layer-by-layer composites
8.4. Concluding remarks
Chapter 9. Summary and prospects
9.1. Summary on graphene
9.2. Prospects
Index
Copyright
Elsevier
Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, United Kingdom
50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States
Copyright © 2020 Elsevier Inc. All rights reserved.
No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).
Notices
Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.
Library of Congress Cataloging-in-Publication Data
A catalog record for this book is available from the Library of Congress
British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library
ISBN: 978-0-12-819576-5
For information on all Elsevier publications visit our website at https://www.elsevier.com/books-and-journals
Publisher: Matthew Deans
Acquisition Editor: Glyn Jones
Editorial Project Manager: Naomi Robertson
Production Project Manager: Sruthi Satheesh
Cover Designer: Mark Rogers
Typeset by TNQ Technologies
Preface
The Nobel Prize in Physics for 2010 was awarded to Profs. A. Geim and K. Novoselov of the University of Manchester for their groundbreaking experiments on graphene. The term graphene
was proposed in 1986 in relation to the terminology used for graphite intercalation compounds. After the Nobel Prize, scientific and technological interest in graphene increased rapidly; as a consequence, a tremendous amount of literature declaring its research target to be graphene has been published. Unfortunately, this rapid growth in interest has caused serious confusion regarding the definition and terminology of graphene even in scientific journals.
Two of the current authors, M. Inagaki and F. Kang, have authored three books in a series, entitled Materials Science and Engineering of Carbon: Fundamentals, Advanced Materials Science and Engineering of Carbon, and Materials Science and Engineering of Carbon: Characterization. It was anticipated that these books would provide a comprehensive understanding of a wide range of carbon materials (graphite, graphitized carbons, carbon blacks, activated carbons, pyrolytic carbons, glass-like carbons, porous carbons, carbon fibers, etc., in addition to diamond, fullerenes, carbon nanotubes, and graphene) through the detailed explanation and discussion of their structures, nanotextures, and fundamental properties. However, the occurrence of misperceptions and the flood-like increase in research papers on graphene-related (graphene-like) materials led these two authors to believe it was necessary to edit a book focusing on graphene-related materials, in addition to the three previously mentioned books, although graphene is a members of the carbon family. Therefore, they invited two authors, K. Takai and S. Tsujimura, who are specialists in the physical chemistry of graphene, and electrochemistry applications to biomedicals, respectively, to cover the widely spread applications of graphene in this book.
In this book, the authors attempted to provide summaries and reviews on graphene and its related materials by differentiating materials with a high perfection of structure (graphene) from those that are highly defective, even with various functional groups attached (reduced graphene oxide). In addition, it is emphasized that the number of layers governs the properties of the flakes of graphene (the characteristics of graphene) which are quite different from those of graphite (many layers stacked) and possible to be obtained on the flake of only a few highly-crystalline layers stacked.
To understand graphene and its related materials, a wide range of fundamental knowledge on various carbon materials is essential: that is, knowledge such as carbonization, graphitization, intercalation, and so on, in addition to basic knowledge about chemistry, physics, biology, and others. For the readers' convenience, it is recommended to consult the three books mentioned previously, which are published by Tsinghua University Press and Elsevier. These books will supply fundamental knowledge about carbon materials and provide an understanding of a broad range of topics in the current book.
It will be truly pleasing to all of the authors if the content of this book delivers useful information to the readers and will lead readers to the correct understanding of graphene and related materials.
Acknowledgments
The authors would like to express their sincere thanks to the people who kindly provided the data and figures for this book. The names and affiliations of the contributing persons are mentioned in the captions of figures. The authors also thank all of the people who took care of this book at Tsinghua University Press and also at Elsevier.
Chapter 1
Introduction
Abstract:
First, a discussion is presented about the definition of graphene. It shows the importance of differentiating between graphene with highly crystalline layers and reduced graphene oxide (rGO) with highly defective layers. In addition, fundamental materials science is briefly explained for carbon materials that are often used as starting materials to prepare graphene and is referred to in the literature to compare graphene and rGO.
Keywords
Carbon materials; Highly crystalline layer; Highly defective layer; Number of layers
1.1 What is graphene?
1.2 Fundamentals of materials science for carbon materials
1.2.1 Classification of carbon materials
1.2.2 Structure and nanotexture of carbon materials
1.2.3 Carbonization and graphitization
1.2.4 Carbon materials
1.2.4.1 Highly oriented graphite materials
1.2.4.2 Synthetic graphite materials
1.2.4.3 Fibrous carbon materials
1.2.4.4 Nanoporous carbons
1.2.4.5 Spherical carbon materials
1.2.4.6 Glass-like carbons
1.3 Construction and purposes of the current book
References
1.1. What is graphene?
The term graphene
was firstly proposed in 1986 [1] and then recommended by the International Union of Pure and Applied Chemistry [2] as the name for a single two-dimensional layer of carbon atoms bonding using sp² hybrid orbitals, which occurs in graphite intercalation compounds. The single carbon layer occurring in the intercalation compounds was proposed to be called graphene
which comes from the suffix -ene
for polycyclic aromatic hydrocarbons such as naphthalene and anthracene and the prefix graph-
from graphite. In the earliest literature [1], the authors noted that it should be adopted for graphite intercalation compounds.
In the first-stage structure of the compounds, a two-dimensional (2D) carbon layer is sandwiched by two intercalate layers and isolated from other carbon layers, although more than two carbon layers are stacked in parallel with regularity in compounds with higher than a second stage, as schematically illustrated in Fig. 1.1.
Graphene
is commonly defined as an isolated single layer of carbon hexagons consisting of sp² hybridized C–C bonding with π-electron clouds. From an engineering point of view, thin flakes consisting of a few layers of carbon atoms, including single-layer graphene, could be important because of their interesting structural, chemical, and physical characteristics. The promising potential applications of graphene and its related materials have been pointed out in technological fields.
Figure 1.1 Stage structure of graphite intercalation compounds.
Numerous reports have been published, particularly after the Nobel Prize in Physics was awarded in 2010 to Profs. A. Geim and K. Novoselov of the University of Manchester for their groundbreaking experiments in graphene [3]. Sudden scientific and technological interest in graphene and its related materials has caused some confusion about the definition and terminology of graphene-related materials, even in scientific journals. A proposal regarding the nomenclature for two-dimensional carbon materials was been presented in the journal Carbon [4]. In much of the literature, however, the term graphene
has not been used according to its strict definition, i.e., a single layer of carbon atoms consisting of sp² hybridized bonds. Some authors do not pay enough attention to how many layers are stacked in their materials, although they have called them graphene
Therefore, here, the terms single-layer graphene
(or monolayer graphene), double-layered graphene
(or bilayered graphene), and multilayered graphene
are used only for products that have confirmed numbers of stacked layers.
Numerous reviews have been published from various viewpoints: providing an overview of graphene by pointing out what is fascinating about it [5–13], focusing on its production in relation to its structure [14–23], applications in electronics [24–30], energy storage applications [31–48], biological applications including sensors [49–68], functionalization including the formation of composites [69–80], mechanical applications [81,82], thermal properties [83,84], doping to improve its properties [85,86], and its toxicity [87,88].
Before summarizing research on graphene-related materials and discussing the results, it is worthwhile to understand the flakes and/or films that determine how many numbers of layers can exhibit the unique functionalities characteristic of graphene, to differentiate it from graphite. Fig. 1.2 compares Raman spectra are for graphite and graphene flakes with different numbers of stacked layers [89]. As shown in Fig. 1.2A, highly oriented pyrolytic graphite (HOPG), for example, shows a sharp and strong G-band and G′-band (2D band). The latter is presented as decomposing into two bands, G′1 and G′2, with heights of roughly ¼ and ½ of the G-band, respectively. On the other hand, single-layer graphene exhibits a sharp and single profile for both G- and G′-bands and the G′-band is much stronger than the G-band by almost four times. As shown in Fig. 1.2B, the G′-band gradually shifts to a higher position, changing its profile to unsymmetrical and decreases its intensity relative to the G-band with an increasing number of stacked layers.
A marked field effect on electrical conduction is a characteristic of graphene. Pioneering works were reported on few-layered graphene flakes [90]. Fig. 1.3A shows resistivity change with gate voltage Vg on few-layered graphene flakes as a function of temperature, revealing a marked field effect, a peak of several kΩ with rapid decay to about 100 Ω at high Vg, and a strong dependence on temperature. This field effect is shown by plotting the relative carrier concentration n/n0 (n0 is the concentration of carriers at 4 K) against temperature for three flakes, few-layered flakes, few-layered but much thicker flakes, and multilayered flakes in Fig. 1.3B. The results clearly reveal that the field effect depends on the number of stacked layers: the smaller the number of layers, the stronger the field effect.
Figure 1.2 Raman spectra measured by a laser with 514-nm wavelength for graphite to single-layer graphene [89] : (A) comparison of graphite and single-layer graphene, (B) Change in G′-band with increasing number of stacked layers. 2D , two-dimensional; a.u., arbitrary units.
Figure 1.3 Field effect for few-layered graphene flakes [90] .
Figure 1.4 Change in thermal conductivity at room temperature with number of layers stacked [91] .
Extremely high thermal conductivity, more than 5000 W/m.K, is expected for single-layer graphene [91]. It depends on the number of layers stacked in the flake. In Fig. 1.4, thermal conductivities experimentally determined from the temperature dependence of the position of the micro-Raman G-band together with theoretical predictions are plotted against the number of stacked layers [91]. A marked dependence of room temperature thermal conductivity on the number of layers is clearly shown. Single-layer graphene has extremely high conductivity, but that of a flake composed of four layers is comparable to bulk graphite with high crystallinity, such as HOPG, although it is high compared with metals.
These three experimental results suggest that we need a flake and/or film composed of less than four layers to obtain characteristics intrinsic to graphene.
1.2. Fundamentals of materials science for carbon materials
Graphene is a member of carbon materials, which include graphite, carbon blacks, carbon nanotubes, carbon fibers, activated carbons, porous carbons, diamond, and fullerenes. To prepare graphene flakes and films, graphite materials such as natural graphite, HOPG, and kish graphite have been employed as starting materials. In addition to graphene, other carbon materials such as activated carbons and porous carbons including carbon foams and carbon nanotubes have been cited in references and/or competitive materials.
Before discussing graphene, the fundamentals of carbon materials science will be briefly explained, emphasizing highly oriented graphite materials that have been used as raw materials of graphene. For a detailed explanation and discussion of carbon materials, readers of this book are suggested to refer to fundamental books written by the current authors (M.I. and F.K.) [92,93].
1.2.1. Classification of carbon materials
Many kinds of carbon materials have been manufactured, synthesized, and widely used in various industries. These carbon materials are proposed to be classified on the basis of their chemical bonds of constituent carbon atoms using sp³, sp², and sp hybrid orbitals. The sp² hybrid bonding of carbon atoms results in two structures: flat layers composed of six-membered carbon rings, which had been represented by graphite but now include graphene; and curved layers, which introduce five-membered carbon rings into six-membered rings, as occurs in fullerenes. Layers composed of sp² orbitals, both flat and curved, are intrinsically anisotropic and have π-electron clouds at both sides of the layer, which creates broad diversity in the structure and properties of the materials. Carbon nanotubes can be placed between fullerene and graphene, because the tips of the tube include five-membered rings (fullerene-like) and its wall is composed of six-membered rings although it is rolled up (graphene-like). Fig. 1.5 shows the classification of carbon materials based on hybrid bonds, together with the diversity of each material.
Because of the anisotropic nature and the presence of π-electron clouds in the carbon materials composed mainly of sp² hybrid orbitals, the number of layers stacked in parallel has a strong influence on their properties. The importance of the number of stacked layers has been pointed out for carbon nanotubes and fullerenes, and now for graphene, as mentioned in the previous section. Large numbers of layers stacked with regularity have been called graphite, and various graphite-related materials have been produced in industries and used as important industrial materials, some of which are listed in Fig. 1.5.
Figure 1.5 Classifications and diversity of carbon materials.
1.2.2. Structure and nanotexture of carbon materials
Most carbon materials composed of flat layers using sp² hybrid orbitals consist of small units of layers stacked in parallel, as shown by the transmission electron microscopy (TEM) lattice fringe image and schematic illustration in Fig. 1.6. These units are called basic structural units (BSU) or crystallites. In a unit, two kinds of layers coexist with stacking regularity: random and regular stacking. The former is called turbostratic stacking and the latter is graphitic stacking (usually written as AB stacking). These BSUs are strongly anisotropic in the nature of their bond, with strong covalent bonding using sp² hybrid orbital along the layer and weak van der Waals bonding across the layers.
The aggregation of these anisotropic nanosized BSUs in the particles results in different textures owing to the different schema of preferred orientations of anisotropic BSUs (i.e., planar, axial, point, and random orientations), as illustrated with some representative carbon materials in Fig. 1.7. The aggregation of BSUs in different schema is called nanotextures. By the planar orientation scheme, films and platelets of highly oriented graphite are produced and some coke particles are principally composed of planar orientation of BSUs. By the axial orientation scheme, fibrous carbon materials are produced from different precursors such as carbon nanotubes and vapor-grown carbon fibers with a coaxial mode of orientation and some mesophase pitch-based carbon fibers with a radial mode. By the point orientation scheme, various carbon spheres are produced, as well as various sized fullerene particles and different nanosized carbon blacks with concentric mode and mesophase spheres with a radial mode. The particles composed of these oriented nanotextures are still anisotropic. In addition, random aggregation of small BSUs occurs in so-called glass-like carbon (glassy carbon), the particles of which are isotropic in nature.
Figure 1.6 Basic structural unit of carbon materials: (A) lattice fringe image and (B) schematic illustration of a unit.
Figure 1.7 Nanotextures in carbon materials on the bases of preferred orientations of basic structural units.
1.2.3. Carbonization and graphitization
Most carbon materials used in industry are produced from organic precursors such as pitches, biomasses, and organic polymers via heat treatment at high temperatures in an inert atmosphere during carbonization and graphitization processes. Fig. 1.8 summarizes changes in chemical and electronic band structures.
Carbonization is performed after pyrolysis of the precursors from 800 to 2000°C, in which the BSUs are formed and their basic aggregation scheme (nanotexture) is established, accompanied by the emission of foreign atoms, oxygen, hydrogen, and nitrogen as gases and the polycondensation of six-membered carbon rings. Because the nanotexture of most carbon materials is established, this process is most important in the production of various carbon and graphite materials. The nanotexture governs the development of crystalline structures in carbon materials during graphitization. During this process, a large amount of shrinkage and the rapid emission of gas species occur that are associated with cracking of carbon particles that occur in many cases. Therefore, the process of carbonization is applied separately from that of graphitization.
Above 2000°C, a change in crystalline structure (the development of a graphite structure) mainly occurs. The development of a graphite structure may be evaluated by different techniques, including x-ray diffraction (XRD), electromagnetic property measurements, Raman spectroscopy, and high-resolution TEM. In BSUs formed during carbonization, turbostratic stacking with an interlayer spacing of about 0.342 nm is randomly changed to graphitic regular stacking with a spacing of 0.3354 nm (the spacing in the graphite crystal) with an increasing heat treatment temperature (HTT) of above 2000°C, which is measured as a decrease in the averaged interlayer spacing, d002, by XRD, associated with the growth of BSU sizes (crystallite size) along the a- and c-axes, La and Lc. The change in d002 with HTT depends on the materials after carbonization (carbon materials). Fig. 1.9 shows the changes in these parameters with HTT for various carbon materials. In needle-like coke with a planar orientation scheme, d002 decreases quickly, approaching the value of a graphite crystal (0.3354 nm), and Lc and La grow rapidly. In glass-like carbon with a random orientation scheme, in contrast, there is almost no decrease in d002 and no appreciable growth in Lc and La even after 3000°C treatment (i.e., no development of a graphite structure). Carbon blacks with a point orientation scheme have intermediate behaviors: the large-sized thermal black shows more improvement in structure than does the small-sized furnace black.
Figure 1.8 Schematic illustration of chemical, crystallographic, and electronic band structures in carbon materials with a planar orientation nanotexture. MW , molecular weight.
Figure 1.9 Changes in x-ray diffraction parameters with heat treatment temperatures (HTT) for various carbon materials: (A) d 002 , (B) L c measured from a 002 diffraction peak, and (C) L a from a 110 peak.
Diffraction peaks of XRD for carbon materials are classified into three groups: 00l, hk0, and hkl, mainly owing to the strong anisotropy of BSUs and the coexistence of two interlayer spacings, as shown in Fig. 1.6B. Diffraction peaks with indices of 00l give averaged interlayer spacing, which decreases gradually from more than 0.344 nm for highly defective layers to 0.3354 nm for graphitic stacking and about 0.342 nm for turbostratic stacking with increasing HTT: in other words, with improving crystallinity of carbon, as shown in Fig. 1.9A. In turbostratic stacking of layers, there is no 3D regularity in stacking (i.e., no l index is defined), and so the diffraction peaks with hk0 indices for a graphitic structure are expressed as hk diffraction peaks for carbon materials mainly consisting of the turbostratic stacking of layers. Regular and random stackings, graphitic and turbostratic, are clearly demonstrated in diffraction profiles of hk0 and hk peaks, as shown for 101 and 10 peaks in Fig. 1.10. The profile of the 10 peak for a turbostratic structure shows a characteristic unsymmetrical profile, and that of a 100 peak for a graphitic structure is sharp and symmetrical and is associated with a 101 peak owing to the formation of 3D stacking regularity. With increasing HTT, an unsymmetrical 10 peak is modulated by the appearance of a 101 peak, together with its sharpened and improving symmetry. Peaks with indices of hkl are caused by the graphite structure, and so 112 peak is often selected as to indicate the formation of a graphite structure, because the 112 peak does not overlap with other peaks although the 101 peak overlaps with the 100 peak, as can be seen in Fig. 1.10.
Figure 1.10 (A) Correspondence between x-ray diffraction (XRD) profile of hk line and (B) transmission electron microscopy (TEM) image. BSU , basic structural units.
Because most carbon particles are more or less anisotropic, except for glass-like carbon with a random orientation scheme, their aggregation into a block gives texture on an larger scale, which may be called microtexture and macrotexture. The technique for evaluating these micro- and macrotextures has not yet been established. Figs. 1.11 and 1.12 demonstrate examples of these textures, which must be controlled for practical applications. In Fig. 1.11, scanning electron microscopy (SEM) images of cross-sections of commercially available isotropic high-density graphite are shown, demonstrating different pore structures [94]. Image processing of these micrographs suggests a relation between pore structure parameters (for example, the pore area) and the mechanical properties of these graphite blocks. To prepare carbon fiber–reinforced plastics and carbon fiber–carbon composites, different macrotextures based on the orientation of carbon fibers are employed to obtain high strength and a high modulus of the composites, as shown in Fig. 1.12.
Figure 1.11 SEM images of the cross-sections of isotropic high-density graphite blocks [94] .
Figure 1.12 Different scheme for reinforcing carbon fibers in composites.
1.2.4. Carbon materials
1.2.4.1. Highly oriented graphite materials
The blocks or platelets, which are composed of big carbon layers stacked in large amount with graphitic regularity, are called highly oriented graphite; the extreme case of this is a single crystal of graphite. In practice, however, it is difficult to obtain single large crystals and almost impossible to get those that are more than a few square millimeters. There are only two ways to find single crystals of graphite: in natural graphite ores and in so-called kish graphite. Resources of high-quality natural graphite are limited on earth and exist in Sri Lanka, Madagascar, and China. Even in these ores, the possibility of finding single crystals of a certain size is slim. Kish graphite is formed by the precipitation of supersaturated carbon from molten iron and cannot be large in size, but some have very high crystallinity and can be called single crystals. Alternatives to single-crystal graphite are HOPG and graphite films derived from some organic polymer precursors such as polyimides via high-temperature treatment.
Natural graphite: Natural graphite is usually recovered as a powder from natural ores through milling and purification processes. After the final purification process, the powder has usually a purity of more than 99 wt%. Some of it consists of flaky particles with a highly crystalline structure and highly oriented nanotexture (flaky graphite), but some consists of aggregates of small crystals (called amorphous graphite) [95]. These two kinds of natural graphite can be differentiated by XRD, as shown in Fig. 1.13, which shows flaky graphite with sharp 100 and 101 peaks but a microcrystalline graphite broad 101 peak. Some flaky graphite contains metastable rhombohedral modification of the graphite structure, possibly as a result of shear stress during milling.
Figure 1.13 X-ray diffraction patterns of flaky and microcrystalline natural graphite [95] .
Kish graphite: A part of carbons dissolved into molten iron at high temperatures is incorporated into the crystal lattice of iron to form alloys (different steels) and another part segregates as graphite. Graphite flakes segregated from molten iron are called kish graphite [96–98]. Relatively large amounts of kish graphite flakes are obtained as a by-product during steel production; all of them do not always have high crystallinity because it depends on the segregation conditions. When they are produced at the temperature at which iron evaporates, kish graphite flakes have a single crystal nature. As shown in Fig. 1.14, flakes are thin with an irregular shape. Regular stacking of layers is observed by SEM and their single crystal nature is confirmed by a well-organized electron channeling pattern. The high crystallinity of kish graphite flakes was also confirmed by measurements of the resistivity ratio, ρ300 K/ρ4.2K, maximum transverse magnetoresistance, (Δρ/ρ0)max, measured at 77 K, and Shubnikov-de-Haas oscillation in magnetoresistance.
Graphite single crystals have been synthesized from molten iron by controlling the segregation process [99,100] and from Al4C3 by transporting decomposition gases [101,102].
Highly oriented pyrolytic graphite: Carbon deposited on a substrate by chemical vapor deposition (CVD) of hydrocarbon gases such as methane and propane at high temperatures can have a well-oriented texture and is called pyrolytic carbon. Pyrolytic carbons have extensively been studied to control their structure and texture by applying different deposition conditions, such as precursor hydrocarbon, the concentration and flow rate, the temperature of deposition, and the geometry of the furnace [103]. By hot-pressing at high temperatures under high pressure, the preferred orientation of graphite crystallites and their crystallinity can be markedly improved. Typical conditions are shown in Fig. 1.15 [104]. The products of this process are called HOPG. To achieve high crystallinity and high orientation, the starting pyrolytic carbons need selected, as well as the hot-pressing conditions.
Figure 1.14 Kish graphite : (A) optical micrograph, (B) scanning electron microscopy image of the edge surface, and (C) electron channeling pattern on the basal plane.
Courtesy of Prof. Y. Hishiyama of Tokyo City University.
Figure 1.15 Procedure of the production of highly oriented pyrolytic graphite.
As shown in Fig. 1.16A, HOPG has a highly oriented nanotexture along its plate, but electron channeling analysis suggests that HOPG consists of randomly oriented a-axes of crystalline domains, although their c-axes are almost perfectly oriented perpendicular to the plate. The channeling pattern on the surface of the plate (Fig. 1.16B) is somewhat distorted compared with that of kish graphite (refer to Fig. 1.14C) and the channeling contrast shows the random orientation of a-axes of the domains (Fig.1.16C).
Graphite films derived from organic films: Highly crystallized graphite films were prepared from films of limited numbers of organic precursors, such as poly(p-phenylene vinylene) (PPV) [105], poly(p-phenylene-1,3,4-oxadiazole) (POD) [106,107], benzimidazobenzo-phenanthroline ladder polymer BBL [107–110], and the commercially available aromatic polyimides Kapton and PPT [111–113], the repeating units of which are shown in Fig. 1.17. Heat treatment of PPV-derived carbon films up to 3000°C gave a graphite film; as shown in its Raman spectrum in Fig. 1.18A, no D-band was detected [105]. The film of polyoxadiazole heat-treated at 2800°C gave sharp and strong 00l diffraction profiles and no hk0 diffraction lines by reflection mode and well-resolved hk0 and hkl lines and no 00l lines by transmission mode, as shown in Fig. 1.18B; this revealed the formation of a highly oriented, well-crystallized graphite film [106].
Figure 1.16 Highly oriented pyrolytic graphite: (A) scanning electron microscopy image of the cross-section, (B) electron channeling pattern, and (C) electron channeling contrast.
Courtesy of Prof. A. Yoshida of Tokyo City University.
Figure 1.17 Molecular repeating units of polymers giving graphite films. PPV , poly( p -phenylene vinylene); POD , poly( p -phenylene-1,3,4-oxadiazole); BBL , benzimidazobenzo-phenanthroline ladder polymer; Kapton; PPT , commercially available aromatic polyimide films.
BBL polymer film was cast onto a glass substrate from its solution of trifluoromethanesulfonic acid (TFMSA) and then heat-treated at high temperatures to 3200°C by sandwiching it between two artificial graphite plates [110]. The resultant ultrathin films, which were less than 100 nm thick, showed an electron channeling pattern similar to that of HOPG and an orientation of graphite basal planes parallel to the film surface, as shown in Fig. 1.19A. By using methanesulfonic acid as solvent for the BBL polymer, film with an orientation of basal planes perpendicular to the film surface was obtained, as shown in Fig. 1.19B [109].
Two films derived from polyimide polytrimethylene terephthalate (PTT) and Kapton by heat treatment at 3200°C under a simple mechanical constraint (sandwiching between two graphite plates) gave high crystallinity comparable to kish graphite and HOPG; ρRT/ρ4.2K (one parameters to evaluate crystallinity of graphite) was 4.90 and 4.79 for PTT- and Kapton-derived film, respectively, although it was 4.7–5.5 for kish graphite and HOPG [112].
Figure 1.18 (A) Change in Raman spectrum with heat treatment temperature for poly( p -phenylene vinylene) film [105] and (B) x-ray diffraction patterns of polyoxadiazole film heat-treated at 2800°C [106] . a.u. , arbitrary units; cps , counts per second.
Figure 1.19 002 lattice fringe images of the cross-section of graphite films prepared from a basic structural units polymer using trifluoromethanesulfonic acid (A) and methanesulfonic acid (B) as solvents and heat-treated at 2800°C [109] .
Exfoliated graphite and flexible graphite sheets: Because large flakes of natural graphite are not commonly available and natural graphite flakes cannot be formed directly into a sheet, a technique for preparing graphite sheets with no binding materials was developed by forming graphite oxides and via their thermal exfoliation and reduction, followed by compression into sheets. They are called flexible graphite sheets and have promoted the application of graphite [114]. Graphite sheets have characteristic advantages such as flexibility, resilience, and the ability to be formed into various shapes easily; in addition, graphite has intrinsic properties such as lubricity, chemical and thermal stability, and high electrical and thermal conductivity.
Natural graphite flakes are chemically or electrochemically intercalated to form a covalent intercalation compound of graphite, graphite oxide, through the use of a mixture of concentrated sulfuric and nitric acids. After rinsing and drying, residue compounds are obtained in which the regular stage structure is lost but some sulfuric acid derivatives remain in the graphite gallery. These residue compounds are then rapidly heated to 900–1200°C, in which intercalates remaining in the graphite gallery are decomposed into gaseous products, yielding marked exfoliation of the pristine graphite. This exfoliated graphite consists of worm-like particles, as shown in Fig. 1.20. They are formed by preferential exfoliation perpendicular to the plane surface of the pristine flakes. Worm-like particles of exfoliated graphite are either molded or rolled into a sheet with no adhesives or binders. These are called flexible graphite sheets and are widely used as seals, gaskets, and packings.
After rinsing and drying of intercalation compounds, the residue compounds are commercially available as expandable graphite and are used as the starting material for thin flakes of graphene oxide in some literature.
1.2.4.2. Synthetic graphite materials
So-called graphite materials have been manufactured industrially into variously sized blocks, rods, and plates for different applications, such as large degree electrodes for electric arc furnaces, various jigs for processing metals, crucibles and heating elements to grow semiconductor crystals, brushes for electric motors, and neutron moderators for nuclear reactors. Graphite materials for these applications have mostly been produced using coke particles as fillers with pitches as the binder through carbonization and graphitization treatment at high temperatures. The production processes are summarized as a block diagram in Fig. 1.21. Briefly, pulverized coke particles are mixed and kneaded with binder pitch at a temperature slightly higher than the softening point of the pitch, in which the nanotexture (either needle-like or mosaic) and particle size of the filler coke and the mixing ratio of the filler to the binder have to be controlled according to the requirements of the applications. After the mixture is formed by extrusion, molding, or isostatic compression, it is heat-treated at 1400–1800°C (calcination or carbonization), followed by heat treatment at a higher temperature up to 2800–3000°C (graphitization) (see Fig. 1.8). The final products of this manufacturing process are called synthetic (artificial) graphite materials. All synthetic graphites are polycrystalline solids; of these, BSUs (crystallites) are much smaller than those in the highly oriented graphites described previously. To produce graphite materials for use in an electric arc furnace as an electrode, coarse-grained coke particles with a needle-like texture are usually employed to achieve high thermal-shock resistance. They are formed by extrusion or molding, and the slight orientation of coke particles cannot be avoided. To produce isotropic high-density graphite blocks, however, fine-grained coke particles have to be used to achieve a high density, and isostatic pressing must be employed for an isotropic nature in the final block. Microtextures of two graphite materials, electrode grade and isotropic grade, are compared in Fig. 1.22.
Figure 1.20 (A) Scanning electron microscopy images of worm-like particles of exfoliated graphite. (B) Appearance of the particles. (C) Cross-section of a particle. (D) Distribution of pore area on cross-section.
Figure 1.21 Production process of synthetic graphite materials.
Most synthetic graphite materials consist of filler coke and a carbonized binder (binder coke). The former is able to attain a higher degree of graphite structure development (graphitization degree) compared with the latter. The degree of graphitization achieved after final graphitization depends on the nanotexture of the starting coke particles, as shown in Fig. 1.23A. Needle-like coke has an extended flow pattern owing to the preferred orientation of crystallites, and so there is a higher degree of graphitization after high-temperature treatment compared with mosaic coke, which has a short range of orientation and is less ordered (Fig. 1.23B). However, the degree of graphitization reached by needle-like coke is far lower than that of highly oriented graphite materials.
Figure 1.22 Polarized light micrographs of cross-sections of synthetic graphites: (A) electrode grade and (B) isotropic grade.
Figure 1.23 Polarized light micrographs of cross-section of particles: (A) needle-like and (B) mosaic coke.
1.2.4.3. Fibrous carbon materials
A variety of fibrous carbon materials have been prepared by different techniques: electric arc-discharge, catalytic CVD, template-assisted CVD, melt-spinning, and electrospinning. Table 1.1 classifies these fibrous carbons according to whether the 002 lattice fringes of carbon layers are short or long and whether their diameters are on the micrometer or nanometer scale. Carbon layers constituting the wall of carbon nanotubes (CNTs) are long and oriented exactly parallel to the tube’s axis, but layers of carbon nanofibers are short and oriented in different modes, called tubular, herringbone, and platelet types, depending on the methods and conditions of synthesis. Carbon fibers with a diameter of about 7 μm are produced industrially by melt-spinning from different precursors, polyacrylonitrile (PAN), isotropic and anisotropic (mesophase) pitches, cellulose, and phenol as filament yarns, clothes, chops, webs, etc. Vapor-grown carbon fibers on the market are produced by catalytic CVD and have tubular nanotextures. By applying electrospinning, various organic polymers can be used as precursors to fabricate carbon nanofibers [115].
Table 1.1
Figure 1.24 Transmission electron microscopy images of CNTs: (A) by catalytic chemical vapor deposition [116] and (B) by electric arc discharge [119] .
Carbon nanotubes: CNTs are synthesized by catalytic CVD during the initial stage of vapor-grown carbon fibers [116,117] and are found in carbon deposits on graphite anodes during arc discharge [118–120], as shown in Fig. 1.24. Under similar arc discharge conditions, a scroll of carbon layers were obtained in 1960, called graphite whiskers [121]. The disproportionation reaction of CO at high pressures and high temperatures with an iron catalyst produces single-walled CNTs (SWCNTs), which are almost free of amorphous carbon [122,123]. Double-walled CNTs (DWCNTs) are successfully obtained at a high yield by catalytic CVD using an Fe catalyst deposited on MgO [124,125].
CNTs aligned perpendicular to the substrate (called arrays) with well-defined patterns were synthesized by CVD of ethylene at 700°C on an Fe thin film (5 nm thick) on an Si substrate [126]. Accelerated growth of SWCNTs and DWCNTs is achieved by using an alcohol as the carbon precursor or by adding water vapor to the precursor gas [127–129]. Vertically aligned SWCNTs (called forest) were grown from ethylene on various metal catalysts in either Ar/H2 or He/H2 by adding a small amount of water vapor to atmospheric gas, as shown in Fig. 1.25. These SWCNT forests grew at a superior high rate [130,131].
Transparent thin sheets 3.4 cm wide, composed of well-aligned MWCNTs, were pulled from the side face of a CNT array [132]. Long yarns and ribbons were continuously spun from the SWCNT array at room temperature [133], passing through a volatile liquid during spinning, which was effective in making the spun yarns dense [134].
Carbon fibers: Commercially available carbon fibers have different nanotextures, particularly in their cross-sections, as illustrated in Fig. 1.26. Carbon fibers derived from isotropic pitch, phenol, and cellulose have a random orientation in cross-sections both parallel and perpendicular to the fiber axis. On the other hand, vapor-grown carbon fibers (VGCFs) consist of well-oriented carbon layers that are parallel along the fiber axis and concentric in the perpendicular cross-section. Mesophase pitch–based carbon fibers are able to control their nanotextures in cross-section, such as straight radial, corrugate radial, and concentric, and consequently their properties.
Figure 1.25 Single-walled carbon nanotube (SWNT) forest [130] .
Figure 1.26 Schematic illustration of nanotexture in carbon fibers. PAN , polyacrylonitrile.
Depending on the nanotexture, structural changes with high temperature treatment are different for each carbon fiber. Fig. 1.27 compares changes in interlayer spacing d002 (one parameters for evaluating the degree of graphitization) with HTT on PAN-based carbon fibers, VGCFs, and anthracene-derived coke. Graphitization was depressed on PAN-based carbon fibers with a random nanotexture, whereas it proceeded rapidly on VGCFs about 10 μm in diameter with a well-developed axial nanotexture, comparable to anthracene-derived coke with a planar nanotexture. The diameter of the fibers has an influence on the development of the graphite structure, VGCFs with a 10-μm diameter have a much higher degree of graphitization than do those 2 μm in diameter.
Figure 1.27 Changes in interlayer spacing d 002 with heat treatment temperature (HTT). PAN , polyacrylonitrile; VGCFs , vapor-grown carbon fibers.
Courtesy of Dr. N. Iwashita.
1.2.4.4. Nanoporous carbons
Pores commonly exist in carbon materials; they are categorized according to whether a specific gaseous molecule, usually nitrogen, can be adsorbed in open and closed pores. Open pores are classified by their widths: micropores less than 2 nm wide, mesopores of 2–50 nm, and macropores more than 50 nm. These pores are controlled by oxidation. The process and the resultant porous carbons are called activation and activated carbons, respectively. After the development of template-assisted carbonization processes, many template materials were proposed to control the pore structure. This process has the advantages of a high carbon yield because there is no scarification of matrix carbon by oxidation (gasification), a homogeneous pore size, easy control of the pore volume, high reproducibility of pore structure, etc.
Activated carbons: The history of activated carbons goes back to the prehistoric era, when charcoals are known to have been used to purify water and as a medicine to stop diarrhea. Granular activated carbons are now prepared from different precursors, including biomasses, and are used in a wide range of industries. Different processes of activation were employed, such as oxidation using diluted oxygen gas, air, water vapor, etc. (physical activation), and oxidation using ZnCl2 and KOH (chemical activation). Using KOH, an extremely high surface area of about 3600 m²/g was reported. In most carbon materials, macropores (>50 nm in size) and mesopores (2–50 nm) coexist with micropores, as shown schematically in Fig. 1.28A. Macropores and mesopores are influenced by the performance of their adsorption as pathways to micropores for adsorbates.
Figure 1.28 Illustrations of the pore structure in (A) in granular activated carbon and (B) activated carbon fiber.
Fibrous activated carbons, or activated carbon fibers (ACFs), have been prepared from different carbon fibers, such as PAN-based, isotropic pitch–based and phenol-based carbon fibers, the pore structure of which is schematically shown in Fig. 1.28B. ACFs have the advantage of a high adsorption–desorption rate because almost all micropores are open at the surface of the fiber where adsorbate molecules can be adsorbed directly into micropores, in contrast to granular activated carbons, for which adsorbates have to pass through the macropore and mesopore to reach the micropore (Fig. 1.28A).
Templated porous carbons: Template-assisted carbonization was successfully applied to prepare nanoporous carbons [135,136]. Table 1.2 summarizes templates classified as hard or soft, together with commonly used carbon precursors. The resultant carbons with pore structure characteristics (microporous or mesoporous, the representative Brunauer–Emmett–Teller (BET) surface area, SBET, and the total pore volume, Vtotal) and the cycle performance of the templates.
Ordered micropores are created by using zeolites. Ordered mesopores are made using mesoporous silicas and block copolymers, but MgO, colloidal silica, metal-organic frameworks (MOFs) etc. result in disordered mesopores, together with micropores. Representative SEM images of MgO-templated and mesoporous SBA 15–templated carbons are shown in Fig. 1.29A and b, respectively. An MgO template has two advantages: it can be removed from the carbon matrix simply by washing with diluted acidic aqueous solutions, and it can be recycled, if needed, in contrast to other templates, zeolites, mesoporous silicas, and colloidal silicas, which need either HF or concentrated NaOH for removal, and block copolymers and MOFs, which are sacrificial.
Table 1.2
AN, acrylonitrile; SBET, Brunauer–Emmett–Teller surface area; EOA, triethyl orthoacetate; FA, furfuryl alcohol; MOF, metal-organic frameworks; MP, mesophase pitch; PET, poly(ethylene terephthalate); PF, phenol-formaldehyde; PhF, phloroglucinol-formaldehyde; PVA, poly(vinyl alcohol); RF, resorcinol-formaldehyde; Vtotal, total pore volume.
Figure 1.29 Scanning electron microscopy images of templated carbons: (A) MgO-templated and (B) mesoporous silica (SBA-15)-templated.
1.2.4.5. Spherical carbon materials
Various spherical carbon materials with a wide range of diameter have been developed that have different nanotextures based on their point orientation schema, whether radial or concentric (see Fig. 1.7). Fig. 1.30 shows the nanotextures of three representative carbon spheres. The concentric alignment of BSUs is found in carbon blacks (Fig. 1.30A), although the nanotexture becomes random at the center of the sphere. Radial alignment (Fig. 1.30C) is found in carbon spherules, which are formed from a mixture of poly(ethylene) and poly(vinyl chloride) by pressure carbonization [137]. An intermediate nanotexture, in which the alignment of BSUs is radial at the surface of the sphere but planar at the center, occurs in mesocarbon microbeads, as explained later.
Carbon blacks: Carbon blacks are formed through incomplete combustion in the gas phase of either gaseous or mist-like hydrocarbons [138]. Carbon blacks are classified on the basis of the reaction process and the raw materials as furnace blacks, channel blacks, lamp blacks, thermal blacks, acetylene blacks and Ketjenblack. They are characterized by different-size spherical primary particles and their coalescence into aggregates (secondary particles). Carbon black has the smallest primary particle size with the most marked aggregation, as shown in TEM images in Fig. 1.31.
Furnace blacks are produced from either the mist of creosote oil or natural gas consisting mainly of methane by their incomplete combustion at 1200–1400°C; they have marked aggregation of primary particles (called structure
in the industry) (Fig. 1.31A). This is the main reason why they are used to reinforce rubber. Lamp blacks are manufactured by burning aromatic oils in shallow open pans with a limited air supply; they are mainly used for inks. Thermal blacks are produced by the thermal decomposition of natural gas; the size of the primary particles is usually large, as shown in Fig. 1.31B, and almost no aggregation is observed. Acetylene blacks are formed through the exothermic decomposition of acetylene gas at a relatively high temperature, 2400°C, in which the primary particles are highly aggregated (Fig.1.31C). Under conditions similar to those for furnace blacks, carbon blacks with high aggregation, compared with acetylene blacks, are produced and named Ketjenblack [139], as shown in Fig. 1.31D. Acetylene black and Ketjenblack are often used as electric conductive additives for electrodes in electrochemical devices because their well-developed aggregation and relatively high electrical conductivity.
Figure 1.30 Nanotextures in spherical carbons.
Figure 1.31 Transmission electron microscopy images of carbon blacks.
Figure 1.32 Mesophase spheres. (A) Polarized light micrograph of spheres formed in a pitch, and (B) scanning electron microscopy image of the spheres separated from matrix pitch (mesocarbon microbeads).
Mesocarbon microbeads: Anisotropic spherical particles are segregated from pitches during heating, as shown in Fig. 1.32A. They are called mesophase spheres and coalesce into so-called bulk mesophase to form coke [140]. The structure of mesophase spheres is close to a radial point orientation scheme, but in its center, the orientation of layers is planar, as illustrated in Fig. 1.30B [140,141]. Mesophase spheres formed in pitches are separated from an isotropic matrix using a solvent, either pyridine or quinoline [142], and are called mesocarbon microbeads (MCMB) (Fig. 1.32B).
The graphitization behaviors were studied for two kinds of MCMBs. One type was separated from quinoline-soluble parts in coal tar pitch after heat treatment at 430°C for 120 min, and the other was from asphalt pitch after heat treatment at 430°C for 60 min [143]. After they were separated from the matrix pitch, MCMBs were graphitized by heat treatment at high temperatures. Their changes were almost the same as those of cokes prepared from the pristine pitches, as shown for coal tar and asphalt pitches in Fig. 1.33A and B, respectively. In other words, the graphitization behavior is governed by that of the mesophase spheres (MCMBs).
1.2.4.6. Glass-like carbons
Glass-like carbons (glassy carbons) are produced by the carbonization of thermosetting resins, such as phenol-formaldehyde, poly(furfuryl alcohol), and cellulose, by very slow heating. They are characterized by an amorphous structure [144]. The heating rate during pyrolysis and carbonization of the precursor has to be slower than the rate of shrinkage to compensate for the evolution of the pores caused by the release of decomposition gases from the precursor block. Their properties are similar to inorganic glasses, such as high hardness, brittle conchoidal fracture, and gas impermeability. This is why they are called glass-like (or glassy) carbon, even though they are not transparent.
Figure 1.33 Changes in x-ray diffraction parameters, d 002 , L a , and L c , with heat treatment temperature (HTT) for mesocarbon microbeads ( open symbols ) compared with coke prepared from pristine pitch ( closed symbols ) [143] : (A) prepared from quinoline solubles heated to 430°C for 120 min, and (B) from asphalt pitch heated to 430°C for 60 min.
In glass-like carbons, the BSUs do not grow with heat treatment above 3000°C. They have large d002 (about 0.344 nm) and very small Lc and La (about 3.5 nm for both), as shown in Fig. 1.9. The following experimental result shows how strongly graphitization of these carbons is depressed. When a rod of glass-like carbon was melted by directly passing an electrical current at a temperature above 3700°C under Ar gas pressure above 10 MPa, a graphite ball was obtained in a crater-like cavity at the middle of the rod, but the wall of the crater retained the characteristics of glass-like carbon, even though it had been heated to a temperature near the melting point of carbon [145]. For complete graphitization of these carbons, heat treatment under high pressure at 30 MPa at a temperature above 1600°C was needed [146,147].
1.3. Construction and purposes of the current book
The purpose of this book is to provide a basic and through understanding of the preparation, structure, properties, and applications of graphene and its related materials. The book has nine chapters.
In Chapter 1, Introduction, a short history of the term graphene
is explained and the conditions for exhibiting its intrinsic characteristics are briefly discussed on the basis of the Raman spectrum, the field effect on electrical conduction, and extremely high thermal conductivity. Fundamental knowledge about the science and engineering of carbon materials is explained, followed by a description of some graphite materials that are frequently used as raw materials of graphene, such as natural graphite, HOPG, and kish graphite, and as the reference and competitive materials such as carbon nanotubes, activated carbons, and carbon blacks.
In Chapter 2, Preparation, the procedure for preparation and conditions for CVD, cleavage (mechanical or chemical), and exfoliation through graphene oxide are summarized by focusing on structural perfection rather than the commonly used classification (i.e., top-down and bottom-up processes). The first two processes (CVD and cleavage) give less-defective layers in the resultant graphene compared with the last one (exfoliation). In the current book, flakes synthesized through CVD of organic precursors and prepared from graphite through mechanical and chemical cleavage are differentiated from those prepared from graphite through graphene oxide by oxidation, exfoliation, and reduction because they are defective, including some functional groups. In this book, therefore, the former is expressed as graphene, and the latter as reduced graphene oxide (rGO).
From Chapters 3 to 7, the properties of graphene and rGO flakes are discussed in relation to their applications: electronic properties and applications in Chapter 3, chemical properties and applications in Chapter 4, mechanical properties and applications in Chapter 5, thermal properties and applications in Chapter 6, and biomedical properties and applications in Chapter 7. The authors expect that this may provide a better and easier understanding of both properties and applications, rather than independent explanations of each property and application in separate chapters.
Chapter 3 discusses the electrical properties of graphene by referring to other carbon allotropes, carbon nanofibers, graphite, etc. Applications of graphene based on its electrical properties are summarized by dividing them into transistors, spintronic and sensor devices, photon detectors, etc.
Chapter 4 summarizes various applications of graphene based on its chemical properties, including electrochemical properties, by classifying energy storage and environment remediation, after discussing the fundamental chemical properties of graphene. The section related to energy storage includes lithium-ion batteries,
Recenzii
Recenzii
Ce cred oamenii despre Graphene
00 evaluări / 0 recenzii