Sie sind auf Seite 1von 94

Accepted Manuscript

A review on liquid-phase exfoliation for scalable production of pure graphene,


wrinkled, crumpled and functionalized graphene and challenges

Ahmad Amiri, Mohammad Naraghi, Goodarz Ahmadi, Mohammadreza


Soleymaniha, Mehdi Shanbedi

PII: S2452-2627(18)30004-7
DOI: https://doi.org/10.1016/j.flatc.2018.03.004
Reference: FLATC 61

To appear in: FlatChem

Received Date: 13 January 2018


Revised Date: 15 March 2018
Accepted Date: 29 March 2018

Please cite this article as: A. Amiri, M. Naraghi, G. Ahmadi, M. Soleymaniha, M. Shanbedi, A review on liquid-
phase exfoliation for scalable production of pure graphene, wrinkled, crumpled and functionalized graphene and
challenges, FlatChem (2018), doi: https://doi.org/10.1016/j.flatc.2018.03.004

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
A review on liquid-phase exfoliation for scalable production of pure

graphene, wrinkled, crumpled and functionalized graphene and challenges

Ahmad Amiri 1,*, Mohammad Naraghi 2, Goodarz Ahmadi 3,*, Mohammadreza


Soleymaniha 1, Mehdi Shanbedi 4

1
Department of Mechanical Engineering, Texas A&M University, College Station, TX
77843, United States of America
2
Department of Aerospace Engineering, Texas A&M University, 3409 TAMU, College
Station, TX 77843-3409, United States of America
3
Department of Mechanical and Aeronautical Engineering, Clarkson University,
Potsdam, NY 13699, United States of America
4
Department of Chemical Engineering, Faculty of Engineering, Ferdowsi University of
Mashhad, Mashhad, Iran

Correspondence and requests for materials should be addressed to A.A. (email:


ahm.amiri@gmail.com and ahmad.amiri@tamu.edu), or M.S. (email:
gahmadi@clarkson.edu)
TABLE OF CONTENTS
Graphical Abstract ................................................................................................................ 3

Abstract ................................................................................................................................. 3

1. Introduction .................................................................................................................... 4

2. Direct Ultrasonic Exfoliation in Liquid ......................................................................... 8

2.1. Organic Solvent-Based Exfoliation ....................................................................... 9

2.1.2. Exfoliation with Low Boiling Point Solvents .................................................... 17

2.2. Stabilizer-Based Exfoliation .................................................................................... 18

2.2.1. Ionic Surfactant Stabilizers ................................................................................ 18

2.2.2. Nonionic Surfactant Stabilizers ......................................................................... 21

2.2.3. Polymeric Stabilizers ......................................................................................... 23

2.2.4. Pyrene Derivative Stabilizers ............................................................................. 24

2.2.5. Aromatic Stabilizers........................................................................................... 28

2.3. Exfoliation with Ionic Solvents ............................................................................... 28

3. Mild dissolution exfoliation ......................................................................................... 29

4. Shear Exfoliation ......................................................................................................... 30

4.1. High shear mixing .................................................................................................... 30

4.2. Wet ball milling ....................................................................................................... 33

5. Microfluidization, homogenization and shear-assisted supercritical techniques ......... 36

6. Electrochemical Exfoliation......................................................................................... 41

7. Functionalization-assisted Exfoliation ......................................................................... 51

7.1. Covalent Attachment of Organic Functionalities ................................................ 57

7.2. Noncovalent Functionalization of Graphene sheets ............................................ 68

8. Future plan for mass production of high quality graphene from GO .......................... 77

9. Conclusions and outlook .................................................................................................. 79

References ................................................................................................................................ 83
Graphical Abstract

Abstract

Graphene, the 2D form of carbon-based material existing as a single layer of atoms

arranged in a honeycomb lattice has set the science and technology sectors alight with interest

in the last decade in view of its astounding electrical and thermal properties, combined with

its mechanical stiffness, strength and elasticity. Mass production of graphene is prerequisite for

its viability and wide applications. The liquid-phase exfoliation of graphite into graphene is one

of the most promising ways to achieve large-scale production at an extremely low cost. Another

3
related main challenge is to find an economical procedure for large scale production of

functionalized single- and few-layered graphene sheets. This review focuses on discussing

different liquid-phase exfoliation methods based on a common mechanical mechanism. It is our

premise that a deep understanding of the exfoliation mechanism can provide fruitful information

on how to efficiently achieve high-quality graphene by optimizing exfoliation techniques. We

highlight the recent progress on liquid-phase exfoliation for graphene production during the last

decade. The emphasis is set on the widely used direct ultrasonic exfoliation methods, stabilizer-

based exfoliation procedures, the newly explored wet-ball milling methods, shear exfoliation

approaches, functionalization-assisted exfoliation techniques, electrochemical exfoliation

routes, and the innovative supercritical fluid approaches. Various synthetic approaches are

categorized and the advantages and disadvantages of different liquid-phase exfoliation

methods are described. In addition, the main concerns regarding the quality and structure of

the graphene sheet produced from aforementioned methods and the importance of a

comprehensive evaluation of the final bulk graphene materials will also be discussed. We

hope this review will point towards a rational direction for the scalable production of graphene.

Keywords: Graphene; Liquid-phase Exfoliation; Shear exfoliation; electrochemical

exfoliation; sonication; Functionalization

1. Introduction

After the discovery of carbon-based nanostructures, two-dimensional (2D) graphene has

attracted intensive research attention because of its unique properties [1]. Mono layer

graphene with remarkable properties such as high thermal conductivity of the order of 5000

W m-1K-1, large specific surface area of 2630 m2g-1, high intrinsic mobility of 2× 105 cm2

V−1 s−1, Young’s modulus of as high as 1.0 TPa, strength of as high as 130 GPa, and

4
electrical conductivity which can surpass silver (in a defect free or nearly defect free

graphene) at room temperature owed to delocalized electrons [2-12]. Owning to its unique

properties such as low flake resistance and good transmittance, graphene was applied in

different electronic devices including the organic light emitting diodes and touch screen

displays. Note that the outstanding mechanical and chemical stability of 2D graphene make it

unique and exclusive to rigid indium tin oxide (ITO).

In addition, possessing comprehensive absorption ranging from ultraviolet to infrared

and ultrafast response, graphene could be used in different photodetectors [13] and various

optical modulators [14]. Graphene has also illustrated superior potential for fabricating highly

effective storage devices, such as solar cells, supercapacitors, and lithium-ion batteries [15,

16].

Due to numerous applications and fields which can directly benefit from large scale

utilization of graphene, there is a need to develop a commercially viable method for mass

production of high quality graphene[17-22]. To date, some methods have been reported

which hold some promise for mass production of mono layer graphene, such as chemical

vapor deposition (CVD), chemical exfoliation, exfoliation by micromechanical cleavage, and

liquid-phase ultrasonic exfoliation. Among these methods, the CVD has been used

extensively for production of graphene. In order to synthesize mono- and few-layered

graphene sheets with CVD method, precursors react on transition metal substrates at high

temperature. The CVD is able to synthesize mono- or few-layered graphene with superior

quality; however, the approach requires manufacturing conditions, which increase the costs

and may raise safety issues, such as high vacuum and high temperature. Note that the CVD

method is also size-limited caused by hydrogen, which is the major problem of this method.

In addition, in CVD the transfer procedure from the surface of metal to target substrates may

induce some defects, lowering the quality of the graphene and associated performance

5
metrics. To increase the performance of the CVD synthesis and the transfer procedures, many

studies have been carried out [23, 24]; however, further discussions of the details of these

procedures are outside the scope of present paper. Interested readers may refer to the

following reading materials on these subjects [25].

Another approach to obtain graphene is the exfoliation of graphite. The bulk graphite as

a layered material possesses strong in-plane covalent bonds and weak out-of-plane van der

Waals (vdW) interactions. The weak vdW interactions can thus provide a means to exfoliate

bulk graphite into thin graphene flakes with thickness of nanometer. This approach utilizes

very cheap raw material, and as such may be very promising for industrial scale-up. Graphite

exfoliation has been achieved via micromechanical cleavage [26, 27]. However, the

micromechanical cleavage has shown low yield of few-layered graphene. Alternatively and

as a low-cost and highly scalable method, chemical exfoliation was implemented to

synthesize mono- or few-layered graphene, in some cases with exfoliation yield of near 100%

[28-30]. This method comprises of synthesizing graphene oxide (GO) through chemical

oxidation of graphite and subsequent exfoliation via ultrasonic, which can be followed by GO

reduction. Obviously, the presence of structural defects restricts the applications of graphene

in electronic and optical devices[6, 31-33]. Recently, Voiry et al. [33] reported a simple,

rapid method to reduce GO into pristine graphene using 1- to 2-second pulses of microwaves

to completely remove the oxygen functional groups. Their method was successful and can be

easily scalable.

Being successful and high performance make liquid-phase exfoliation method a high

potential route for mass production of graphene. More recently, other novel liquid-phase

exfoliation methods were introduced that show significant performance in forming 2D

carbon-based nanostructures [34]. Liquid-phase exfoliation refers to a group of approaches

that exfoliate bulk graphite into thin graphene (mono- or few-layered graphene) directly in

6
the liquid media. Elimination of the chemical oxidation step is an advantage of this method.

Liquid-phase exfoliation includes:

i) Various ultrasonic exfoliation techniques such as liquid phase exfoliation by

surfactants, organic solvents, ionic liquids, and salts.

ii) Electrochemical exfoliation in different liquid media.

iii) Shear exfoliation methods such as high shear mixing, wet ball milling,

microfluidization and homogenization.

iv) In situ functionalization and exfoliation

The liquid-phase exfoliation strategy has attracted considerable attention as a cost-

effective, efficient, and extraordinarily versatile method that has the potential for large-scale

production of defect-free graphene[34, 35]. Although a number of liquid-phase exfoliation

approaches for synthesis of scalable few-layered graphene have been emerging[36], there are

still difficulties with processing graphene, in particular, due to the weak solubility and the

poor colloidal stability of graphene in most common solvents[37]. This issue transcends to

graphene-reinforced composites as well, with sever performance penalties, as for example,

pristine graphene tends to agglomerate in composites. Therefore, in order to improve the

interactivity and solubility of graphene, functionalization was reported as a potential solution

in the literature[38, 39]. With appropriate functionalization, the graphene can remain

dispersed over a reasonable period of time. In view of many great advantages of liquid phase

exfoliation and in situ functionalization, and the rapid progress that is being made in the

literature, this article aims to provide a comprehensive review on the synthesis methods of

mono- and few-layered graphene. In addition, the effective functionalization strategies for

solving the lack of dispersibility of graphene in different media are also described. Herein,

liquid phase exfoliation methods are classified into four main sections e.g., direct ultrasonic

exfoliation in various solvents such as organic solvents, low boiling point solvents, ionic and

7
non-ionic solvents, etc., shear exfoliation, electrochemical exfoliation and functionalization-

assisted exfoliation. In direct ultrasonic exfoliation, presence of different types of stabilizers

and their effects on the exfoliation procedure are also considered and discussed in detail.

Shear exfoliation methods include high shear mixing methods, wet-ball milling routes and

microfluidization and homogenization techniques. In electrochemical exfoliation, the effects

of type of electrolyte, concentration and voltage are discussed. Finally, with dividing

functionalization-assisted exfoliation into two covalent and non-covalent functionalization-

assisted exfoliation methods, it is tried to cover all liquid phase exfoliation methods.

2. Direct Ultrasonic Exfoliation in Liquid

Direct ultrasonic exfoliation in liquid is referred to different approaches that synthesize

graphene via direct ultrasonication of the bulk graphite in liquid media. In this method, two

factors can critically influence the yield of exfoliation of the bulk graphite. The first factor is

the energy input to the graphite in liquid media by the bath sonication or probe sonication.

This is because the bulk graphite is effectively exfoliated upon exposure to the ultrasonic

waves. Shear forces or cavitation bubbles created from such waves can generate high amount

of energy with the collapse of bubbles in liquids that can change the main structure of bulk

graphite from layered structure to single- or few-layered graphene [40].

The second important factor is the liquid media used. There is commonly an energy

barrier in the interlayers of the graphite, and in the subsequent stabilization of nanosheets via

interfacial interactions. Different liquid media (ionic liquids, aqueous solutions of stabilizers,

and organic solvent) behave differently in decreasing the potential energy barrier. To

properly classify direct ultrasonic exfoliation, this part of the literature review is organized

into 4 sections based on the liquid media used.

8
2.1. Organic Solvent-Based Exfoliation

2.1.1. Exfoliation in Suitable Solvents

This approach includes the sonication of the bulk graphite in an organic solvent and the

subsequent purification procedure [41]. In this method, the interfacial tension between the

bulk graphite and organic solvent is the main factor influencing their interactions. As

mentioned above, an appropriate organic solvent can decrease the potential energy between

adjacent layers in the bulk graphite to overcome the vdW interactions between layers[34, 42,

43]. In addition, the solvent–nanosheets interactions could balance the inter-sheet attractive

forces to stabilize the dispersion of graphene sheets in the suspension[34, 42]. There have

been a number of investigations on the interaction between the bulk graphite and organic

solvents in pursuit of identifying the suitable solvents for efficient exfoliation, and

enhancement of the stability of concentrated dispersion of graphene sheets. These are

described in this section.

In a comprehensive study, Coleman’s et al.[34] investigated the liquid exfoliation

performance of the bulk graphite with a series of commonly-used organic solvents. They

concluded that solvents with surface tension of about 40 mJ m-2 such as N-methyl-2-

pyrrolidone (NMP) and N,N-dimethylformamide (DMF) are the best candidates for mass-

production of graphene sheets. In fact, surface tension of 40 mJ m-2 is relatively close to the

surface energy of graphene, which is 68 mJ m-2 (Fig. 1a).

Note that the Raman spectrum of large graphene flakes does not have a noticable D

band, implying the lack of structural defects after the application of ultrasonic procedure for

30 min. On the other hand, small graphene flakes show a weak D band, which is due to the

edge defect as is seen in Fig. 1b. Figs. 1c and 1d show some transmission electron

microscopy (TEM) of some graphene flakes with size of a few micrometres [34].

Interestingly, Hernandez et al.[34] reported that flakes with thickness of more than a few

9
layers are rare, indicating high performance liquid exfoliation. They concluded that graphite

was extensively exfoliated, and mainly single-layered and few-layered graphene was

produced. Analyzing a large number of TEM images and paying attention to the uniformity

of the flake edges, they found 28% of total number of flakes was monolayer graphene after

sonication process in NMP. Finally, they reported a dispersion concentration of 0.01 mg mL-
1
and a flak size range of 500 nm to 3 μm. The resulting weight fraction, however, was still

too small to meet the industrial demands of large scale graphene production. Therefore, the

search for alternative organic solvents for production commercially viable graphene has yet

to continue.

Fig. 1. a) Graphite concentration measured after centrifugation for a range of solvents plotted versus
solvent surface tension, b) Raman spectra of the graphite and few-layered graphene c) and d) TEM images of
graphene. Reprinted with permission from Ref [34].

10
In order to synthesize homogenous dispersions of graphene nanosheets, Hamilton et

al.[44] selected ortho-dichlorobenzene as a nonpolar solvent. Due to the surface tension of

36.6 mJ m-2 and via - stacking, ortho-dichlorobenzene can easily interact with graphene.

But, the resulting weight concentration was improved to only 0.03 mg mL-1, which is still

quite low. In order to enhance the exfoliation yield, Bourlinos et al. [45] suggested use of a

peculiar series of perfluorinated aromatic molecules such as pentafluorobenzonitrile

(C6F5CN), hexafluorobenzene (C6F6), pentafluoropyridine (C5F5N), and octafluorotoluene

(C6F5CF3). According to their results, in addition to the surface energy matching of the

solvent and graphene, the donor–acceptor interactions play an important role as a driving

force for the exfoliation of the bulk graphite. Overall, they reported that the

pentafluorobenzonitrile (C6F5CN) had the most promising results, showing the graphene

sheets with thickness of 0.5–2 nm reaching a concentration of 0.1 mg mL-1. The sequence of

exfoliation yield was C6F5CN > C6F6 > C5F5N > C6F5CF3. Also, a comparison between

above-mentioned solvents and the analogous hydrocarbon solvents such as benzene, and

toluene showed that similar solvents failed to produce stable suspension at high

concentrations.

In addition to the selection of appropriate solvents, other parameters including

ultrasonication power and time were also studied to enhance the exfoliation performance. For

instance, with a low-power ultrasonication in NMP but using a long time period (up to 460

h), Khan et al.[46] reported increasing the graphene concentrations up to 1.2 mg mL-1. The

corresponding concentration of graphene versus sonication time is presented in Fig. 2a. It is

seen that the concentrations of flakes in solutions increases as the sonication time increases.

However, Fig. 2b shows that the flake dimensions reduce with sonication time. The

reduction of flake size is proportional to the reverse of square root of time (t-1/2).

Undoubtedly, long sonication times resulted in inducing numerous edge- and basal plane

11
defects. Later Khan et al.[47] reported a higher exfoliation performance (2 mg mL-1) by using

a sonic tip. To enhance the exfoliation yield, they also offered a novel method in which a pre-

sonication and centrifugation procedure was performed for eliminating big particles

(unexfoliated graphite). Then, the mixture was sonicated via a bath sonication for 24 h in

NMP again[47]. They reported a significant graphene concentration of 63 mg mL-1, with a

flake size of about 1µm × 0.5 µm and thickness of 3–4 layers on average. These high-

concentrated solutions will be useful for practical applications such as the manufacturing of

conducting films and reinforcement of composites [48].

Fig. 2. a) Concentration of graphene dispersion versus sonication time (hr). The inset illustrates solutions
including graphene after 6 h and 180 h. b) Mean layer number, length, and width of flakes as a function of
sonication time. Reprinted with permission from Ref [46].

Recent advances have also been made to match surface tension of a given solvent to that

of graphite as a means to assist with exfoliation. These efforts suggest that using surface

tension, rather than Hildebrand and Hansen parameters, to screen solvents for graphene

would be more appropriate [49]. Recently, Amiri et al.[2] reported an effective and

economical method for the scalable synthesis of few-layered graphene sheets by the

microwave-assisted functionalization. They could synthesize single-layered and few-layered

12
graphene sheets via dispersion and exfoliation of functionalized graphite in ethylene glycol.

They reached a high yield of exfoliation. They reported the graphene concentrations up to 1.3

mg mL-1. With reference to the SEM images of their study, all the graphite flakes expanded

after treatment. Selected area electron diffraction (SAED) results showed that almost 84 % of

entire exfoliated graphene sheets were single-layered.

According to the SEM results, all graphite flakes were expanded and exfoliated (Fig. 3a-

c). It is worth mentioning that FESEM clearly showed intense cleavage surfaces of graphite

by functionalization under microwave irradiation and sonication. Fig. 3d shows what appear

to be a graphene monolayer and a graphene bilayer, respectively. Fig. 3d is particularly

interesting as almost the middle side of the flake consists of at least two layers (white dot),

whereas on the edge side, a monolayer graphene (red dot). By analyzing a large number of

TEM images and their electron diffraction patterns based on the intensity ratio of

I{1100}/I{2110}, and considering the uniformity of the flake edges, the authors generated flake

thickness statistics as shown in Fig. 3e. The distribution of the number of layers per flakes

was obtained by analyzing 3 groups of sheets (almost 200 sheets). Fig. 3f presents typical

AFM images in which no graphene sheets were identified as being layered over other sheets,

and interestingly, all of the sheets were single-layer sheets. As shown in Fig. 3g, the resulting

samples indicated that all of the sheets had thicknesses of less than 1 nm.

13
Fig. 3. (a-c) SEM images of chemically-assisted exfoliated, highly-porous, single-layered graphene, (d)
TEM of thermally-treated graphene, (e) Histogram of the ratios of the intensity of the {1100} and {2110}
diffraction peaks for all the diffraction patterns collected, (f) AFM ichnography and (g) cross-section contour of
thermally-treated graphene. Reprinted with permission from Ref [2].

Geo et al. [50] presented a method for large-scale graphene production by ultrasound-

assisted exfoliation of natural graphite in supercritical CO2/H2O medium. They coupled a

pressure reactor with an ultrasonic generator to produce intense knocking force (Fig. 4a). The

intense impact force generated from high-pressure acoustic cavitation in a supercritical

CO2 (scCO2)/H2O system and the superior penetration ability of scCO 2 were combined to

enhance the exfoliation efficiency of natural graphite into mono- and few-layer graphene.

Under optimal conditions, the graphene yield could reach more than 50% with 93% of the

graphene sheets being less than three layers, and the suspension concentration could be

greater than 2.5 g/L. In this approach, from raw material feeding to the discharge of products,

natural graphite was directly exfoliated into high-quality graphene sheets in a few hours with

a considerably high yield and concentration by using only CO2, H2O, and ethanol. This

approach should be a feasible and promising for high-quality graphene production on a large

scale and at a low cost. According to their results, as the system pressure increased from 8

MPa to 16 MPa, the graphene yield increased significantly, reaching 51.3% at 16 MPa after

4 h reaction time (Fig. 4b). In fact, the ultrasonic energy released from acoustic cavitation at

14
16 MPa is high enough to continuously deconstruct the closely stacked graphite layers and,

thus, to continuously maintain the increased graphene yield. When ultrasonic power was

changed from 0 W to 2000 W, a sharp increase in graphene yield from almost 0% to more

than 51.3% obtained (Fig. 4c). The critical role of ultrasonic cavitation on the graphene yield

is clearly demonstrated from this result. During the production process, both making the

initial cracks on the side of the stacked graphite layers, and intercalating scCO2 molecules

into the interlayers through cracks required the assistance of ultrasonic cavitation. The

ultrasonic cavitation threshold is determined by the system pressure and was quite high in this

16 MPa system. Therefore, when the output power of the ultrasonic generator was too low to

meet the required energy demand, ultrasonic cavitation hardly occurred. Without the

assistance of ultrasonic cavitation, reliance only on the penetration ability of scCO2 was not

sufficient to overcome the vdW interactions between graphite layers.

15
Fig. 4. (a) Pilot plant for the large-scale production of graphene by ultrasound-assisted exfoliation of

natural graphite in scCO2/PVP/H2O medium. (UB: ultrasonic bath; TC: Temperature controller; PI: Pressure

indicator; PC: Pressure controller; BPR: Back pressure regulator; CV: Check valve; ①. Raw material supply

system; ②. CO2 feeding system; ③. Pre-heating/mixing system; ④. Ultrasound-coupled supercritical reactor;

⑤. Products collector; ⑥. Exhaust gas collector.). (b) Effect of system pressure on graphene yield. System

temperature: 315 K; aqueous solution fraction: 20 vol%; ultrasonic power: 2 kW. (c) Effect of ultrasonic power

on graphene yield. System pressure: 16 MPa; temperature: 315 K; aqueous solution fraction: 20 vol%.

Reprinted with permission from Ref [50].

16
2.1.2. Exfoliation with Low Boiling Point Solvents

Despite great advantages for the dispersion of graphene in the aforementioned solvents

such as NMP and DMF, there are also some disadvantages such as their high boiling point

and toxicity. In particular, these undesirable solvent properties of can hinder their

applications and deteriorate the device performance[51, 52]. For instance, O’Neill et al.[51]

showed that the suspension of graphene with a high colloidal stability and concentration can

be obtained with low boiling point solvents, such as isopropanol and chloroform. They found

that the resulting graphene flake had the lateral dimensions of 1 µm × 0.35 µm, and the

thickness of almost 10 layers on average is a suspension with the concentration of 0.5 mg

mL-1. Using a similar approach, Choi et al.[52] used a volatile solvent (propanol) for

exfoliation of graphite to graphene and reported a weight concentration of 1 mg mL-1. Quick

evaporation of propanol accelerated the deposition procedure of graphene nanosheets onto

substrates without (prior to) considerable flake aggregation. In addition, Qian et al.[53]

studied the effect of acetonitrile (ACN, boiling point of 81.6 ºC) as a low boiling point

solvent on exfoliation yield. Combined with a solvothermal procedure for producing adequate

energy, ACN molecules could diffuse into the interlayers of graphite and overcome the

potential barrier. They reported that a stable dispersion including monolayer and bilayer

graphene was synthesized after sonication and centrifugation. Their exfoliation yield was 10–

12 wt%.

In addition to exfoliation with low boiling point solvents, a solvent exchange technique

was suggested by Zhang et al.[54] for transferring graphene dispersions from high boiling

point solvents into low boiling point solvents. In this method, after preparing NMP-based

graphene solution, it was filtered and re-mixed into ethanol. This procedure shows a

considerably better stability when compared with the direct exfoliation in ethanol. However,

the reason for the stabilizing mechanism is not yet fully understood.

17
Additional investigations were also performed with methanol, toluene, and

dichloromethane, however, the graphene flakes were precipitated after centrifugation.

Catheline et al. [55] employed solvents with low boiling point and toxicity. They synthesized

some novel low boiling solvents with dissolution of KC8 in tetrahydrofuran, methyl-

tetrahydrofuran and cyclopentylmethylether. Graphite intercalated by exposing to afore-

mentioned solvents. Graphenide solutions (solutions of negatively charged graphene flakes)

were finally obtained. They reported a scalable method for production of single- and bi-layer

graphene sheets with an average lateral size of over one micron.

2.2. Stabilizer-Based Exfoliation

Based on the presented reviews in Section 1.2, it may be concluded that the direct

exfoliation in organic solvents is quick and simple, but some of the associated steps as well as

most of the materials used pose health hazards which is a concern for industrial scalability.

To enhance environmental friendlyness, another method was proposed to exfoliate bulk

graphite to graphene in liquid media in the presence of stabilizers such as surfactants,

polymers and pyrene derivatives. Stabilizers play a key role in tuning the surface tension of

the aqueous solution, which can increase the exfoliation yield efficiently. Reviews of the

related works on different kinds of stabilizer are presented in this section.

2.2.1. Ionic Surfactant Stabilizers

Sodium dodecylbenzene sulfonate (SDBS), which is an ionic surfactant was employed to

exfoliate bulk graphite to graphene by Lotya et al.[56]. They reported that the obtained

graphene flakes had dimensions in the range of 0.1 to 3 µm with the thickness distribution of

43% less than 5 layers. Fourier transform infrared (FTIR) and Raman spectra showed the

presence of a small percentage of oxidation over the exfoliation procedure. They also reached

18
the dispersion concentration of 0.1 mg mL-1. The prepared sample remained stable for over 6

weeks due to the strong electrostatic repulsion. Later, Lotya et al.[57] used sodium cholate

(SC) as an ionic surfactant to increase the dispersion concentration of graphene up to 0.3 mg

mL-1. After long time sonication of 400 h, TEM analysis showed the graphene flakes with a

lateral size of 1 µm and 400 nm and more than 20% of flakes were mono layer.

To improve the electrical conductivity of this type of films, De et al.[58] employed an

annealing stage at 500 ºC. Green and Hersam[59] also employed the density gradient

ultracentrifugation (DGU) method to separate graphene flakes based on their thickness. As a

result, in terms of sheet resistance and transmittance, the performance of graphene-based

films generated by DGU method was enhanced as compared with the common sedimentation

centrifugation. In a similar work, Hasan et al.[60] used a mild sonication for exfoliation of

the bulk graphite in aqueous solution of sodium deoxycholate (SDC) bile salt. With SDC,

they could generate denser colloidal sample including graphene flakes with better stability.

This enhancement in density was attributed to the highly-hydrophobic index of SDC, which

significantly increases the SDC absorption on the graphitic surfaces. As a stabilizer, 7,7,8,8-

tetracyanoquinodimethane (TCNQ) anion was used by Hao et al.[61] to synthesize water-

based graphene solution by using expanded graphite. Via - stacking interactions, 7,7,8,8-

tetracyanoquinodimethane could be absorbed onto the surface of graphene flakes. Also, the

negative charge imposed by 7,7,8,8-tetracyanoquinodimethane do not allow the exfoliated

graphene sheets to aggregate due to the inherent electrostatic repulsion. Finally, they reported

a successful exfoliation that produced only mono- and few-layered graphene nanosheets,

though with some defects. However, the weight concentration of obtained samples was fairly

low, of about 15–20 µg mL-1.

Generally, surfactants have both hydrophobic and hydrophilic chains (tail or head). The

hydrophobic tail groups could absorb carbon-based nanostructures (non-polar particles)

19
through vdW interactions. Complementary to that, the hydrophilic head groups are prone to

dissociation and charge the flakes. So, graphene flakes in the presence of surfactants are

stabilized in aqueous solution by the electrostatic repulsion between flakes. Zeta potential (ζ)

is typically used to measure the electric potential of aqueous dispersions.

Like anionic surfactants, cationic surfactants have been also applied as stabilizer for

exfoliation of graphite to graphene. As a cationic surfactant, cetyltrimethylammonium

bromide (CTAB) and acetic acid play essential roles in exfoliation of highly ordered

pyrolytic graphite (HOPG) to few-layered graphene flakes through rather mild

sonication[62]. According to the results of Vadukumpully et al., the hydrophobic group of

long alkyl chains can get adsorbed to the surface of graphene (hydrophobic interactions).

Also, the electrostatic charge of hydrophilic chains can prevent restacking and agglomeration.

They concluded that the obtained graphene with their method had the thickness of 1.18 nm

with 85% of which were less than four-layeres. In addition, the average length and width of

flakes were, respectively, 0.7 µm and 0.5 µm. Fig. 5 shows the topographic view of the

graphene flakes spin-coated on mica, AFM image of a graphene sheet, height profile of the

panel b, and statistical analysis of the thickness, length and width of the flakes. In order to

increase the dispersion concentration and exfoliation yield with surfactants, an ultrasonic

exfoliation procedure with the continuous addition of surfactants was proposed by

Notley[63], resulting in significant increase in exfoliation yield. With this method, the surface

tension of solutions could be kept at an optimum value of 41 mJ m-2 (comparable to the

surface energy of graphene, which is 68 mJ m-2), representing an efficient exfoliation and

significant growth in the weight concentration of samples. This approach was successful with

the use of different types of surfactants, either ionic or non-ionic.

20
Fig. 5. (A) Topographic view of the graphene layers. (B) AFM image of a single flake. (C) Height profile of the
AFM image in panel B. Statistical analysis of the AFM images of 60 flakes. (D) Thickness. (E) Length and (F)
Width of the flakes. Reprinted with permission from Ref [62].

2.2.2. Nonionic Surfactant Stabilizers

According to Guardia et al.[64], due to the stronger steric repulsion compared to the

electrostatic repulsion, non-ionic surfactants can prepare the graphene-based solution with

higher dispersibility and stability compared with the ionic ones. They prepared several

water-based graphene solutions by using direct bath sonication with various non-ionic

surfactants and compare the results with different ionic ones. Their findings presented in Fig.

6 show that the weight concentrations of graphene dispersion in the presence of non-ionic

surfactants are significantly larger than that for the ionic ones. P-123, which was the best

non-ionic surfactant, led to the weight concentration of 0.9 mg mL-1. The concentration was

21
further grown to almost 1.5 mg mL-1 after extending the sonication period to 5 h. In addition,

the paper-like films obtained from this procedure had a remarkably high electrical

conductivity of 3600 S m-1.

A successful process was suggested by Geng et al.[65] to exfoliate bulk graphite in NMP

containing nonionic porphyrin. Organic ammonium ion and porphyrin were used for

exfoliation via - interactions with graphene. The survival of undisturbed sp2 carbon

networks was the main result of this novel method. In this case the stability of non-ionic

surfactants was also caused by the steric effects. Also, the long hydrophilic chains of

surfactants can disperse into water easily. More interestingly, if the graphene flakes get close

to each other, the hydrophilic head groups of surfactants can interact to induce osmotic

repulsion.

Coleman’s group [66] investigated the effect steric repulsive potential barrier on the

dispersion of graphene using different surfactants. They studied 12 different ionic and

nonionic surfactants for identifying the main stabilization mechanism in surfactant-assisted

graphite exfoliation. They found that the dispersion concentration of graphene growss

linearly as the steric repulsive potential barrier increased.

22
Fig. 6. Concentration of graphene dispersions obtained by various types of surfactants. Reprinted with
permission from Ref [64].

2.2.3. Polymeric Stabilizers

Bulk graphite can be exfoliated in different media in the presence of various polymers

such as polyvinyl chloride (PVC) and poly (methyl methacrylate) (PMMA). For example, in

order to synthesize non-oxidized graphene, Bourlinos and his co-worker[67] sonicated the

graphite powders in water and polyvinylpyrrolidone (PVP) directly. The PVP has good

solubility in aqueous media and affinity to graphite surfaces and is biocompatible. The

resulting graphene was reported to be mostly few layered with weight concentration of 1 mg

mL-1. Steric (or depletion) effect was the main mechanism for stabilization of PVP-graphene.

23
Using ethyl cellulose as a stabilizing polymer, Liang and Hersam[68] reported a high

yield of graphene exfoliation in ethanol. They reached the maximum dispersion concentration

of 1.02 mg mL-1 via iterative solvent exchange. May et al.[69] performed a study for

identifying suitable type of polymer-solvent combinations for exfoliate of bulk graphite. They

established a simple model based on Hildebrand solubility parameters of nanosheets, polymer

and solvent. Despite the low graphene concentration on a scale of µg mL-1, the proposed

model was successful. Recently, Xu et al.[70] obtained graphene dispersions with good

concentrations using tetrahydrofuran (THF)/ hyperbranched polyethylene (HBPE). They

proposed using a solvent/polymer pair to exfoliate graphite via absorbing HBPE on graphene

flakes to generate steric stabilization for avoiding restacking. More interestingly, the

characterization instruments and morphological studies showed no defect on the structure of

graphene nanosheets. In addition, the produced graphene flakes had between 2 and 4 layers

with length of 0.5 µm and width of 0.2 µm. They reported the concentration of 3.4 mg mL-1

for graphene dispersion in chloroform after solvent evaporation.

In addition to the significant stabilization effect of polymers in the exfoliation

procedures, they can be utilized to enhance the graphene wettability. To address this issue,

Skaltsas and co-workers[71] added amphiphilic block copolymers such as poly[styrene-b-(2-

vinylpyridine)] (PS-b-P2VP) and poly(isoprene-b-acrylic acid) (PI- b -PAA) to switch

solubility of graphene from the organic to aqueous phase. A tip sonication was used in the

presence of o-dichlorobenzene or NMP, however, the quality of graphene, the concentration

of dispersion and the lateral size of flakes were small.

2.2.4. Pyrene Derivative Stabilizers

In the exfoliation of graphite, typically the higher concentration of small molecular

surfactants leads to higher concentration of graphene nanosheets in the dispersions.

24
Nevertheless, the additional surfactants or polymers which are not bonding to the exfoliated

graphene in the final product reduce the exfoliation yield and may adversely affect the

performance of composites, electronic devices and films. Moreover, removing the additional

surfactants or polymers from the exfoliated graphene is not trivial. Therefore, finding

alternative materials with low weight fraction for mass-stabilizing of graphene flakes is

critical. Pyrene (Py) molecules and some of its derivatives shown in Fig. 7 could be

promising alternative materials for high-yield exfoliation of the bulk graphite. The -

interactions between graphene and the planar surfaces of this kind of stabilizers play a key

role in the exfoliation process. Adsorption of these compounds onto the graphene surface

occurs through - interactions between the planar -conjugated surfaces, by reducing the

surface free energy of the dispersion. In these interactions both aromatic planar surfaces share

the electrons of -orbitals through a non-covalent bond [41, 72-79].

Fig. 7. Molecular structures of pyrene derivatives with their names and corresponding acronyms for stabilizing
graphene in the liquid media. Reprinted with permission from Ref [75].

In the presence of tetrasodium 1,3,6,8-pyrenetetrasulfonic acid (Py-4SO3), Dong et

al.[79] presented an efficient exfoliation technique for preparing monolayer graphene with

25
only probe sonication. Their statistical analysis of AFM images showed that the majority of

graphene sheets were mono-layered. Their AFM images are shown in Fig. 8.

Fig. 8. (a) A typical AFM image of graphene flakes dispersed in TPA/D 2O solution. (b) The thickness
distribution of all the graphene flakes observed in panel (a). (c) AFM image of a mechanically exfoliated
monolayer graphene. Reprinted with permission from Ref [79].

In another study, pyrene derivatives, 1-pyrenecarboxylic acid (PCA) was used by An et

al.[80] to produce water-based graphene dispersion. Using a method of sonication of bulk

graphite in aqueous solutions of pyrene molecules functionalized with water-soluble groups,

including Py-NH2 and Py-4SO3, Zhang et al.[73] could synthesize monolayer graphene flakes

with high quality Interestingly, after annealing the graphene nanosheets at 1000 ºC, superior

26
conductive transparent films with electrical conductivity of 181200 S m-1 and transmittance

in the visible light range of 90% were produced. Hence, in addition to the stabilizing the

exfoliated graphene solutions, the pyrene molecules may also play a role as the fillers

(nanographene) to treat the structural defects on the surface of graphene sheets.

A supercritical fluid exfoliation with 1-pyrene sulfonic acid sodium (Py-1SO3) was

suggested by Jang et al.[78] in a mixture of ethanol/water. The Py-1SO3 molecules can attach

on the surface of graphene flakes and act as electron-withdrawing groups, resulting in an

electron transfer from graphene flakes to Py-1SO3 molecules. Two main advantages were

reported with the addition of Py-1SO3 molecules. First, it could increase the exfoliation yield

of bulk graphite into monolayer and bilayer graphene sheets up to 60%. Second, the Li-ions

storage capacity was also increased. Similarly, Yang et al.[77] used Py-1SO3 for exfoliation

of bulk graphite to graphene sheets with only bath sonication in aqueous media. To enhance

the exfoliation yield, Lee et al. [74] synthesized an aromatic amphiphile molecules. These

molecules have four pyrene units as an aromatic segment and a hydrophilic dendron. With

non-destructive - stacking interactions, these molecules could easily attach on the surface

of graphene flakes and subsequently facilitate the production of water-based graphene

dispersions. They reported the concentration up to 1.5 mg mL-1 via sonication in

water/methanol media.

To investigate the effects of different pyrene derivatives on exfoliation yield, Parviz et

al.[75] used various pyrene derivatives and determined the dispersion yield of graphene

sheets. They found that Py-1SO3 showed the best exfoliation performance compared with all

other pyrene derivatives shown in Fig. 7. They could reach the graphene concentration yield

up to 0.8–1 mg mL-1 in the presence of Py-1SO3. They also showed that the functional groups

with larger electronegativity is more suitable for being adsorbed on the surface of graphene

sheets. Therefore, they concluded that pyrene derivatives including sulfonyl groups can be

27
the most effective stabilizer for producing graphene flakes over a wide range of pH. This can

be carried out sue to the stronger interactions of sulfonyl groups with the graphene surface

like SDBS or SDS. The effect of charges present corresponding to functional groups should

be effective as well. In another comparative study, Schlierf et al.[76] reported a novel

exfoliation method in aqueous media in the presence of different pyrene sulfonic acid sodium

salts such as Py-1SO3, 6,8-dihydroxy-1,3-pyrenedisulfonic acid disodium salt (Py-2SO3), 8-

hydroxypyrene-1,3,6-trisulfonic acid trisodium salt (Py-3SO3) and Py-4SO3. Their results

showed that the molecular dipole was not essential, however, it can facilitate adsorption of

the molecule on the graphene flakes, enhancing lateral displacement of water molecules

collocated between aromatic cores of the organic dye and graphene [41].

2.2.5. Aromatic Stabilizers

Similar to pyrene molecules and polymers, some polycyclic aromatic hydrocarbons such

as tetrapotassium salt of coronene tetracarboxylic acid [81] and perylenebisimide-based

bolaamphiphile detergent [82] have the ability to increase exfoliation yield of graphite. As

highly water-soluble detergents, this type of aromatic stabilizers with bulky planar aromatic

surface provides a strong interaction with the surface of graphene nanosheets via synergistic

noncovalent charge-transfer and - stacking interactions. Moreover, and similar to other

stabilizers, the negative charge imposed on the graphene surfaces prevent them from inter and

intra re-stacking, thus, leading to high level of stabilization for graphene dispersions.

2.3. Exfoliation with Ionic Solvents

Semi-organic salts with a low melting point (mostly lower than 100 oC) are commonly

used as ionic solvents [83], which have absorbed much attention due to their promising

properties including suitable thermal stability, ionic conductivity, and non-flammability.

28
According to the high dispersibility of carbon nanotubes in ionic solvents [84], it is

anticipated that ionic solvents may also be utilized for stabilizing graphene flakes via

Coulombic interaction. The exfoliation of bulk graphite into graphene nanosheets via tip

sonication in a ionic solvent, namely, 1-butyl-3-methylimidazolium bis(trifl

uoromethanesulfonyl)imide ([Bmim][Tf2N] was reported by Wang et al.[85]. They found a

rather high exfoliation yield reaching a stable graphene suspension of 0.95 mg mL-1. They

also reported that the graphene nanosheets were few-layered with micrometer sizes. More

interestingly, 1-hexyl-3-methylimidazolium hexafluorophosphate (HMIH) was found to be a

significantly strong ionic solvent for exfoliation of bulk graphite, leading to a concentration

of graphene dispersion up to 5.33 mg mL-1 via direct sonication [86, 87]. Quantitative

analysis showed that the graphene nanosheets had the width of 4 μm and an average thickness

of 2 nm, representing almost 6–7 atomic layers. Sanjeeva Rao et al.[88] described an efficient

soft processing approach for continuous synthesis of high-quality, few-layer graphene

nanosheets. They used electrochemical exfoliation of graphite in the presence of

environmentally friendly glycine-bisulfate ionic complex under ambient reaction conditions.

The resulting graphene nanosheets were mostly 2–5 layered. They suggested that the

formation of surface molecule nuclei via the polymerization of intercalated monomeric

HSO4− and SO42− ions is the main plausible electrochemical exfoliation mechanism.

3. Mild dissolution exfoliation

Valle´s et al. [89] presented a mild dissolution route without use of sonication, starting

from neutral graphite and avoiding any kind of sonication that yielded large size graphene

flakes. They demonstrated that alkali metal graphite intercalation compounds (GICs) readily

and spontaneously exfoliate in N-methylpyrrolidone (NMP), yielding stable solutions of

negatively charged graphene sheets and ribbons. With depositing sheets on a variety of

29
substrates, they produced graphenes sheets in a scalable fashion. The method was sonication-

free, and only completed with a mild dissolution of graphite in the presence of reductive

agent. To this end, they synthesized well-documented GICs, such as the ternary potassium

salt K(THF)XC24. Exposure of potassium GIC to NMP readily and spontaneously exfoliated

the intercalated graphite and yielded air-sensitive solutions of reduced graphene sheets,

affording a virtually unlimited amount of one-atom thick layers. In another study, the same

research group reported that more concentrated and chemically simpler solutions can be

obtained by mild, spontaneous, dissolution of the parent GIC KC8 in NMP [90]. Similar to

GICs-K(THF)XC24, the new solutions contained negatively charged graphene sheets and K+

ions. They reported a concentration of 0.7 mg.ml-1 and yield of 35% out of the starting

material.

4. Shear Exfoliation

4.1. High shear mixing

Shear mixing is a highly effective method to exfoliate and disperse bulk graphite and

other layered-materials that are weekly bonded along interfaces of layers in solvents[41, 91].

For this purpose, bulk graphite was pretreated with sulfuric acid to weaken the interlayer

bonding, followed by conducting shear force exfoliation. However, the thickness of obtained

sheets was high, ranging from 10 nm to 100 nm. In fact, the exfoliation performance was a

function of the intercalation procedure rather than shear force, a clear roadblock for mass

production of graphene. Recently, Paton and coworkers[35] addressed this issue. They

showed that mass production of graphene sheets without defect is possible with use of high

shear mixing of bulk graphite in proper solvents (Fig. 9). To reach high yield exfoliation,

mixers with rotating blades were utilized to form high shear rates in solvents, in which bulk

graphite were loaded (Fig 9a-d). The interaction between solvents and the bulk graphite

permits the exfoliation. The solvents can also be used for stabilizing the produced nanosheets.

30
The TEM images of graphene nanosheets are shown in Fig. 9e-h. Paton et al. [35] reported a

histogram of nanosheet thickness and the presence of monolayers was confirmed by AFM

and Raman spectroscopy, as shown in Fig. 9i. Accordingly, the exfoliation procedure

occurred at the local shear rate up to 104 S-1 as revealed in Fig. 9k. The scalability of this

approach was studied by Paton et al. [35] since it is a function of the initial graphite

concentration (Ci), mixing time (t), rotor diameter (D), liquid volume (V) and rotor speed (N).

They found that there is a linear relationship between the production rate of 2D materials and

the volume of solvents and provided initial conditions for mass production of 2D materials

(Fig. 9j). They could exfoliate huge amount of graphite with a production rate of 1.44 g h-1

for graphene, far larger than that of the ultrasonication approaches. This method could

provide defect-free nanosheets with very low thickness (less than 10 layers) and appropriate

dimensions. More interestingly, shear exfoliation can be used for exfoliation of different

multi-layered material into mono- and few-layered materials such as BN, WS2, MoSe2 and

MoTe2, as shown in Fig. 9l.

For industrial applications, preparing a large quantity of mono- and few-layered

graphene nanosheets with high quality is a must and this method can be a reasonable

approach to reach this goal. Also, the graphene thin films obtained from this method showed

a high electrical conductivity of 400 S cm-1; therefore, it can be used for solar cells

applications as well as in micro-supercapacitors.

31
Fig. 9. (a, b, c) Set-up of high shear mixer with close-up view of rotor and stator. (d), Mass-produced graphene–
NMP dispersions via shear exfoliation. (e, f, g h) TEM images of graphene nanosheets produced by shear
exfoliation. (i), Histogram of graphene thickness and Raman spectrum (inset). (k) Diagram of rotor speed, N
versus diameter, D (the red line shows a minimum shear rate of 104 S-1). (f) TEM image of exfoliated BN flake.
Reprinted with permission from Ref [35].

Shinde et al.[92] developed a shear-assisted electrochemical exfoliation procedure to

produce large graphene flakes with a high percentage of mono- and few-layered graphene.

The suggested mechanism includes intercalation of sulfate ions and then shear-assisted

exfoliation. Similarly, Tran et al.[93] reported a shear-assisted exfoliation method for mass

production of few-layered graphene. They showed a promising high-yield exfoliation of

graphite into few-layered graphene by high shear mixing of bulk graphite and solvent in the

Taylor vortex flow regime. They also showed that the obtained graphene was almost defect-

free.

Another recently emerging method uses mixer driven fluid dynamics. The device for

realizing this method is relatively simple and easily available. Based on a high-shear rotor-

stator mixer (for example, the maximum speed of 9500 rpm in one of the case), Paton et al.

[35], Liu et al. [94] and Varrla et al. [95] demonstrated a shear-assisted large-scale

exfoliation for producing dispersions of graphene flakes. Varrla et al. [95] could produce

large quantities of defect-free graphene using a kitchen blender and household detergent. They

32
investigated both graphene concentration and production rate with the mixing parameters: mixing

time, initial graphite concentration, rotor speed and liquid volume. They reported that the

production rate is invariant with mixing time and to increase strongly with mixing volume, results

which are important considerations for scale-up. They showed that the concentration of exfoliated

graphene increases linearly with time resulting in a time-independent production rate. Also, far

higher performance than sonication-based methods make this method a great alternative.

Importantly, the concentration decays very weakly as the volume was increased resulting in a

production rate that increased with volume, a property that is critical for scale-up.

4.2. Wet ball milling

Wet ball milling is one of the most effective shear exfoliation methods. Teng et al. [96]

directly exfoliated graphite in N-methyl-2-pyrrolidone (NMP) into graphene dispersion

through a combination of large and small ball milling. The proposed method of making

graphene has three advantages: (i) the wet ball-milling process can fast produce large-

volume, high-concentration, plane-defect-free, few-layer graphene dispersion at high yield;

and (ii) the filtrated NMP solvent can be reused and help to avoid environment pollution and

lower cost; (iii) due to annealing at high temperature and mechanical compression, the

physical contact between graphene sheets improved and thus facilitate electron and phonon

transport from one graphene sheet to adjacent one, resulting in higher thermal and electrical

conductivities (electrical conductivity: 2231 S cm−1, thermal conductivity: 1529 W m-1 K−1).

The route developed here would be highly useful in fabrication of highly conductive

graphene paper for practical applications. Their proposed method includes direct exfoliation

of graphite into graphene in NMP through ball milling in the presence of a large and small

zirconium oxide (ZrO2) ball. NMP was selected as a solvent because its surface energy is

close to that of graphene. From the view point of thermodynamics, the small difference of

33
surface energy would decrease their mixing enthalpy and thus enable the direct exfoliation of

graphite without the use of surfactant [96]. Ball milling was selected because it is a well-

established industrial grinding technique and can be scaled up easily. During grinding, large

ZrO2 balls fragment large graphite particles into small ones through high-speed crash, while

small ZrO2 balls detach small graphite particles into many thin graphene sheets through

shear. The exfoliation time (6 h) needed in this procedure is much shorter than ultrasonic

exfoliation time (typically several days) [46]. Moreover, the proposed wet ball-milling

technique has the potential to mass-produce graphene dispersion because it is a well-

established industrial grinding technique (Fig. 10a). The representative TEM image of

exfoliated graphene and its electron diffraction pattern are shown in Fig. 10b. The estimated

average number of layers per graphene and length were reported about 4.4 and 648 nm (Fig.

10c). The intensity ratio of the D band and G band, ID/IG, slightly increases from 0.03 for raw

graphite to 0.25 for the exfoliated graphene, indicating that a few defects were introduced

during ball milling (Fig. 10d). According to the Raman and TEM results (Fig. 10e and f),

there is a linear relationship between the IG/ID and length (L) of sheets, suggesting that the

defects introduced during ball milling are mainly located at nanosheet edges, not in the basal

plane [35]. In a word, large-volume, high-concentration, plane-defect-free, few-layer

graphene dispersion was fast produced from graphite at high yield using a wet ball-milling

method.

34
Fig. 10. a) Digital photo of centrifuged graphene/NMP dispersion with concentration 2.6 mg mL −1after ball

milling. b) Representative TEM image of exfoliated graphene, (Insert) Electron diffraction pattern of exfoliated

graphene. c) Length distribution of ball-milling exfoliated graphene as counted by TEM. d) Comparison of

Raman spectra of raw graphite and exfoliated graphene. e) Raman spectra of size-selected graphene showing

that the intensity of D peak at 1346 cm−1 increases with the increase of centrifugation force. f) ID/IG of size-

selected graphene as a function of the reciprocal of their length (1/L), showing a linear relationship. Reprinted

with permission from Ref [35].

35
Zhao et al. [97] introduced an efficient wet ball milling method to

exfoliate graphite platelets into graphenes in a liquid medium. Graphite nanosheets with a

thickness of 30 to 80 nm were dispersed into N,N-dimethylformamide (DMF) and exfoliated by

shear-force-dominated ball milling carried out in a planetary mill. After high-speed

centrifugation, irregular shaped single- and few-layer graphene sheets (≤3 layers) having a

thickness around 0.8–1.8 nm were found from the supernatant. The electrical conductivity of

the graphene powder was ∼1200 S m−1 at room temperature [97].

Similarly, Amiri et al. [98] proposed an efficient method in the presence of a group of

alcohols at very low temperature. They proposed a new method for large-scale production of

mono- and few-layered graphene via liquid phase exfoliation with the use of wet ball milling

in the presence of organic solvents at extremely low temperatures. The wet ball milling

combined with the temperature modulated high surface energy solvents affords exfoliation of

bulk graphite into graphenes in a fast, scalable, cost effective and environmentally friendly

process. Thorough analysis of as-prepared graphene flakes on a statistically significant

number of samples demonstrated that more than 61% of the flakes comprise less than 5

layers, while ∼14% of the flakes were monolayer graphene. Combined with the ∼30% yield

of few-layer graphene out of the graphite precursor, this method demonstrated incredible

efficiency in just 45 min. In the presence of methanol, the method resulted in formation of

predominantly bi-layer graphene, which is more difficult to obtain in scalable fashion, than

mono-layer graphene.

5. Microfluidization, homogenization and shear-assisted supercritical techniques

Microfluidization is a homogenization technique whereby high pressure (up to 207 MPa)

is applied to a fluid, forcing it to pass through a microchannel [99]. he key advantage over

sonication and shear-mi ing is that high > 106 s−1 is applied to the whole fluid volume not

36
just locally. Microfluidization was used for the mass-production of few layer graphene.

Karagiannidis et al. [99] reported the exfoliation of graphite in aqueous solutions under high

shear rate [∼ 108 s−1 ] turbulent flow conditions, with a 100% exfoliation yield. The material

is stabilized without centrifugation at concentrations up to 100 g/L using

carboxymethylcellulose sodium salt to formulate conductive printable inks. SEM (Fig. 11a

and b) imaging was employed to compare the lateral size of the starting flakes and of

exfoliated ones after 100 cycles. Samples after 100 cycles showed very uniform flakes in

terms of size. A statistical analysis of over 80 particles (Fig. 11c) of the starting graphite

reveals a lateral size (defined as the longest dimension) up to ∼32 μm. During the following

microfluidization, the lateral size of particles was reduced, accompanied by a narrowing of

the flake distribution. After 100 cycles, the mean flake size was reduced to ∼1 μm. Atomic

force microscopy (AFM) was also employed to determine the thickness and aspect ratio of

flakes after 20 and 100 cycles. After 20 cycles, the thickness showed a log-normal

distribution peaked at ∼ 10 nm (Fig. 11d), with a mean value of ∼ 19 nm. The distribution

shifted towards a lower thickness (Fig. 11d), with a peak at thickness of ∼ 7.4 nm, mean

thickness at ~ 12 nm after 100 cycles. It should be mentioned that the aspect ratio increased

from 41 to 59 as the processing cycle increased from 20 to 100.

37
Fig. 11. SEM image of (a) pristine graphite flakes and (b) obtained graphene after 100 cycles. (c) Histograms of

lateral flake size for the starting material and after 5, 20, and 100 cycles. (d) Flake thickness distribution and

(e) Aspect ratio after 20 and 100 cycles, as measured by AFM. Reprinted with permission from Ref [99].

Paton et al. [100] presented a method for mass production of few-layer graphene by

liquid phase exfoliation using a microfluidizer. They demonstrated that microfluidizers are

able to produce FLG directly from natural flake graphite, without the need for any pre-

treatments such as oxidation or intercalation. They reported a significant decrease in the

length and thickness of flakes with increasing the centrifuge speed. For example, the average

thickness and length of flakes falls from ~16 layers and ~2 µm to 2.1 layers and 136 nm as

the centrifuge speed increases from 1250 rpm to 16 000 rpm, respectively. his ability to

select the flake size needed for each application increases the industrial relevance of this

method, as the thinner (and so higher value) fraction can be removed from the bulk of the

thicker material. They have also highlighted the importance of post-exfoliation separation

steps in selecting the composition of flake sizes in the final product. Such separation steps are

required for most liquid phase exfoliation processes, and yet the evaluation of this process

step is often neglected in reports.

As one of the homogenization techniques, a process for mass producing graphene

nanoplatelets was reported by Cesareo et al. [101], which comprised expanding flakes of

intercalated graphite and collection of the same in a dispersing medium with forming of a

dispersion that is subjected to exfoliation and size reduction treatment carried out by high

pressure homogenization in a high shear homogenizer. A dispersion of graphene is obtained

in the form of nanoplatelets, at least 90% of which have a lateral size from 50 to 50,000 nm

and a thickness from 0.34 to 50 nm.

Recently, Li et al. [102] developed a procedure of shear-assisted supercritical

CO2 exfoliation for producing high-quality and large-quantity graphene nanosheets. A self-

38
developed apparatus which contains four parts including gas cylinder, chiller, pump and

reactor was used to exfoliate parent graphite into graphene (Fig. 12a). The mechanism of

exfoliation relies on diffusion of supercritical CO2 fluid between the interlayer of graphite

due to its high diffusivity and low viscosity. When high-speed shear stress was applied to

CO2 fluid, graphite powder was expanded and delaminated effectively into graphene sheets.

Experimental results showed that high temperature and high rotation speed of CO2 fluid

accelerate the exfoliation process. Also, it is found that the graphene synthesized by SSCE

consists of 90% exfoliated sheets with less than 10 layers and ∼70% between 5 and 8 layers,

confirmed by atomic force microscopy and transmission electron microscopy

characterizations (Fig. 12b and c). SSCE procedure provided a simple, fast, and economical

method of producing large-scale and high-quality graphene.

39
Fig. 12. (a) The schematic drawing of experiment device. (b) TEM images of graphene sheet (the inset showing

electron diffraction pattern) in low-resolution, (c) distribution of the number of graphene layer (the inset

showing AFM images of graphene sheets). Reprinted with permission from Ref [102].

Similarly, Song et al. [103] proposed a green scalable approach for producing high-

quality graphene by using a shear mixer in supercritical CO2. The total exfoliation yield

reached 63.2%, and up to 79% of the produced graphene contained less than five layers, of

which monolayer, bilayer and trilayer represented 27%, 25% and 14%, respectively. The

average lateral size of the resulting graphene was 5 μm. hey also reported that the

synergistic effects of the fluid dynamic force generated by the shear mixer coupled with the

40
supercritical fluid CO2 were the main factors contributing to the exfoliation of graphite into

pristine graphene sheets. The films fabricated by the prepared graphene had a transparency of

83% at a resistance of 0.83 kΩ/sq and maintained stable electrical and mechanical

performances even after bending through more than 500 cycles.

6. Electrochemical Exfoliation

Electrochemical exfoliation of graphite was suggested as a fast method in a two-

electrode system. Simplicity of operation procedure, high yield exfoliation under different

conditions, and the potential for mass production of mono layer graphene are advantageous

for electrochemical exfoliation [104]. The main mechanism for electrochemical exfoliation is

the intercalation of ionic species into graphite under an electrochemical bias to expand the

graphite layers, which can facilitate ultrasonic exfoliation as the second phase. Different

anionic or cationic species have been employed to accelerate the intercalation phase. To

perform anionic intercalation, bulk graphite material is installed in the setup as anode

electrode and another material like platinum (Pt) as the cathode electrode. Su et al. [105] used

a graphite anode and Pt cathode to exfoliate graphite into the graphene with different

electrolyte solutions, as shown in Fig. 13a. They tested a quick electrochemical exfoliation in

the presence of H2SO4-KOH solutions. They also studied the effects of different acids such as

HBr, HCl, HNO3 and H2SO4, and concluded that only H2SO4 is able to exfoliate bulk

graphite. Note that KOH was applied for attenuating the severe oxidation effect of sulfuric

acid to avoid introducing defects in graphene sheets. Despite the presence of defects, over

65% of exfoliated graphene sheets were very thin with a thickness less than 2 nm. Similarly,

Parvez et al.[106] investigated the effect of higher acidic solution to exfoliate graphite into

graphene, resulting in high-performance electrochemically exfoliation (Fig. 13b). To clear the

main mechanism, water oxidation under bias voltage formed some radicals (e.g. hydroxyl and

41
oxygen) which could easily react with edge sites and grain boundaries of graphite electrode.

The presence of these defects could have facilitated the intercalation of SO42- into graphite

interlayers. Therefore, the decomposition or release of gaseous SO2 as well as anion

depolarization expanded the graphite. As a result, they could introduce a high yield of

exfoliation method, representing that over 80% of graphite were exfoliated into graphene

sheets with 1–3 layers. In addition, they also investigated the effects of different inorganic

salts on the electrochemical exfoliation of graphite in aqueous solutions [107, 108]. Their

method had a yield of 85% and produced large graphene sheets with a lateral size up to 44

μm and thickness less than 4 layers. Fabrication of a highly conductive graphene film was

another advantage of this method.

42
Fig. 13. (a) Electrochemical exfoliation of bulk graphite in the presence of acids. (b) Suggested mechanism for
electrochemical exfoliation. Reprinted with permission from Ref [106].

Sulfonate salts are also suitable candidates for intercalating and subsequently exfoliation

of bulk graphite. As an efficient electrolyte, poly (sodium-4-styrenesulfonate) was used by

Wang et al.[109] to synthesize graphene sheets under a voltage of 5 V. The yield of this

method was 15% for production of mono- or few-layered graphene sheets. Wang and

coworker.[109] demonstrated that the graphene dispersion has remained stable for 6 months

without noticeable sedimentation. The proposed exfoliation mechanisms included the edge-

to-face interaction between aromatic ring of poly(sodium-4-styrenesulfonate). In a similar

43
approach, sodium dodecylbenzene sulfonate solution was selected as the electrolyte for

electrochemical production of mono- and few-layered graphene by using 30 V for 48 h[41].

Sodium dodecylbenzene sulfonate plays two important roles in electrochemically exfoliation

of graphite. It acts as both: intercalant and surfactant. Mensing and coworkers [110] proposed

a quick electrochemical approach to produce graphene-metal phthalocyanines hybrid in the

presence of copper phthalocyanine tetrasulfonic acid as electrolyte. The functionalized

graphene was obtained using electrolysis of two graphite rod-electrodes and subsequent

ultrasonication. Kuila and coworkers[111] used 6-amino-4-hydroxy-2-naphthalene-sulfonic

acid as the electrolyte and surface modifier for exfoliation of graphite anode.

Functionalization of graphene and the presence of defects in graphene nanosheets were

confirmed via different characterization instruments.

Liu et al.[112] reported an electrochemical exfoliation method in the presence of 1-octyl-

3-methyl-imidazolium hexafluorophosphate as an ionic liquid by using a two-graphite

electrode test-rig. They managed to produce the graphene sheets with average thickness of

1.1 nm and the dimensions of 700 nm × 500 nm. According to their results[112], 1-octyl-3-

methylimidazolium free radicals have enough potential to interact with -electrons of

graphene sheets, resulting an interaction between graphene sheets and ionic liquids and

subsequently facilitating the exfoliation procedure. Lu et al.[113] utilized a new type of

electrolyte, water-miscible ionic liquids such as 1-butyl-3-methylimidazolium

tetrafluoroborate [BMIm][BF4], to fabricate fluorescent carbon nanoribbons, nanoparticles,

and graphene from graphite electrodes. According to their results, water as the main impurity

in ionic liquids played a key role in performing ionic liquid-assisted electrochemical

exfoliation. Using the fabrication of novel hydrogen-bonded networks, water seems to disrupt

the internal organization of ionic liquids, causing anionic intercalation in the ionic liquid,

expansion and exfoliation of graphite into carbon nanoparticles, nanoribbons and graphene

44
sheets [113]. In addition, the ionic liquids electrolyte with water content higher than 10%

could form soluble and oxidized carbon nanostructures. Due to high potential of this method

of exfoliation, some other anionic species have also been evaluated as a means to enhance the

exfoliation of graphite anodes. For example, 9-anthracene carboxylate ion was applied as the

electrolyte to exfoliate the graphite into the graphene sheets at constant voltage of 20 V [114].

Also, graphene sheets were functionalized with 9-anthracene carboxylate ion during the

procedure, resulting in oxidation of graphite over the electrochemical exfoliation. According

to their proposed mechanism, 9-anthracene carboxylate ions were attached on the surface of

graphite anode via electrostatic interaction, subsequently exfoliated into the graphene sheets

under the electrical potential.

As another example, phosphate buffer solution was selected to perform electrochemical

exfoliation of bulk graphite via cyclic voltammetry (Zeng and coworkers [115]). They

showed that functionalization of graphite samples with oxygen in the intercalation procedure

could weaken the interlayer attraction and, therefore, facilitate the exfoliation procedure upon

the gas evolution.

According to the recent results [104], the potentials needed for intercalating of anions are

commonly higher than that of the potentials required for the oxidation of bulk graphite.

Therefore, the formation of functionalized graphene sheets with some defects is inevitable,

and this can be a harmful factor for electronic applications. In fact, after disrupting the -

electron system with functionalization, it seems impossible to reestablish the electronic

structure of graphene, even by reduction [41]. Therefore, the cathodic exfoliation was

proposed to solve the aforementioned problem. Accordingly, Zhou and coworkers [116]

proposed a method for electrochemical exfoliation of a graphite cathode at low voltage (DC,

5 V) in the presence of Na+ / DMSO as the intercalant. To produce the graphene sheets with

lower amount of functional groups (oxygen functional groups) and defects, sonication was

45
employed [116]. Similarly, Yang and coworkers [117] presented a high-performance cathodic

exfoliation of graphite in an ionic liquid namely N -butyl, methylpyrrolidinium bis(trifl

uoromethylsulfonyl)-imide (BMPTF2N). Their proposed mechanism of cathodic exfoliation

in BMPTF2N was the intercalation of ionic liquids cation, [BMP]+, under highly negatively

charge, which provide conditions for graphite expansion. They also synthesized porous

graphene sheets once KOH was used as the activation material.

Li+ can be also a promising alternative to intercalate into the bulk graphite [118, 119].

For instance, Zeng et al.[118] presented an efficient approach to generate mono-layered

MoS2, WS2, TiS2, TaS2, ZrS2 and graphene by a controllable lithiation procedure. They first

put bulk layered material such as graphite into an electrochemical test-rig as cathode, and the

Li foils were selected as anode for preparing the required Li ions. This procedure was

accomplished within 6 h at ambient temperature and the production of mono-layered

materials yielded up to 92%. Wang et al. [120] reported a high-performance method for

exfoliation of graphite (yield of 70%). To this end, the negative graphite electrode was

charged electrochemically and subsequently expanded in an electrolyte of Li salts and

propylene carbonate [120].

Shi et al. [121] introduced a non-electrified electrochemical e foliation method with a

high yield of ∼80% for large-scale production of high-quality graphene sheets. By direct

electrochemical reaction between graphite powders and metallic Li in 1 M LiPF6/PC

electrolyte, continuous graphite exfoliation was carried out (Fig. 14a). Authors claimed that

the non-electrified electrochemical exfoliation route can be easily scaled-up for mass

production of few-layer graphene. Over a lab-level experiment, they could collect 16.25 g of

pure graphene sheets out of 20 g of graphite, representing a high yield of 80% (Fig. 14b).

From HRTEM image, the layer distance of the obtained four-layer sheets is approximately

0.36 nm (Fig. 14c). To confirm the quality of samples, the oxygen content of obtained

46
graphene was measured via XPS, showing approximately 3.48%, which is slightly greater

than that of parent graphite precursor (1.29%). The received sheets had relatively low defect

density (Raman spectroscopy, ID/IG = 0.45), with a C/O ratio (XPS) of 27.74 (Fig. 14d).

Fig. 14. (a) Schematic illustration of the non-electrified electrochemical exfoliation method for large-scale

production of high-quality graphene in 1 M LiPF6/PC electrolyte. (b) HRTEM images of high-quality graphene.

(c) Elemental contents of the high-quality graphene. Reprinted with permission from Ref [121].

Tian et al. [122] introduced a one-step quick electrochemical method for preparing

water-dispersible graphene. Aided by Oxone, water-dispersible graphene of 2–5 atomic

layers was electrochemically fabricated from graphite in one step without any extra

stabilizers or additives. Oxone as a novel electrolyte and the electrochemically enhanced

oxidation were responsible for the excellent aqueous dispersibility, effective exfoliation and

the reasonably high yield (up to 60.1%). A two electrodes system was applied in the presence

of 0.05 mol L−1 aqueous Oxone solution as the electrolyte (as shown in Fig. 15a). High

47
anodic potential of 50 V was loaded on two electrodes, the graphite anode expanded with

numerous debris detaching from the anode and constant bubbles being generated. The whole

process lasted for about 4 minutes and the exfoliated graphene dissolved in the electrolyte,

resulting in a brown suspension with obvious aggregations (Fig. 15a). The mechanism of

exfoliation presented by authors is shown in Fig. 15b. In fact, the abundance of SO42−,

HSO5− and HSO4− anions were produced with dissolving of Oxone. They produced active

oxidative radicals of sulfate radical or hydroxyl radical to oxidize the graphite during its

exfoliation. Over the procedure, oxygen-contained functional groups can decorate onto the

graphene surface through destroying the sp2 conjugate carbon structures. These functional

groups could provide a high colloidal stability for produced graphene. The HR-TEM image

(Fig. 15c) presents the well-defined structure of a bi-layer graphene. The selected area

electron diffraction pattern in Fig. 15c exhibits a typical 6-fold symmetric diffraction with

strong diffraction, indicating the conserved crystallinity of graphene. Fig. 15d shows the

typical AFM image of produced graphene with 2.5 nm height, indicating four atomic layers.

They could achieve a dispersion of ∼1.04 mg mL−1 without obvious sediment during 2

months standing (Fig. 15e). The analysis of significant number of sheets (45) statistically

showed that the lateral size distribution is in the range of 1-5 μm (Fig. 15f). Also, the

statistical analysis of 31 sheets illustrated that the obtained sample has the thickness

distribution of 73.4% less than five layers (Fig. 15g).

48
Fig. 15. (a) Schematic illustration of electrochemical exfoliation setup (left) and electrochemical exfoliation

process with increase of reacting time (right). (b) Mechanism of electrochemical exfoliation of water dispersible

graphene, including intercalation, oxidation, exfoliation and solution. (c) HR-TEM image and SAED pattern of

produced graphene. (d) AFM image of graphene. (e) Photograph of 2L aqueous graphene dispersion with a

concentration of 1.04 mg mL−1. (f) Statistical lateral size analysis on SEM images. (g) Statistical thickness

analysis on AFM images. Reprinted with permission from Ref [122].

Although electrochemical exfoliation of graphite has drawn increasing attention over the

last few years as a potentially scalable method, lack of control on procedure once

electrochemical procedure has initiated and separation of big portions of graphite rods can

hamper its promotion to the market. Generally, this method involves the use of an aqueous or

non-aqueous electrolyte (Table 1) and an electrical current to drive structural expansion at a

49
graphite electrode. Mostly, a graphite rod is used as working electrodes and controlling the

procedure to begin from cap of electrodes is complex. While this method is not equipment-

intensive and is typically performed under ambient conditions and can be considered as an

eco-friendly method as compared to other chemical/sonication routes, the presence of

oxidation or reduction reactions can decrease the quality of graphene sheets, as can be seen

the ratio of C/O and ID/IG of the samples in Table 1. Most importantly, electrochemical

exfoliation procedures can only be used for producing gram scale quantities of graphene

materials at the laboratory-scale. Relying on operating voltages, graphite precursors and

electrolytes, the quality can be tunable.

Table 1. Summary of electrochemical exfoliation of graphite in different electrolytes.

Electrolyte Working bias Yield of Number/thickness C/O ID/IG Stable Production Ref

graphene of graphene layers graphene rate

concentration

Mixture of (1) +2.5V, 1 min, 5–8 wt% less than 2 nm – 0.5–1.0 0.085 mg mL−1 – [105]

H2SO4 (2) switching (65%) (473 nm) in DMF

and KOH (+10 V, 2s; -10

V, 5s)

0.1 M sulfuric + 10 V, 10 min ≈60 wt% 1–3 layers (80%) 12.3 0.40 (488 nm) 1 mg mL−1 4.2 g h−1 [106]

acid
in DMF

Glycine-bisulfate (1) +1V, 5 min; – < 3 layers (≈60%) 8.8 0.70 (633 nm) 2.2 mg mL−1 up to 0.18 g [88]
−1
solution (2) +3V, 5 min in NMP h
−1
Mixture of ±20V, 10 min – < 3 layers (≈80%) 26.2 <0.45 (532 0.2 mg mL in 1.54 g h−1 [123]

melamine nm) DMF

and sulfuric acid

0.1 M + 10V, 10 min ≈75 wt% 1–3 layers (85%) 17.2 0.25 (532 nm) 2.5 mg mL−1 in 16.3 g h−1 [107]

sodium/potassiu DMF

m/ammonium

sulfate

0.1 M SDS (1) +1.6V, 12 h; – – – 0.68 (532 nm) – – [124]

(Sodium dodecyl (2) -1V, 2 h

sulfate)

0.01 M SDS + (1) +5V, 30 min; – bilayers (54%) 6.7 (400 °C 0.61 (532 nm) – – [125]

0.15 M (2) +6V, 1 h annealed)

50
Na2SO4

0.1 M H2SO4 10V, 60 min; 70% Few layers 16.67 0.11 (514 nm) – 0.5 g h-1 [126]

Mixture of (1) +1V, 10 min; – 3–6 layers (95%) 17.2 0.67 (633 nm) – – [127]
(2) +3V, 10 min
NaOH

and H2O2

0.1 M (1) From – 3V to 30–40% – – >1(functionali 0.1 mg mL−1 – [128]

LiClO4+1.80 –5V, 5 min; (2) - (functionali zed graphene) in DMF

mM TBAP in 60 5V, 24 h zed (532 nm) (functionalized

mL propylene graphene) graphene)

carbonate

Mixture of LiCl, Cyclic – 6–10 layers – 0.3 (mainly) – 0.5–2 g h−1 [129]

Et3NHCl in Voltammetry (mainly) (633 nm)

DMSO Scan from 0 to -6

0.01 M AgClO4 Cyclic – few layer (2 nm – – 0.01 mg mL−1 – [130]

and 0.1 M TBAP Voltammetry thickness)


in NMP
in NMP Scan from 0 to -6

NaClO4 in from 0 to + 5 V; 61% single or double – 0.79 (after 0.24 mg mL−1 – [131]

acetonitrile 30 min layer (69%) reduction) in DMF

(633 nm)

7. Functionalization-assisted Exfoliation

Functionalization-assisted exfoliation is a high-yield method for the exfoliation of bulk

graphite into mono layer graphene via direct sonication in different solvents. Covalently

oxidation-assisted exfoliation is the most common method of the functionalization-assisted

exfoliation. Graphite oxide as the first resulting material can be exfoliated in different media

such as water by direct sonication. The presence of polar functional groups which can interact

with an organic solvent, and the increased d-spacing of the material which lowers the vdW

interactions between graphite layers, were suggested as the main reasons for increasing the

exfoliation yield [31, 132]. Using functionalization of graphite oxide with isocyanate,

Stankovich et al.[133] performed the first graphite oxide exfoliation in organic solvents. Such

reactions decrease the percentage of hydrogen bonding in the oxide layers by changing

hydrogen bond donors, e.g. hydroxyl groups, into carbamate esters and/or amides. They

51
reached the stable monolayer graphene dispersions of 1 mg mL-
1
in dimethylformamide (DMF) and N-methyl-2-pyrrolidone (NMP). In a similar study [134],

poly(tert-butyl acrylate)-functionalized graphite oxide were employed to prepare a suitable

dispersion of graphene in toluene at 1 mg mL-1 [134]. Recently, exfoliation of GO in certain

organic solvents such as DMF, NMP, tetrahydrofuran (THF), and ethylene glycol (EG) has

also been performed with direct sonication [135]. Paredes et al. reported synthesis of stable

dispersions of 1 mg mL−1 in the presence of organic solvents. In addition to all the

advantages of graphite oxide, the oxidation of bulk graphite produces different reactive

oxygen-containing functional groups on the basal planes and edges [136, 137]. Oxygen-

containing functional groups such as hydroxyl and epoxy, carboxyl and ketone groups can

impart hydrophilicity to the graphene network, resulting in stable dispersion in water via

direct sonication (Fig. 16) [138, 139].

Fig. 16. Scheme for the production of water-based graphene oxide suspension from graphite (a) a photo of bulk
graphite powder, (b) a SEM image of the layer structure of graphite oxide, (c) a water-based graphene oxide
suspension, (d) a AFM image of graphene oxide placed on a mica substrate, and (e) chemical structure of
graphene oxide with different functional groups. Reprinted with permission from Ref [139].

52
Graphene oxide is heavily functionalized with many permanent chemical defects, such as

holes introduced into the basal plane. These holes are not readily healed even upon annealing

[140, 141]. Some efforts have been done to remove holes. For example, reduced graphene

oxide (RGO) were produced via the reduction of an aqueous graphene oxide suspension with

hydrazine (NH2NH2) or dimethylhydrazine (C2H8N2) as a pure material [132, 142]. Due to

the promising electrical and thermal properties as well as high surface area, RGO material

has shown great potential in a wide range of applications. Nevertheless, preparing stable and

homogeneous suspensions of RGO platelets has been challenging. In addition, the chemical

reduction of the materials with hydrazine led to agglomerated powders, which created

another problem. To address these problems, a mixture of water and organic solvents (9 parts

organic solvent and 1 part water) were employed [143] and subsequently was reduced by

hydrazine, leading a stable suspension of RGO for depositing films with a relatively high

electrical conductivity (Fig. 17).

53
Fig. 17. The procedure of preparing homogeneous colloidal suspensions of RGO. Reprinted with permission
from Ref [143].

The production of chemically converted graphene from the reduction of graphene oxide

is a convenient method for obtaining large quantities of graphene[31]; however, even with

efficient reducing agents such as hydrazine or H2, and annealing at high temperature, the

crystalline structure of graphene is not restored. Graphene oxide is heavily functionalized

with many permanent defects, such as holes, introduced into the basal plane. These holes are

not readily healed even with annealing [140, 141]. While thermal annealing works well in

healing some of holes, it does not seem to be cost-effective [38].

54
As mentioned in introduction, a novel technique for overcoming above-mentioning

drawbacks for production of single-layered and/or few-layered sheets of graphene is the

liquid phase exfoliation of graphite in the presence of high surface-tension organic solvents

and direct sonication [38]. However, due to the lack of easily-miscible functional groups,

such as polymers to decrease interlayer attractions, the graphene sheets suspended in the

high-surface tension base fluids tend to aggregate and re-stack [56]. Also, due to the strong π–

π interactions, liquid-phase exfoliated graphene cannot reach stable dispersion in base fluids,

leading high level of aggregation[38]. However, due to poor graphene-solubility, the

exfoliation performances of most of the solvents suggested in previous attempts were quite

low. Also, in some cases, the obtained samples are only partially exfoliated and still contain

extensive amount of staked graphitic layers. To increase the efficiency of exfoliation with this

method, the chemical functionalization of graphite with different functional groups, such as

4-bromophenyl, provides a new approach for improving the solubility of graphene sheets in

polar, aprotic, organic solvents, allowing the exfoliation of bulk graphite with fewer problems

[28].

Earlier, Sun et al.[38] concluded that graphite can be changed to graphene by covalent

functionalization of the graphite and the synthesis of a stable suspension in the presence of

DMF without any added surfactant or stabilizer. They achieved both exfoliation and

functionalization with a faster procedure than those of shear- or stabilizer-based exfoliation

methods and obtained appropriate edge-functionalization and intact pristine graphene

structure in the interior basal planes. Moreover, they directly synthesized functionalized

graphene with good colloidal stability in organic solvents. They concluded that the edges of

the expanded graphite are more accessible than the interior basal plane surfaces that are

stacked with strong π–π interactions. Due to higher solubility of produced graphene with this

method than pure graphene, exfoliation yield increased as compared to other methods when

55
parent graphite is the raw material. In addition, the morphological results showed that more

than 70% of the graphene flakes had layers with thicknesses less than 5 nm, providing a high

yield of soluble graphene. They concluded that the proposed preparation approach is

significantly simpler than the multi-step approaches needed for preparing graphene oxide

followed by reduction to chemically-converted graphene, and it is an effective method to

produce soluble graphene with high yield (Fig. 18).

Fig. 18. (a) Histogram of number of layers per graphene flakes, (b) SEM image of graphene flake, (c & d) AFM
of graphene on mica and (e and f) the height profiles along the black imaged lines above in panels c and d,
respectively. Reprinted with permission from Ref [38].

56
Amiri et al. [3] also proposed a chemically-assisted exfoliated graphene method for

production of highly-crumpled, few-layered graphene (HCG) and highly-crumpled nitrogen-

doped graphene (HCNDG) with high surface area. These materials were produced by

microwave-assisted liquid-phase exfoliation of bulk graphite. The mechanism for the

exfoliation of the graphite to HCG and HCNDG includes the generation of semi-stable

diazonium ions, which then initiate a radical reaction with the flakes. Then, treated graphite

with higher dispersibility in organic solvents such as DMF opens a new gateway for

exfoliation of expanded graphite into graphene. This stabilizer-free solution of

graphene/DMF may be useful when impurities are the source of new problems.

The covalent functionalization of graphene sheets is carried out by direct bonding

between functional groups and sp2-hybridized carbon or initiation of covalent bonds between

oxygen-containing groups of GO and organic functional groups. Also, non-covalent

functionalization is another method to exfoliate graphene. In addition to section 2.2, some of

non-covalent functionalization methods are reviewed in detail in section 7.2.

7.1. Covalent Attachment of Organic Functionalities

According to the recent studies on different carbon nanostructures [144-151], organic

free radicals can be classified as the most promising organic groups for the reaction with

sp2 carbons of graphene. Sinitskii et al133. and Kosynkin et al. 134


prepared a reactive free

radical by heating of a diazonium salt to attack the sp2 carbon atoms of graphene, resulting in

nitrophenyls-functionalized graphene. Upon chemically-unzipping of carbon nanotubes

(CNT), the graphene sheets were generated. Similarly, Niyogi et al.[145] proposed the

covalent functionalization of graphene sheets with nitrophenyls. X-ray photoelectron

spectroscopy (XPS) was used to analyze the covalent attachment of the nitrobenzyl groups on

the surface of graphene (Fig. 19). There are two peaks at 406 and 400 eV in the N1s XPS

57
spectrum of the nitrobenzyl-functionalized graphene, representing the presence of the

nitrogen of NO2 and reduced nitrogen, respectively. Note that the reactions with diazonium

salts have been employed to functionalize different carbon-based nanostructures [28, 144,

146, 152, 153].

Fig.19. N1s and C1s XPS spectra of the functionalized GNR and GNRs. Reprinted with permission from Ref
[145].

Fang et al. [154] reported a promising approach to functionalize graphene nanosheets via

a diazonium addition. They functionalized individual and multi-layer graphene sheets (size

range: 20–40 nm) with linear polystyrene (Fig. 20). According to their results and those of

Ruoff and coworkers [155, 156], as the size of graphene sheets decreases, the interlayer

cohesive energy reduces, implying lower effective viscosity of suspension.

58
Fig. 20. The functionalization procedure of graphene nanosheets by polystyrene. Reprinted with permission
from Ref [154].

The ratio of sp2 hybridized carbon to sp3 ones (ID/IG) in the graphitic lattice is helpful to

indicate the degree of functionalization. Raman spectroscopy is also a powerful instrument to

measure the ID/IG ratio. It should be noted that pure graphene is a 2D sp2 hybridized carbon

sheet and the presence of sp3 carbons in the lattice points to the presence of defects. Note that

such defects may be present in the basal edges and/or the plane. Fig. 21 panel (a) shows that

the ID/IG ratio of samples increases from 1.7 to 2 after diazonium addition reaction [157]. The

functionalization of graphene sheets with phenyl groups is confirmed by the presence of a

sharp D band at 1343 cm–1 seen in Fig. 21 (b), which is in good agreement with the presence

of defects on the graphene surface.

59
Fig. 21. Radical addition reaction mechanism. (a) Raman spectra of a mono layer graphene after the and before
functionalization and (b) optical image of a functionalized mono layer graphene with some holes or defects.
Reprinted with permission from Ref [157].

Recently, Sarsam et al.[158] proposed a new and quick synthesis method for

synthesizing triethanolamine-treated graphene nanoplatelets. Using an electrophilic addition

reaction, they could functionalize graphene nanoplatelets via molecules including hydroxyl

(OH) groups e.g. triethanolamine. They could synthesized water-based graphene

nanoplatelets nanofluids with high stability [158]. The triethanolamine-functionalized

graphene nanoplatelets with different weight concentrations and specific surface areas

showed appropriate stability in aqueous media. The direct coupling of graphene nanoplatelets

with free radical of triethanolamine was confirmed by several characterization methods as

shown in Fig. 22. The higher ID/IG ratio of triethanolamine-treated graphene nanoplatelets

compared to that of the pristine sample, the appearance of some meaningful bands in FTIR

spectrum and the presence of different elements in energy dispersive X-ray spectroscopy

(EDS) results confirms the successful attachment of functional groups on graphene

nanoplatelets.

60
Fig. 22. (a) Schematic of Electrophilic addition reaction. (b) FTIR spectra (c) Raman spectra, and (d) energy
dispersive X-ray spectroscopy traces of the triethanolamine-treated graphene nanoplatelets (EDS traces are
presented for different specific surface areas). Reprinted with permission from Ref [158].

Similar to free radicals, dienophiles also can easily decorate on the graphene surface with

quick reaction with sp2 carbons. As one of the dienophiles, Azomethine ylide has been

commonly used to functionalize different carbon-based nanostructures. In fact, Azomethine

ylide reacts mostly through a 1,3 dipolar cycloaddition [159]. In fact, 1,3 dipolar

cycloaddition reaction can be employed for a number of organic derivatives and plays a key

role in increasing the performance of different applications such as nanoelectronic devices,

polymer composites, drug delivery, solar cells and biotechnology [160-162]. Georgakilas and

coworkers[163] produced graphene sheets via direct sonication of pristine graphite in organic

61
solvents, and subsequently performed the substitution of graphene sheets with pyrrolidine

rings by a 1,3 dipolar cycloaddition of azomethine ylide [164].

The dihydroxyl phenyl-functionalized graphene sheets were generated perpendicular to

the graphene surface after the attachment of azomethine ylide precursors, which is shown in

Fig. 23 [164]. Due to the introduction of hydroxyl groups onto the graphene sheets, their

dispersibility was significantly enhanced in polar solvents (e.g. ethanol). The significant

increase of the ratio of ID/IG verified the appearance of sp3 planar carbon atoms, implying

successful functionalization (Fig. 23).

Fig. 23. The 1,3-dipolar cycloaddition reaction for decorating the surface of graphene with azomethine ylide
(left). Ethanol-based Functionalized graphene nanoplatelets dispersion (top, right). Raman spectra of (a)
pristine graphene and (b) pyrrolidine functionalized graphene (bottom, right). Reprinted with permission from
Ref [164].

62
Park and Ruoff [156] reported the preparation of large few-layered graphene sheets with

sizes ranging between 500 nm and 1 μm; their morphological studies using TEM and AFM

are shown in Fig. 24. Their AFM images illustrate a mono layered graphene with a

thickness of 1.5 nm. According to the resent studies [156, 165], monolayer graphene without

any modification has a thickness between 0.6 and 0.9 nm; therefore, the increase in graphene

layer thickness was attributed to the functional groups across the surface of graphene sheets.

Fig. 24. (a, b, c) TEM images and (d, e) AFM images of the functionalized graphene sheets. Reprinted with
permission from Ref [156].

Dipolar cycloaddition can be employed for several aldehydes to yield the degree of

functionalization [159]. Due to this flexibility, this reaction procedure has attracted numerous

researchers in this field of study. Zhang et al.[166] utilized tetraphenylporphyrin aldehyde

and sarcosin as precursors to attach tetraphenylporphyrin (TPP) on the surface of graphene

sheets. Also, Quintana and coworkers[167] used paraformaldehyde as precursor to

functionalize graphene flakes with 1,3-dipolar cycloaddition reaction and NH2-terminated α-

amino acid was obtained. To selectively bind gold nanorods, the decorated amino groups

were applied.

63
Upon the creation of the reactive intermediate nitrene, phenyl and alkyl azides can easily

attach to the graphene surface. Therefore, to produce different organic graphene derivatives,

some researchers used the effective reaction of organic azides with graphene sheets. For

example, Liu and coworkers[168] employed the aforementioned method to decorate different

functionalities on graphene sheets via a three-membered aziridine ring. They used this

method for functionalization of graphene sheets with perfluorophenylazides (PFPA) with

thermal and photochemical methods. Fig. 25 shows the functionalization procedure of

pristine graphite with perfluorophenylazides. Solubility of the PFPA-treated graphene sheets

was significantly higher than that of pristine samples.

Fig. 25. Edge-functionalization of pristine graphene with perfluorophenylazides. Reprinted with permission
from Ref [168].

64
Nitrene addition reaction is a suitable method for functionalization of graphene sheets.

Strom et al. [169] presented a high performance approach to functionalize graphene sheets

via nitrene addition of azido-phenylalanine [Phe(N3)]. This procedure was used to exfoliate

bulk graphite into few layer graphene. Moreover, nitrene addition of azido-

phenylalanine provided a direct method to reach high-degree of functionalization in graphene

sheets. They could introduce a high degree of functionalization, representing 1

phenylalanine substituent per 13 carbons.

Vadukumpully et al. [170] also reported a quick and efficient method for covalent

functionalization of graphene sheets. They functionalized dispersible surfactant wrapped

graphene sheets (CTAB-graphene) with different alkylazides. Applied alkyl chains had

several groups such as hexyl, hydroxyl-undecanyl, dodecyl, and carboxy-undecanyl [170].

According to their results 11-azidoundecanoic acid is a highly suitable type of azide for

improving dispersion and the alkylazides-functionalized samples illustrated high

dispersibility in organic solvents like toluene and acetone. The functionalization of graphene

sheets with various alkylazides is shown in Fig. 26 panel (a). The corresponding Raman

results is presented in Fig. 26 panel (b), showing an increase in the ID/IG ratio after

functionalization with nitrene addition method [170]. Consistent with Raman spectrums, the

amount of nitrene play a key role in increasing the degree of functionalization.

65
Fig. 26. (a) Functionalization of graphene sheets with alkyl nitrenes. (b) Raman spectra of graphene sheet
before (1) and after functionalization of graphene sheets with 11-azidoundecanoic acid in 1:1 (2) and 10:1 (3)
ratios. Reprinted with permission from Ref [170].

He and Gao[171] introduced an efficient and useful nitrene cycloaddition reaction for

functionalization of graphene sheets with polymeric chains and different functional groups.

Interestingly, this functionalization method can be applied to covalently bond different

functional moieties and polymers to the graphene sheets, representing functionalized 2-D

graphene sheets and 2-D macromolecular brushes, respectively. As a result, the treated

samples demonstrated higher chemical and thermal stabilities compared to that of graphene

oxide. The modified graphene sheets showed high electrically conductivity, great stability

and dispersion in solvents such as DMF, NMP, DMSO, THF, and acetone [171]. Fig. 27

illustrates the photographs of functionalized graphene nanosheets dispersion in various

solvents after sonication. A higher dispersibility of graphene sheets in various solvents are

obvious after covalently functionalization.

66
Fig. 27. The photographs of functionalized graphene nanosheets dispersion in solvents after sonication:
hydroxyl-functionalized graphene in water (1) and DMF (2), carboxyl-functionalized graphene in water (3) and
DMF (4), poly(ethylene glycol)-functionalized graphene in water (5) and DMF (6), long alkyl chain-
functionalized graphene in chloroform (7) and toluene (8), polystyrene-functionalized graphene in chloroform
(9) and toluene (10), amino-functionalized graphene in water/chloroform (12), graphene oxide in chloroform
(13), and reduced graphene oxide in DMF (14). Reprinted with permission from Ref [171].

Similarly, Choi et al.[172] proposed a method for covalent functionalization of graphene

sheets with thermally generated nitrene. Upon selecting the amount of nitrene in the

functionalization procedure, the band gap of the functionalized graphene sheets was

controlled. Zhong et al.[173] also proposed a facile method to produce aryne-functionalized

graphene sheets via aryne cycloaddition. Using 2-(trimethylsilyl)aryl triate as a precursor

along with reactive benzyne intermediate were employed to do aryne cycloaddition. The

aryne-functionalized graphene sheets showed a significant stability in different solvents such

as DMF, o-DCB, ethanol, chloroform, and water. The functionalization procedure resulted in

the attachment of a four-membered ring, decorating graphene surface with the aromatic arene

rings (Fig. 28). Additionally, different interesting derivatives have been synthesized by

substituting organic groups with the aryl rings. Fig. 28 is also shows stable and uniform

dispersion of functionalized graphene sheets in DMF [173].

67
Fig. 28. (top) The schematic of chemical functionalization of graphene sheets via aryne cycloaddition. (Bottom
left) Photographs of graphene sheets and aryne-modified graphene sheets dispersions in DMF: (a) graphene
sheets, (b) F- aryne-modified graphene sheets, (c) Me- aryne-modified graphene sheets and (d) H- aryne-
modified graphene sheets. (Bottom right) (B) UV–vis absorption spectra of Me- aryne-modified graphene sheets
dispersed in ethanol with different concentrations. Reprinted with permission from Ref [173].

7.2.Noncovalent Functionalization of Graphene sheets

Upon manipulating vdW interaction and/or exposing ionic interaction between a

molecule and graphene sheets, noncovalent functionalization of graphene is occurred.

Jaegfeldt et al.[174] reported that the pyrene moiety have a severe affinity toward the basal

plane of graphite via π-stacking. Similarly, a stable water-based graphene nanoplatelets

dispersion was synthesized using 1-pyrenebutyrate as one of the water-soluble pyrene

derivatives [175]. To make the stable water-based graphene nanoplatelets dispersions,

graphene oxide was initially reacted with pyrenebutyric acid and subsequently the

pyrenebutyric acid-treated graphene oxide was reduced chemically in the presence of

hydrazine [175]. They finally produced graphene oxide and pyrenebutyric-treated graphene

nanoplatelets (Fig. 29) with thickness of 1.3 and 1.7 nm, respectively.

68
Fig. 29. (left) AFM images of graphene oxide and (right) pyrenebutyric-treated graphene nanoplatelets on mica.
Reprinted with permission from Ref [175].

In another study, Wang et al.[176] non-covalently functionalized the graphene sheets

with pyrene butanoic acid succidymidyl ester to enhance the power conversion performance

of graphene. An et al.[177] also prepared 1-pyrenecarboxylic acid-treated graphene films

with noncovalent (π-stacked) pyrenecarboxylic acid with graphene sheets to prepare films

with excellent optical and molecular sensing properties.

Yang et al.[178] synthesized stable water-based graphene nanoplatelets suspensions with

a high volume fractions (0.6–2 mg/mL) using the reduction of graphite oxide. They used

sodium lignosulfonate, sodium carboxymethyl cellulose, and pyrene-containing

hydroxypropyl cellulose (HPC-Py) to reduce graphite oxide chemically. Kodali et al. [179]

proposed a method for noncovalent functionalization of graphene sheets via pyrenebutanoic

acid-succinimidyl ester, resulting in production of flakes without disrupting structure. In a

majority of noncovalent functional groups, the attachment of the substituted molecules and

graphene was attributed to the vdW interactions [180]. Recently, Su et al.[181] offered a

69
promising method for functionalization of graphene sheets with pyrene and perylenediimide.

They could synthesize a stable aqueous dispersion including mostly mono and double layer

graphene sheets.

As another type of noncovalent functional groups, DNA can be employed as stabilizer

for developing the stability of graphene in aqueous media and other basefluids. For instance,

Liang et al.[182] non-covalently functionalized graphene sheets with single-stranded DNA.

To this end, they first oxidized graphite to graphene oxide and subsequently performed the

hydrazine reduction in the presence of single-stranded DNA [182]. The produced graphene

nanosheets presented a high concentration in aqueous media, ranging as high as 2.5 mg/L.

Using a similar approach, in order to deposit gold nanoparticles on graphene sheets, Liu

et al.[183] produced thiolated DNA oligonucleotides-functionalized graphene nanoplatelets.

This procedure included adsorbing thiolated DNA oligonucleotides on graphene oxide

nanoplatelets and reducing thiolated DNA oligonucleotides-functionalized graphene oxide

via hydrazine to thiolated DNA oligonucleotides-functionalized reduced graphene oxide

[183]. The resulting functionalized materials had a unique solubility in aqueous media.

Subsequently, gold nanoparticles were poured into the aqueous solutions of thiolated DNA

oligonucleotides-functionalized reduced graphene oxide to form a homogenous suspension.

The reported mechanism is shown in Fig. 30.

70
Fig. 30. DNA coating, metal coating and aqueous dispersion of graphene oxide and reduced graphene oxide.
Reprinted with permission from Ref [183].

It is well known that methyl green has a positive charge with high solubility in aqueous

media, therefore it is a promising candidate for adsorbing on the graphite surface via π–π

stacking, weakening the powerful vdW interactions between layers. Hence, the adsorption of

methyl green onto graphite can dramatically increases the solubility of exfoliated graphene

and graphite in aqueous media. To address this issue, Tu et al. [184] functionalized graphene

flakes via water-soluble picket-fence iron porphyrin. As shown in Fig. 31, they presented a

stable and homogenous picket-fence iron porphyrin-treated reduced graphene oxide in water

[184].

71
Fig. 31. TEM images of (left) 5,10,15,20-tetrakis [αααα-2-trismethylammoniome thylphenyl] porphyrin iron
(III) pentachloride/Reduced graphene oxide; and (right) photographs of a) Reduced graphene oxide, b)
5,10,15,20-tetrakis [αααα-2-trismethylammoniome thylphenyl] porphyrin iron (III) pentachloride, and c)
5,10,15,20-tetrakis [αααα-2-trismethylammoniome thylphenyl] porphyrin iron (III) pentachloride (0.1 mg mL-1)
dispensed in water. Reprinted with permission from Ref [184].

In a similar way, porphyrin-functionalized graphene nanoplatelets were synthesized via

hydrazine reduction of graphene oxide with water-soluble porphyrins [185]. Zhang et

al.[186] proposed a method for nonocovalently functionalization of graphene nanoribbons

with water-soluble iron(III) meso-tetrakis(N-methylpyridinum-4-yl) porphyrin. They

employed π–π noncovalent interaction on an electrode surface to increase the degree of

functionalization. Note that the resulting iron(III) meso-tetrakis(N-methylpyridinum-4-yl)

porphyrin-adsorbed graphene nanoribbons composite film illustrated unique electrocatalysis

for being used as a biosensor for the detection of glucose in human serum [186].

Bai et al.[187] reported a noncovalent functionalization method to synthesize sulfonated

polyaniline-graphene nanoplatelets. The resulting products were highly soluble in water and

remained stable for more than two weeks. Because of the electrostatic repulsion between

negatively charged sulfonated polyaniline-graphene nanoplatelets as well as severe π–π

stacking between the graphene basal planes and the main structure of sulfonated polyaniline,

the resulting sulfonated polyaniline-graphene nanoplatelets presented suitable stability at pH

of 1–2 [187].

72
Li and his coworkers[188] synthesized an extremely soluble congo red-treated graphene

nanoplatelets composite by direct coupling. Congo red was a non-covalent functional group.

They claimed that the resulting materials have significant solubility in different organic

solvents as well as in water. It can be seen in Fig. 32 that the congo red-treated graphene

nanoplatelets were fully dispersed in dimethylformamide, water, ethanol, dimethyl sulfoxide,

and methanol and partially dispersed in many other organic solvents. The mechanism of high

dispersibility and stability of congo red-treated graphene nanoplatelets was explained in

terms of Hansen solubility parameters.

Fig. 32. Solubility photographs of congo red-treated graphene nanoplatelets nanocomposites in (a) toluene, (b)
benzene, (c) tetrahydrofuran, (d) ethyl acetate, (e) isopropanol, (f) chloroform, (g) acetone, (h) acetonitrile, (i)
dimethylformamide, (j) methanol, (k) dimethyl sulfoxide, (l) water, (m) N-Methyl-2-pyrrolidone, (n) ethanol, (o)
isoamyl alcohol, (p) m-cresol, (q) pyridine, and (r) 1,4-dioxane. Reprinted with permission from Ref [188].

Water-soluble aromatic electroactive dye has also been to non-covalently functionalize

graphene sheets by Liu and coworkers [189]. They prepared methylene green-treated

graphene during the chemical reduction of graphene oxide with hydrazine. Methylene green-

73
treated graphene showed a well-solubility in water through coulomb repulsion between

methylene green-treated graphene nanoplatelets [189]. Also, Fang et al.[190] investigated

noncovalent functionalization of reduced graphene oxide with chitosan. According to their

results, amino and hydroxyl groups of chitosan can be captured by graphene thorough

hydrogen bonding and zwitterionic interactions, showing considerable solubility in water. It

should be noted that the water-based graphene suspension is biocompatible.

As another type of non-covalent groups, polymers were introduced recently to increase

solubility of graphene sheets in aqueous media [191]. Poly(3,4-ethylene dioxythiophene):

poly(styrene sulfonate) (PEDOT:PSS) was used by Jo et al.[191] for noncovalent

functionalization of reduced graphene oxide nanoplatelets. They reported a significant

solution stability for PEDOT:PSS-treated RGO suspension in aqueous media. Ghosh et

al.[81] proposed a facile and quick method for synthesizing stable water-based single- and

few-layer graphenes dispersions via non-covalent functionalization with a coronene

carboxylate acceptor molecule.

To form a stable aqueous solution of single- and double-layer graphene sheets, a

noncovalent functionalization of graphene sheets with thionine was proposed by Chen and

coworkers [192]. Material characterization ttechniques such as optical spectroscopy

confirmed the e istence of π–π interactions between the thionine and graphene sheets.

Interestingly, the conductivity of the hybrid material including functionalized graphene with

thionine was almost 7 orders of magnitude larger than that observed for graphene oxide

[192]. Similarly, Liu and coworkers[193] could synthesize poly(N-isopropylacrylamide)

[PNIPAAm]-treated graphene upon π–π stacking interactions, which resulted in an aqueous

suspensions with high stability.

Yang and coworkers[194] also proposed an efficient technique for noncovalent

functionalization of graphene sheets with poly(2,5-bis(3-sulfonatopropoxy)-1,4-ethynyl-

74
phenylene-alt-1,4-ethynylphenylene) sodium salt (PPE-SO3–). Highly soluble poly(2,5-bis(3-

sulfonatopropoxy)-1,4-ethynyl-phenylene-alt-1,4-ethynylphenylene) sodium salt (PPE-SO3–)

in water play a key role in providing highly conductive and stable water-based graphene

dispersions for long periods of time. he π stacking interaction between graphene

nanoplatelets and PPE-SO3– is the main proposed mechanism. More interestingly, the

successful adsorption of PPE-SO3– by graphene sheets not only was used as a stabilizer for

aqueous media but also donated negative charges to the resulting graphene sheets that

provided suitable conditions for further functionalization of graphene sheets. Hemin– treated

graphene hybrid nanoplatelets (H-GNs) [195] has also shown high solubility and stability in

water, resulting from π–π interactions. Wang et al. [196] proposed a green method for the

formation of soluble reduced graphene oxide. Upon the reduction of graphene oxide in the

presence of green tea, the resulting green tea-treated graphene sheets showed an appropriate

stability and solubility in aqueous solution and a variety of organic solvents. Park et al.[197]

synthesized a stable and biocompatible polyoxyethylene sorbitan laurate (TWEEN)-adsorbed

RGO platelets with highly stability in aqueous media. This procedure took advantage of the

solubility of TWEEN paper in water and the capability for being adsorbed on graphene

sheets.

The key role of Pluronic copolymer in dispersing graphene in aqueous media was

investigated by Zu et al. [198], which led to the formation of stable aqueous graphene

solutions with three months stability. Liu and coworkers[199] also synthesized an aqueous

dispersion of graphene sheets through the chemically-reduction of graphene oxide using

hydrazine hydrate in the existence of a cationic polyelectrolyte namely poly[(2-

ethyldimethylammonioethyl methacrylate ethyl sulfate)-co-(1-vinylpyrrolidone)] (PQ11).

The resulting PQ11-attached graphene nanoplatelets remained stable for several months. As a

75
positively charged polymer, PQ11 determines the promising capability to reduce silver salts

to Ag nanoparticles.

Amine-terminated polymers were used by Choi and coworkers[200] to disperse graphene

sheets in different organic solvents through noncovalent functionalization. In fact, the

carboxylate groups on reduced graphene oxide formed noncovalent functionalization sites for

the protonated amine terminal groups of end-functional polymers. In fact, the phase transfer

of graphene nanoplatelets from a water phase to an organic phase under sonication procedure

can be facilitated by the noncovalent functionalization. They investigated the solubility and

stability of reduced graphene oxide in different organic solvents, and the photographs of

dispersions after one month confirmed their claim, as shown in Fig. 33. In addition, UV–vis

absorption spectroscopy was also applied to verify the dispersion of the reduced graphene

oxides [200].

Fig. 33. (a) Photographs of reduced graphene dispersed in water and 16 different organic solvents via 5 h
sonication. The photographs were taken 1 month after the preparation of the reduced graphene dispersion. (b)
UV-Vis absorption spectra of reduced graphene dispersed in good solvents. Reprinted with permission from Ref
[200].

76
Qi et al. [201] prepared a graphene-based composite through synthesizing an

amphiphilic reduced graphene oxide using a novel coil-rod-coil conjugated triblock

copolymer as the π–π binding stabilizer. his graphene-based composite showed high-

solubility not only in organic solvents with low polarity but also in water-miscible solvents

with high polarity. As can be seen in Fig. 34, EG-OPE includes two hydrophilic PEG coils

and a lipophilic π-conjugated oligomer (Fig. 34). Due to the amphiphilicity and π-conjugation

of PEG-OPE, the resulting PEG-OPE-treated reduced graphene oxide produce a sandwich

structure, providing soluble composite in a variety of solvents e.g. water and toluene [201].

Fig. 34. (a) The main chemical structure of PEG-OPE. (b) The synthesis procedure of PEG-OPE–treated
Reduced graphene oxide in H2O. (c) Photograph of (A) graphene oxide and (B) Reduced graphene oxide in
water and (C) PEG-OPE–treated Reduced graphene oxide and (D) PEG-OPE in methanol. Reprinted with
permission from Ref [201].

8. Future plan for mass production of high quality graphene from GO

While GO with large lateral dimensions has an exfoliation yield of ~100%, there is a few

complex methods to completely remove the oxygen functional groups. Recently, Voiry et al.

77
[33] reported a simple, rapid method to reduce GO into pristine graphene using 1- to 2-

second pulses of microwaves to completely remove the oxygen functional groups, and the

microwave-reduced GO (MW-rGO) was the final product. Some studies has been reported

the irradiation of GO with microwaves [202, 203], but the reduction efficiency has been low

and the rGO remains highly disordered. In fact, Voiry et al. [33] employed a pre-treatment

via thermal annealing to reduce GO before exposure to microwaves, which resulted in

conductive GO. The produced conductive GO could absorb microwaves, leading to rapid

heating of the GO, desorption of oxygen functional groups and reordering of the graphene

basal plane. Their method was successful and can be considered as a scalable method. To this

end, the authors employed HRTEM to show morphological structure at atomic scale. The

MW-rGO exhibited highly ordered structure at the atomic scale (Fig. 35a). Fig. 35b shows

high-resolution C1s spectra for microwaved reduced graphene oxide (MW-rGO), pristine

graphene oxide (GO), reduced graphene oxide (rGO), graphene grown by chemical vapor

deposition (CVD graphene), and graphite. XPS results indicated that MW-rGO was

completely reduced with an in-plane oxygen concentration of 4%, which is much lower than

that theoretically predicted for rGO after annealing to 1500 K [204]. It can be seen in Fig. 35

panel c that the spectrum obtained for MW-rGO is similar to the spectrum of CVD graphene,

with the presence of a high and symmetrical 2D band together with a minimal D band. Sharp

Raman peaks indicate high crystallinity of MW-rGO and demonstrate the quality of

microwave reduction. They also did 62 measurements on about five different MW-rGO

samples to obtain the ratio of I2D/IG. Results are shown in Fig. 35d. I2D/IG ratios and La values

for MW-rGO approached those of graphene via microwave treatment method. Also, the ratios

were substantially higher than the values for rGO and dispersed graphene. It indicates the

high potential of proposed method for mass-production of large graphene sheets.

78
Fig. 35. (a) HR-TEM of MW-rGO showing highly ordered structure. (b) High-resolution XPS spectra from the

C1s regions for microwaved reduced graphene oxide (MW-rGO) compared with pristine graphene oxide (GO),

reduced graphene oxide (rGO), graphene grown by chemical vapor deposition (CVD graphene), and graphite.

(c) Raman spectra of MW-rGO and other graphene-based samples. (d) Evolution of the I2D/IG ratio versus the

crystal size (La) for MW-rGO, GO, rGO, highly ordered pyrolytic graphite (HOPG), dispersed graphene, and

graphene from (3) (i.e., CVD graphene). Reprinted with permission from Ref [33].

9. Conclusions and outlook

Commercialization of graphene is still one the biggest challenges in the carbon field

despite the development of several methods for its production and mass-preparation of

graphene (mono- or few-layer) is of great importance. As a common method with high

79
potential for scalability, LPE is discussed herein. From the chemistry point of view, the LPE

is a method not only to enhance exfoliation yield but also to avoid graphene re-aggregation

due to van der Waals attraction, thereby compromising the effort made during exfoliation.

From the technology point of view, the LPE method is greatly relevant as many applications

rely on large-scale production using low-cost methods such as ink-jet printing. In this review,

different wet chemical routes for synthesizing graphene have been demonstrated and their

advantages and limitations have discussed as summarized in Table 2. Liquid-phase

exfoliation of graphite into graphene sheets is a mild, flexible and potentially up-scalable

method to synthesize graphene sheets with high quality using simple equipment.

Table 2. Summary of the advantages and disadvantages associated with graphene production by LPE.

Graphene exfoliation Advantages Disadvantages

methods

Direct Ultrasonic Exfoliation in Potentially up-scalable method Low yield of thin graphene layers

Liquid High-quality Small flake size (<5 μm)

Simple preparation method Limited dispersability (for non-

Non-oxidative route in some cases functionalized graphene)

Limited processability

High fix and variable costs

Shear Exfoliation Scalable with low fix and variable costs Low yield of mono layer graphene

Environmentally benign Inhomogeneous flakes

Non-oxidative route Time consuming

Simple preparation method

High yield for production of few-layer

graphene

High-quality

Microfluidization, homogenization Ease of processing Low yield of thin graphene layers

and shear-assisted supercritical Potentially fast method Scalable with high fix and variable

80
techniques costs

Electronic properties are affected

Electrochemical High yield Slight oxidation

exfoliation Fast production Inhomogeneous flake thickness

Cost effectiveness

Good electronic properties

High processability

Environmentally benign

Scalable

Functionalization-assisted High yield Poor quality and/or defective

Exfoliation High dispersability structures

High processability Potentially dangerous process

Synthesizing functionalized sheets Complex process

Time consuming

Use of toxic chemicals

Poor electronic properties in cases

with high degree of functionalization

like GO

In this review article, we have highlighted a variety of methods using liquid phase

exfoliation for mass production of mono- or few-layered graphene sheets. Providing

information for selecting a cost-effective, efficient and facile method for scalable production

of large quantities of few-layered graphene is the main goal of the present review article. As

with all processing methods of different materials, liquid phase exfoliation methods of

graphite have certain advantages and disadvantages. Direct ultrasonic exfoliation of bulk

graphite is non-oxidative and potentially scalable, however high fix and variable costs for

mass-production of graphene can be one of the main problem with this method. Possessing

low yield is another drawback of this method and hinders its promotion in industrial sectors.

81
Shear exfoliation method in some of the organic solvents such as NMP, DMF, DMA, etc. or

stabilizer-treated aqueous media resulted in improved exfoliation yield with defect free,

unoxidized graphene sheets obtained in liquid volumes of hundreds of liters. Yet, the yield of

thin graphene layer remains low with limited processability and dispersibility. Wet-ball

milling could provide enough driving force to enhance the performance of shear exfoliation.

However, the exfoliated graphene sheets obtained from shear or ultrasonic exfoliation tend to

strongly aggregate due to vdW attraction once solidification, which spoil the exfoliation

yield. Functionalization, in particular with charged functional groups, can improve the yield

of LPE methods with excellent solution processability to produce single layer graphene, but it

suffers from extremely poor quality or structural inhomogeneity and complex procedure.

Being costly can be considered as another drawback of functionalization-assisted methods.

Also, removing functional groups is a difficult task, impossible to remove the whole defects

and not cost-effective in most of the cases. Voiry et al. [33] reported a simple, rapid method

to reduce GO into pristine graphene using 1- to 2-second pulses of microwaves to completely

remove the oxygen functional groups, and the pure graphene was the final product. It can

open a new gateway to mass-production of high-quality graphene if microwave method has

the protentional of scale up. Before publishing this article, the defects within GO cannot be

fully restored with current reduction methods. In addition, the preparation of functionalized

graphene such as graphene oxide needs the usage of huge amount of strong acids and/or

oxidizing agents and the procedure generally takes up to several days, which hinders their

industrial scale production. Obviously, electrochemical exfoliation is an outstanding

approach of LPE to synthesize high quality graphene via a wet chemical route. This process

provided a great solution to overcome the restrictions of functionalization-assisted

exfoliation, shear and ultrasonic exfoliation methods by producing high-quality graphene

with low degree of oxidation at a large scale within a short time period and in an

82
environmentally friendly manner. However, there are some challenges with this approach

such as synthesizing graphene with non-homogeneous flake size and layer distribution. In

spite of many researches in this filed, finding a method for mass-production of graphene with

ability of tuning dimensions and thickness of flakes is critical. he supercritical fluid

exfoliation methods and microfluidization techniques are also very promising methods for the

preparation of graphene because ease of processing and a high yield production of

monolayers. A combination of supercritical fluid e foliation methods, microfluidization,

electrochemical and shear exfoliation like wet-ball milling have a huge potential for scalable

production of mono layer graphene. Also, graphene prepared by a scalable method is crucial

for their real applications in conductive ink, composite materials, supercapacitors, fuel cells,

batteries, etc. In particular, not only the application performance of graphene must

outperform but also the production cost should be less than the existing carbon materials such

as carbon black, carbon nanotubes, activated carbon, in order to motivate the industries to

adopt this new material.

References
[1] A.K. Rasheed, M. Khalid, W. Rashmi, T.C.S.M. Gupta, A. Chan, Renewable and
Sustainable Energy Reviews, 63 (2016) 346-362.
[2] Ahmad Amiri, Mehdi Shanbedi, Goodarz Ahmadi, Hossein Eshghi, S. N. Kazi, B.
T. Chew, M. Savari, M.N.M. Zubir1, Scientific Reports, 6 (2016) 32686.
[3] A. Amiri, G. Ahmadi, M. Shanbedi, M. Savari, S. Kazi, B. Chew, Scientific
reports, 5 (2015).
[4] A. Amiri, M. Shanbedi, B. Chew, S. Kazi, K. Solangi, Chem. Eng. J., 289 (2016)
583-595.
[5] A.K. Geim, K.S. Novoselov, Nature materials, 6 (2007) 183-191.
[6] Y. Zhu, S. Murali, W. Cai, X. Li, J.W. Suk, J.R. Potts, R.S. Ruoff, Advanced
materials, 22 (2010) 3906-3924.
[7] X. Huang, Z. Yin, S. Wu, X. Qi, Q. He, Q. Zhang, Q. Yan, F. Boey, H. Zhang,
small, 7 (2011) 1876-1902.
[8] K.S. Novoselov, V. Fal, L. Colombo, P. Gellert, M. Schwab, K. Kim, Nature, 490
(2012) 192-200.
[9] J. Abraham, K.S. Vasu, C.D. Williams, K. Gopinadhan, Y. Su, C.T. Cherian, J.
Dix, E. Prestat, S.J. Haigh, I.V. Grigorieva, Nature Nanotechnology, (2017).
[10] Y. Yang, X. Yang, X. Zou, S. Wu, D. Wan, A. Cao, L. Liao, Q. Yuan, X. Duan,
Advanced Functional Materials, 27 (2017).

83
[11] H. Sun, L. Mei, J. Liang, Z. Zhao, C. Lee, H. Fei, M. Ding, J. Lau, M. Li, C.
Wang, Science, 356 (2017) 599-604.
[12] C. Lee, X. Wei, J.W. Kysar, J. Hone, science, 321 (2008) 385-388.
[13] J. Li, L. Niu, Z. Zheng, F. Yan, Advanced Materials, 26 (2014) 5239-5273.
[14] B. Sensale-Rodriguez, T. Fang, R. Yan, M.M. Kelly, D. Jena, L. Liu, H.G. Xing,
Applied Physics Letters, 99 (2011) 113104.
[15] M.V. Shelke, H. Gullapalli, K. Kalaga, M.T.F. Rodrigues, R.R. Devarapalli, R.
Vajtai, P.M. Ajayan, Advanced Materials Interfaces, 4 (2017).
[16] V. Strauss, K. Marsh, M.D. Kowal, M. El‐ Kady, R.B. Kaner, Advanced
Materials, (2018).
[17] F. Bonaccorso, A. Lombardo, T. Hasan, Z. Sun, L. Colombo, A.C. Ferrari,
Materials Today, 15 (2012) 564-589.
[18] S. Deng, V. Berry, Materials Today, 19 (2016) 197-212.
[19] E.P. Randviir, D.A.C. Brownson, C.E. Banks, Materials Today, 17 (2014) 426-
432.
[20] Y.L. Zhong, Z. Tian, G.P. Simon, D. Li, Materials Today, 18 (2015) 73-78.
[21] G. Bepete, E. Anglaret, L. Ortolani, V. Morandi, K. Huang, A. Pénicaud, C.
Drummond, Nature chemistry, 9 (2017) 347.
22 G. epete, A. nicaud, C. Drummond, E. Anglaret, The Journal of Physical
Chemistry C, 120 (2016) 28204-28214.
[23] Y. Wang, Y. Zheng, X. Xu, E. Dubuisson, Q. Bao, J. Lu, K.P. Loh, ACS nano, 5
(2011) 9927-9933.
[24] A.N. Obraztsov, Nature nanotechnology, 4 (2009) 212-213.
[25] C. Mattevi, H. Kim, M. Chhowalla, Journal of Materials Chemistry, 21 (2011)
3324-3334.
[26] A. Ferrari, J. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S.
Piscanec, D. Jiang, K. Novoselov, S. Roth, Physical review letters, 97 (2006) 187401.
[27] P.W. Sutter, J.-I. Flege, E.A. Sutter, Nature materials, 7 (2008) 406-411.
[28] J.R. Lomeda, C.D. Doyle, D.V. Kosynkin, W.-F. Hwang, J.M. Tour, Journal of
the American Chemical Society, 130 (2008) 16201-16206.
[29] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev,
L.B. Alemany, W. Lu, J.M. Tour, ACS nano, 4 (2010) 4806-4814.
[30] Z. Sun, D.K. James, J.M. Tour, The Journal of Physical Chemistry Letters, 2
(2011) 2425-2432.
[31] S. Stankovich, D.A. Dikin, R.D. Piner, K.A. Kohlhaas, A. Kleinhammes, Y. Jia,
Y. Wu, S.T. Nguyen, R.S. Ruoff, carbon, 45 (2007) 1558-1565.
[32] H.A. Becerril, J. Mao, Z. Liu, R.M. Stoltenberg, Z. Bao, Y. Chen, ACS nano, 2
(2008) 463-470.
[33] D. Voiry, J. Yang, J. Kupferberg, R. Fullon, C. Lee, H.Y. Jeong, H.S. Shin, M.
Chhowalla, Science, 353 (2016) 1413-1416.
[34] Y. Hernandez, V. Nicolosi, M. Lotya, F.M. Blighe, Z. Sun, S. De, I. McGovern,
B. Holland, M. Byrne, Y.K. Gun'Ko, Nature nanotechnology, 3 (2008) 563-568.
35 K.R. aton, E. Varrla, C. ackes, R.J. Smith, U. Khan, A. O’Neill, C. oland, M.
Lotya, O.M. Istrate, P. King, Nature materials, 13 (2014) 624.
[36] R.S. Edwards, K.S. Coleman, Nanoscale, 5 (2013) 38-51.
[37] M. Ayan-Varela, J. Paredes, S. Villar-Rodil, R. Rozada, A. Martinez-Alonso, J.
Tascón, Carbon, 75 (2014) 390-400.
[38] Z. Sun, S.-i. Kohama, Z. Zhang, J.R. Lomeda, J.M. Tour, Nano Research, 3
(2010) 117-125.

84
[39] D.W. Johnson, B.P. Dobson, K.S. Coleman, Current Opinion in Colloid &
Interface Science, 20 (2015) 367-382.
[40] T.J. Mason, J.P. Lorimer, Applied Sonochemistry: Uses of Power Ultrasound in
Chemistry and Processing, (2002) 1-24.
[41] L. Niu, J.N. Coleman, H. Zhang, H. Shin, M. Chhowalla, Z. Zheng, Small, 12
(2016) 272-293.
[42] J.N. Israelachvili, Intermolecular and surface forces: revised third edition,
Academic press, 2011.
[43] P. Blake, P.D. Brimicombe, R.R. Nair, T.J. Booth, D. Jiang, F. Schedin, L.A.
Ponomarenko, S.V. Morozov, H.F. Gleeson, E.W. Hill, Nano letters, 8 (2008) 1704-1708.
[44] C.E. Hamilton, J.R. Lomeda, Z. Sun, J.M. Tour, A.R. Barron, Nano letters, 9
(2009) 3460-3462.
[45] A.B. Bourlinos, V. Georgakilas, R. Zboril, T.A. Steriotis, A.K. Stubos, Small, 5
(2009) 1841-1845.
[46] U. Khan, A. O'Neill, M. Lotya, S. De, J.N. Coleman, Small, 6 (2010) 864-871.
47 U. Khan, H. orwal, A. O’Neill, K. Nawaz, . May, J.N. Coleman, Langmuir, 27
(2011) 9077-9082.
48 U. Khan, . May, A. O’Neill, J.N. Coleman, Carbon, 48 (2010) 4035-4041.
[49] J. Shen, Y. He, J. Wu, C. Gao, K. Keyshar, X. Zhang, Y. Yang, M. Ye, R. Vajtai,
J. Lou, Nano letters, 15 (2015) 5449-5454.
[50] H. Gao, K. Zhu, G. Hu, C. Xue, Chemical Engineering Journal, 308 (2017) 872-
879.
51 A. O’Neill, U. Khan, .N. Nirmalraj, J. oland, J.N. Coleman, he Journal of
Physical Chemistry C, 115 (2011) 5422-5428.
[52] E.-Y. Choi, W. San Choi, Y.B. Lee, Y.-Y. Noh, Nanotechnology, 22 (2011)
365601.
[53] W. Qian, R. Hao, Y. Hou, Y. Tian, C. Shen, H. Gao, X. Liang, Nano Research, 2
(2009) 706-712.
[54] X. Zhang, A.C. Coleman, N. Katsonis, W.R. Browne, B.J. Van Wees, B.L.
Feringa, Chemical Communications, 46 (2010) 7539-7541.
[55] A. Catheline, L. Ortolani, V. Morandi, M. Melle-Franco, C. Drummond, C.
Zakri, A. Pénicaud, Soft Matter, 8 (2012) 7882-7887.
[56] M. Lotya, Y. Hernandez, P.J. King, R.J. Smith, V. Nicolosi, L.S. Karlsson, F.M.
Blighe, S. De, Z. Wang, I. McGovern, Journal of the American Chemical Society, 131 (2009)
3611-3620.
[57] M. Lotya, P.J. King, U. Khan, S. De, J.N. Coleman, ACS nano, 4 (2010) 3155-
3162.
[58] S. De, P.J. King, M. Lotya, A. O'Neill, E.M. Doherty, Y. Hernandez, G.S.
Duesberg, J.N. Coleman, Small, 6 (2010) 458-464.
[59] A.A. Green, M.C. Hersam, Nano letters, 9 (2009) 4031-4036.
[60] T. Hasan, F. Torrisi, Z. Sun, D. Popa, V. Nicolosi, G. Privitera, F. Bonaccorso,
A. Ferrari, physica status solidi (b), 247 (2010) 2953-2957.
[61] R. Hao, W. Qian, L. Zhang, Y. Hou, Chemical Communications, (2008) 6576-
6578.
[62] S. Vadukumpully, J. Paul, S. Valiyaveettil, Carbon, 47 (2009) 3288-3294.
[63] S.M. Notley, Langmuir, 28 (2012) 14110-14113.
[64] L. Guardia, M. Fernández-Merino, J. Paredes, P. Solis-Fernandez, S. Villar-
Rodil, A. Martinez-Alonso, J. Tascón, Carbon, 49 (2011) 1653-1662.
[65] J. Geng, B.-S. Kong, S.B. Yang, H.-T. Jung, Chemical Communications, 46
(2010) 5091-5093.

85
[66] R.J. Smith, M. Lotya, J.N. Coleman, New Journal of Physics, 12 (2010) 125008.
[67] A.B. Bourlinos, V. Georgakilas, R. Zboril, T.A. Steriotis, A.K. Stubos, C.
Trapalis, Solid State Commun., 149 (2009) 2172-2176.
[68] Y.T. Liang, M.C. Hersam, Journal of the American Chemical Society, 132
(2010) 17661-17663.
[69] P. May, U. Khan, J.M. Hughes, J.N. Coleman, The Journal of Physical
Chemistry C, 116 (2012) 24390-24391.
[70] L. Xu, J.-W. McGraw, F. Gao, M. Grundy, Z. Ye, Z. Gu, J.L. Shepherd, The
Journal of Physical Chemistry C, 117 (2013) 10730-10742.
[71] T. Skaltsas, N. Karousis, H.-J. Yan, C.-R. Wang, S. Pispas, N. Tagmatarchis,
Journal of Materials Chemistry, 22 (2012) 21507-21512.
[72] F. Liu, J.Y. Choi, T.S. Seo, Chemical Communications, 46 (2010) 2844-2846.
[73] M. Zhang, R.R. Parajuli, D. Mastrogiovanni, B. Dai, P. Lo, W. Cheung, R.
Brukh, P.L. Chiu, T. Zhou, Z. Liu, Small, 6 (2010) 1100-1107.
[74] D.-W. Lee, T. Kim, M. Lee, Chemical Communications, 47 (2011) 8259-8261.
[75] D. Parviz, S. Das, H.T. Ahmed, F. Irin, S. Bhattacharia, M.J. Green, Acs Nano, 6
(2012) 8857-8867.
[76] A. Schlierf, H. Yang, E. Gebremedhn, E. Treossi, L. Ortolani, L. Chen, A.
Minoia, V. Morandi, P. Samori, C. Casiraghi, Nanoscale, 5 (2013) 4205-4216.
[77] H. Yang, Y. Hernandez, A. Schlierf, A. Felten, A. Eckmann, S. Johal, P. Louette,
J.-J. Pireaux, X. Feng, K. Mullen, Carbon, 53 (2013) 357-365.
[78] J.-H. Jang, D. Rangappa, Y.-U. Kwon, I. Honma, Journal of Materials
Chemistry, 21 (2011) 3462-3466.
[79] X. Dong, Y. Shi, Y. Zhao, D. Chen, J. Ye, Y. Yao, F. Gao, Z. Ni, T. Yu, Z. Shen,
Physical review letters, 102 (2009) 135501.
[80] X. An, T. Simmons, R. Shah, C. Wolfe, K.M. Lewis, M. Washington, S.K.
Nayak, S. Talapatra, S. Kar, Nano letters, 10 (2010) 4295-4301.
[81] A. Ghosh, K.V. Rao, S.J. George, C. Rao, Chemistry–A European Journal, 16
(2010) 2700-2704.
[82] J.M. Englert, J. Röhrl, C.D. Schmidt, R. Graupner, M. Hundhausen, F. Hauke, A.
Hirsch, Advanced Materials, 21 (2009) 4265-4269.
[83] X. Zhou, T. Wu, K. Ding, B. Hu, M. Hou, B. Han, Chemical Communications,
46 (2010) 386-388.
[84] S. Gunasekaran, M. Dharmendirakumar, High Perform. Polym., 26 (2014) 274-
282.
[85] X. Wang, P.F. Fulvio, G.A. Baker, G.M. Veith, R.R. Unocic, S.M. Mahurin, M.
Chi, S. Dai, Chemical Communications, 46 (2010) 4487-4489.
[86] V. Alzari, D. Nuvoli, R. Sanna, S. Scognamillo, M. Piccinini, J.M. Kenny, G.
Malucelli, A. Mariani, Journal of Materials Chemistry, 21 (2011) 16544-16549.
[87] D. Nuvoli, L. Valentini, V. Alzari, S. Scognamillo, S.B. Bon, M. Piccinini, J.
Illescas, A. Mariani, Journal of Materials Chemistry, 21 (2011) 3428-3431.
[88] K.S. Rao, J. Sentilnathan, H.W. Cho, J.J. Wu, M. Yoshimura, Advanced
Functional Materials, 25 (2015) 298-305.
[89] C. Vallés, C. Drummond, H. Saadaoui, C.A. Furtado, M. He, O. Roubeau, L.
Ortolani, M. Monthioux, A. Pénicaud, Journal of the American Chemical Society, 130 (2008)
15802-15804.
[90] A. Catheline, C. Valles, C. Drummond, L. Ortolani, V. Morandi, M. Marcaccio,
M. Iurlo, F. Paolucci, A. Pénicaud, Chemical Communications, 47 (2011) 5470-5472.
[91] D. Murphy, G. Hull Jr, The Journal of Chemical Physics, 62 (1975) 973-978.

86
[92] D.B. Shinde, J. Brenker, C.D. Easton, R.F. Tabor, A. Neild, M. Majumder,
Langmuir, 32 (2016) 3552-3559.
[93] T.S. Tran, S.J. Park, S.S. Yoo, T.-R. Lee, T. Kim, RSC Advances, 6 (2016)
12003-12008.
[94] L. Liu, Z. Shen, M. Yi, X. Zhang, S. Ma, RSC Advances, 4 (2014) 36464-36470.
[95] E. Varrla, K.R. Paton, C. Backes, A. Harvey, R.J. Smith, J. McCauley, J.N.
Coleman, Nanoscale, 6 (2014) 11810-11819.
[96] C. Teng, D. Xie, J. Wang, Z. Yang, G. Ren, Y. Zhu, Advanced Functional
Materials, 27 (2017).
[97] W. Zhao, M. Fang, F. Wu, H. Wu, L. Wang, G. Chen, Journal of materials
chemistry, 20 (2010) 5817-5819.
[98] A. Amiri, M.N.M. Zubir, A.M. Dimiev, K. Teng, M. Shanbedi, S. Kazi, S.B.
Rozali, Chemical Engineering Journal, (2017).
[99] P.G. Karagiannidis, S.A. Hodge, L. Lombardi, F. Tomarchio, N. Decorde, S.
Milana, I. Goykhman, Y. Su, S.V. Mesite, D.N. Johnstone, ACS nano, 11 (2017) 2742-2755.
[100] K.R. Paton, J. Anderson, A.J. Pollard, T. Sainsbury, Materials Research
Express, 4 (2017) 025604.
[101] G. Cesareo, M.R. Parrini, L.G. Rizzi, in, Google Patents, 2017.
[102] L. Li, J. Xu, G. Li, X. Jia, Y. Li, F. Yang, L. Zhang, C. Xu, J. Gao, Y. Liu,
Chemical Engineering Journal, 284 (2016) 78-84.
[103] N. Song, J. Jia, W. Wang, Y. Gao, Y. Zhao, Y. Chen, Chemical Engineering
Journal, 298 (2016) 198-205.
[104] A. Abdelkader, A. Cooper, R. Dryfe, I. Kinloch, Nanoscale, 7 (2015) 6944-
6956.
[105] C.-Y. Su, A.-Y. Lu, Y. Xu, F.-R. Chen, A.N. Khlobystov, L.-J. Li, ACS nano, 5
(2011) 2332-2339.
[106 K. arvez, R. i, S.R. uniredd, Y. Hernandez, F. Hinkel, S. Wang, . Feng, K.
M llen, ACS nano, 7 (2013) 3598-3606.
107 K. arvez, Z.-S. Wu, R. i, . iu, R. Graf, . Feng, K. M llen, Journal of the
American Chemical Society, 136 (2014) 6083-6091.
[108] Z.S. Wu, Z. Liu, K. Parvez, X. Feng, K. Müllen, Advanced Materials, 27 (2015)
3669-3675.
[109] G. Wang, B. Wang, J. Park, Y. Wang, B. Sun, J. Yao, Carbon, 47 (2009) 3242-
3246.
[110] J.P. Mensing, T. Kerdcharoen, C. Sriprachuabwong, A. Wisitsoraat, D.
Phokharatkul, T. Lomas, A. Tuantranont, Journal of Materials Chemistry, 22 (2012) 17094-
17099.
[111] T. Kuila, P. Khanra, N.H. Kim, S.K. Choi, H.J. Yun, J.H. Lee, Nanotechnology,
24 (2013) 365706.
[112] N. Liu, F. Luo, H. Wu, Y. Liu, C. Zhang, J. Chen, Adv. Funct. Mater., 18
(2008) 1518-1525.
[113] J. Lu, J.-x. Yang, J. Wang, A. Lim, S. Wang, K.P. Loh, ACS Nano, 3 (2009)
2367-2375.
[114] P. Khanra, T. Kuila, S.H. Bae, N.H. Kim, J.H. Lee, Journal of Materials
Chemistry, 22 (2012) 24403-24410.
[115] F. Zeng, Z. Sun, X. Sang, D. Diamond, K.T. Lau, X. Liu, D.S. Su,
ChemSusChem, 4 (2011) 1587-1591.
[116] M. Zhou, J. Tang, Q. Cheng, G. Xu, P. Cui, L.-C. Qin, Chemical Physics
Letters, 572 (2013) 61-65.

87
[117] Y. Yang, F. Lu, Z. Zhou, W. Song, Q. Chen, X. Ji, Electrochimica Acta, 113
(2013) 9-16.
[118] Z. Zeng, Z. Yin, X. Huang, H. Li, Q. He, G. Lu, F. Boey, H. Zhang,
Angewandte Chemie International Edition, 50 (2011) 11093-11097.
[119] X. Rocquefelte, F. Boucher, P. Gressier, G. Ouvrard, P. Blaha, K. Schwarz,
Physical Review B, 62 (2000) 2397.
[120] J. Wang, K.K. Manga, Q. Bao, K.P. Loh, Journal of the American Chemical
Society, 133 (2011) 8888-8891.
[121] P.C. Shi, J.P. Guo, X. Liang, S. Cheng, H. Zheng, Y. Wang, C.H. Chen, H.F.
Xiang, Carbon, 126 (2018) 507-513.
[122] S. Tian, S. Yang, T. Huang, J. Sun, H. Wang, X. Pu, L. Tian, P. He, G. Ding, X.
Xie, Carbon, 111 (2017) 617-621.
[123] C.-H. Chen, S.-W. Yang, M.-C. Chuang, W.-Y. Woon, C.-Y. Su, Nanoscale, 7
(2015) 15362-15373.
124 M. Alanyalıoğlu, J.J. Segura, J. Oro-Sole, N. Casan-Pastor, Carbon, 50 (2012)
142-152.
[125] J. Liu, M. Notarianni, G. Will, V.T. Tiong, H. Wang, N. Motta, Langmuir, 29
(2013) 13307-13314.
[126] B.D. Ossonon, D. Bélanger, Carbon, 111 (2017) 83-93.
[127] K.S. Rao, J. Senthilnathan, Y.-F. Liu, M. Yoshimura, Scientific reports, 4
(2014) 4237.
[128] Y.L. Zhong, T.M. Swager, Journal of the American Chemical Society, 134
(2012) 17896-17899.
[129] A.M. Abdelkader, I.A. Kinloch, R.A. Dryfe, ACS applied materials &
interfaces, 6 (2014) 1632-1639.
[130] A.J. Cooper, N.R. Wilson, I.A. Kinloch, R.A. Dryfe, Carbon, 66 (2014) 340-
350.
[131] Z.Y. Xia, G. Giambastiani, C. Christodoulou, M.V. Nardi, N. Koch, E. Treossi,
V. Bellani, S. Pezzini, F. Corticelli, V. Morandi, ChemPlusChem, 79 (2014) 439-446.
[132] S. Stankovich, R.D. Piner, X. Chen, N. Wu, S.T. Nguyen, R.S. Ruoff, Journal
of Materials Chemistry, 16 (2006) 155-158.
[133] S. Stankovich, R.D. Piner, S.T. Nguyen, R.S. Ruoff, Carbon, 44 (2006) 3342-
3347.
[134] G.L. Li, G. Liu, M. Li, D. Wan, K. Neoh, E. Kang, The Journal of Physical
Chemistry C, 114 (2010) 12742-12748.
[135] J.I. Paredes, S. Villar-Rodil, A. Martínez-Alonso, J.M.D. Tascón, Langmuir, 24
(2008) 10560-10564.
[136] W. Cai, R.D. Piner, F.J. Stadermann, S. Park, M.A. Shaibat, Y. Ishii, D. Yang,
A. Velamakanni, S.J. An, M. Stoller, Science, 321 (2008) 1815-1817.
[137] H. He, J. Klinowski, M. Forster, A. Lerf, Chemical physics letters, 287 (1998)
53-56.
[138] S. Park, K.-S. Lee, G. Bozoklu, W. Cai, S.T. Nguyen, R.S. Ruoff, ACS nano, 2
(2008) 572-578.
[139] S. Park, R.S. Ruoff, Current Opinion in Colloid & Interface Science, 20 (2015)
322-328.
[140] R. Ruoff, Nature Nanotechnology, 3 (2008) 10-11.
[141] H. Wang, J.T. Robinson, X. Li, H. Dai, Journal of the American Chemical
Society, 131 (2009) 9910-9911.
[142] S. Stankovich, D.A. Dikin, G.H. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A.
Stach, R.D. Piner, S.T. Nguyen, R.S. Ruoff, nature, 442 (2006) 282-286.

88
[143] S. Park, J. An, I. Jung, R.D. Piner, S.J. An, X. Li, A. Velamakanni, R.S. Ruoff,
Nano letters, 9 (2009) 1593-1597.
[144] R. Sharma, J.H. Baik, C.J. Perera, M.S. Strano, Nano letters, 10 (2010) 398-
405.
[145] S. Niyogi, E. Bekyarova, M.E. Itkis, H. Zhang, K. Shepperd, J. Hicks, M.
Sprinkle, C. Berger, C.N. Lau, W.A. Deheer, Nano letters, 10 (2010) 4061-4066.
[146] A. Amiri, M. Maghrebi, M. Baniadam, S.Z. Heris, Appl. Surf. Sci., 257 (2011)
10261-10266.
[147] A. Amiri, R. Sadri, M. Shanbedi, G. Ahmadi, B. Chew, S. Kazi, M. Dahari,
Energy Conversion and Management, 92 (2015) 322-330.
[148] A. Amiri, M. Shanbedi, H. Amiri, S.Z. Heris, S. Kazi, B. Chew, H. Eshghi,
Appl. Therm. Eng., 71 (2014) 450-459.
[149] A. Amiri, M. Shanbedi, H. Eshghi, S.Z. Heris, M. Baniadam, The Journal of
Physical Chemistry C, 116 (2012) 3369-3375.
[150] H.Z. Zardini, A. Amiri, M. Shanbedi, M. Maghrebi, M. Baniadam, Colloids and
Surfaces B: Biointerfaces, 92 (2012) 196-202.
[151] M.Z. Hossain, M.A. Walsh, M.C. Hersam, Journal of the American Chemical
Society, 132 (2010) 15399-15403.
[152] Z. Jin, J.R. Lomeda, B.K. Price, W. Lu, Y. Zhu, J.M. Tour, Chemistry of
Materials, 21 (2009) 3045-3047.
[153] B. Ghiadi, M. Baniadam, M. Maghrebi, A. Amiri, Russian Journal of Physical
Chemistry A, 87 (2013) 649-653.
[154] M. Fang, K. Wang, H. Lu, Y. Yang, S. Nutt, Journal of Materials Chemistry, 19
(2009) 7098-7105.
[155] Y. Zhu, M.D. Stoller, W. Cai, A. Velamakanni, R.D. Piner, D. Chen, R.S.
Ruoff, ACS nano, 4 (2010) 1227-1233.
[156] S. Park, R.S. Ruoff, Nature nanotechnology, 4 (2009) 217-224.
[157] H. Liu, S. Ryu, Z. Chen, M.L. Steigerwald, C. Nuckolls, L.E. Brus, Journal of
the American Chemical Society, 131 (2009) 17099-17101.
[158] W.S. Sarsam, A. Amiri, M.N.M. Zubir, H. Yarmand, S.N. Kazi, A. Badarudin,
Colloids and Surfaces A: Physicochemical and Engineering Aspects, 500 (2016) 17-31.
[159] V. Georgakilas, M. Otyepka, A.B. Bourlinos, V. Chandra, N. Kim, K.C. Kemp,
P. Hobza, R. Zboril, K.S. Kim, Chem. Rev., 112 (2012) 6156-6214.
[160] V. Georgakilas, A. Bourlinos, D. Gournis, T. Tsoufis, C. Trapalis, A. Mateo-
Alonso, M. Prato, Journal of the American Chemical Society, 130 (2008) 8733-8740.
[161] K. Kostarelos, L. Lacerda, G. Pastorin, W. Wu, S. Wieckowski, J. Luangsivilay,
S. Godefroy, D. Pantarotto, J.-P. Briand, S. Muller, Nature nanotechnology, 2 (2007) 108-
113.
[162] M. Dabiri, M. Kasmaei, P. Salari, S.K. Movahed, RSC Advances, (2016).
[163] A.B. Bourlinos, V. Georgakilas, R. Zboril, T.A. Steriotis, A.K. Stubos, Small, 5
(2009) 1841-1845.
[164] V. Georgakilas, A.B. Bourlinos, R. Zboril, T.A. Steriotis, P. Dallas, A.K.
Stubos, C. Trapalis, Chemical Communications, 46 (2010) 1766-1768.
[165] P. Nemes-Incze, Z. Osváth, K. Kamarás, L. Biró, Carbon, 46 (2008) 1435-1442.
[166] X. Zhang, L. Hou, A. Cnossen, A.C. Coleman, O. Ivashenko, P. Rudolf, B.J.
van Wees, W.R. Browne, B.L. Feringa, Chemistry–A European Journal, 17 (2011) 8957-
8964.
[167] M. Quintana, K. Spyrou, M. Grzelczak, W.R. Browne, P. Rudolf, M. Prato,
ACS nano, 4 (2010) 3527-3533.
[168] L.-H. Liu, M.M. Lerner, M. Yan, Nano letters, 10 (2010) 3754-3756.

89
[169] T.A. Strom, E.P. Dillon, C.E. Hamilton, A.R. Barron, Chemical
Communications, 46 (2010) 4097-4099.
[170] S. Vadukumpully, J. Gupta, Y. Zhang, G.Q. Xu, S. Valiyaveettil, Nanoscale, 3
(2011) 303-308.
[171] H. He, C. Gao, Chemistry of Materials, 22 (2010) 5054-5064.
[172] J. Choi, K.-j. Kim, B. Kim, H. Lee, S. Kim, The Journal of Physical Chemistry
C, 113 (2009) 9433-9435.
[173] X. Zhong, J. Jin, S. Li, Z. Niu, W. Hu, R. Li, J. Ma, Chemical Communications,
46 (2010) 7340-7342.
[174] H. Jaegfeldt, T. Kuwana, G. Johansson, Journal of the American Chemical
Society, 105 (1983) 1805-1814.
[175] Y. Xu, H. Bai, G. Lu, C. Li, G. Shi, Journal of the American Chemical Society,
130 (2008) 5856-5857.
[176] Y. Wang, X. Chen, Y. Zhong, F. Zhu, K.P. Loh, Applied Physics Letters, 95
(2009) 063302.
[177] X. An, T.W. Butler, M. Washington, S.K. Nayak, S. Kar, ACS nano, 5 (2011)
1003-1011.
[178] Q. Yang, X. Pan, F. Huang, K. Li, The Journal of Physical Chemistry C, 114
(2010) 3811-3816.
[179] V.K. Kodali, J. Scrimgeour, S. Kim, J.H. Hankinson, K.M. Carroll, W.A. De
Heer, C. Berger, J.E. Curtis, Langmuir, 27 (2010) 863-865.
[180] M. Lopes, A. Candini, M. Urdampilleta, A. Reserbat-Plantey, V. Bellini, S.
Klyatskaya, L. Marty, M. Ruben, M. Affronte, W. Wernsdorfer, ACS nano, 4 (2010) 7531-
7537.
[181] Q. Su, S. Pang, V. Alijani, C. Li, X. Feng, K. Müllen, Advanced Materials, 21
(2009) 3191-3195.
[182] Y. Liang, D. Wu, X. Feng, K. Müllen, Advanced materials, 21 (2009) 1679-
1683.
[183] J. Liu, Y. Li, Y. Li, J. Li, Z. Deng, Journal of Materials Chemistry, 20 (2010)
900-906.
[184] W. Tu, J. Lei, S. Zhang, H. Ju, Chemistry–A European Journal, 16 (2010)
10771-10777.
[185] J. Geng, H.-T. Jung, The Journal of Physical Chemistry C, 114 (2010) 8227-
8234.
[186] S. Zhang, S. Tang, J. Lei, H. Dong, H. Ju, Journal of electroanalytical
chemistry, 656 (2011) 285-288.
[187] H. Bai, Y. Xu, L. Zhao, C. Li, G. Shi, Chemical Communications, (2009) 1667-
1669.
[188] F. Li, Y. Bao, J. Chai, Q. Zhang, D. Han, L. Niu, Langmuir, 26 (2010) 12314-
12320.
[189] H. Liu, J. Gao, M. Xue, N. Zhu, M. Zhang, T. Cao, Langmuir, 25 (2009) 12006-
12010.
[190] M. Fang, J. Long, W. Zhao, L. Wang, G. Chen, Langmuir, 26 (2010) 16771-
16774.
[191] K. Jo, T. Lee, H.J. Choi, J.H. Park, D.J. Lee, D.W. Lee, B.-S. Kim, Langmuir,
27 (2011) 2014-2018.
[192] C. Chen, W. Zhai, D. Lu, H. Zhang, W. Zheng, Mater. Res. Bull., 46 (2011)
583-587.
[193] J. Liu, W. Yang, L. Tao, D. Li, C. Boyer, T.P. Davis, J. Polym. Sci., Part A:
Polym. Chem., 48 (2010) 425-433.

90
[194] H. Yang, Q. Zhang, C. Shan, F. Li, D. Han, L. Niu, Langmuir, 26 (2010) 6708-
6712.
[195] Y. Guo, L. Deng, J. Li, S. Guo, E. Wang, S. Dong, ACS nano, 5 (2011) 1282-
1290.
[196] Y. Wang, Z. Shi, J. Yin, ACS applied materials & interfaces, 3 (2011) 1127-
1133.
[197] S. Park, N. Mohanty, J.W. Suk, A. Nagaraja, J. An, R.D. Piner, W. Cai, D.R.
Dreyer, V. Berry, R.S. Ruoff, Advanced Materials, 22 (2010) 1736-1740.
[198] S.-Z. Zu, B.-H. Han, The Journal of Physical Chemistry C, 113 (2009) 13651-
13657.
[199] S. Liu, J. Tian, L. Wang, H. Li, Y. Zhang, X. Sun, Macromolecules, 43 (2010)
10078-10083.
[200] E.-Y. Choi, T.H. Han, J. Hong, J.E. Kim, S.H. Lee, H.W. Kim, S.O. Kim,
Journal of Materials Chemistry, 20 (2010) 1907-1912.
[201] X. Qi, K.Y. Pu, H. Li, X. Zhou, S. Wu, Q.L. Fan, B. Liu, F. Boey, W. Huang,
H. Zhang, Angewandte Chemie International Edition, 49 (2010) 9426-9429.
[202] Y. Zhu, S. Murali, M.D. Stoller, A. Velamakanni, R.D. Piner, R.S. Ruoff,
Carbon, 48 (2010) 2118-2122.
[203] W. Chen, L. Yan, P.R. Bangal, Carbon, 48 (2010) 1146-1152.
[204] A. Bagri, C. Mattevi, M. Acik, Y.J. Chabal, M. Chhowalla, V.B. Shenoy,
Nature chemistry, 2 (2010) 581.

91
 Classification of different methods for mass-production of mono- and few layer graphene is done.

 In situ liquid-phase exfoliation methods are reviewed.

 Challenges in liquid‐ phase exfoliation and processing of graphene are reviewed.

 Concerns regarding the quality and structure of the produced graphene sheet with various methods are

discussed.

92
93

Das könnte Ihnen auch gefallen