Sunteți pe pagina 1din 47

Accepted Manuscript

Plasma-functionalized exfoliated multilayered graphene as cement reinforcement

Ma Shanlene D.C. Dela Vega, Magdaleno R. Vasquez, Jr.

PII: S1359-8368(18)32747-1
DOI: https://doi.org/10.1016/j.compositesb.2018.12.055
Reference: JCOMB 6384

To appear in: Composites Part B

Received Date: 23 August 2018


Revised Date: 29 November 2018
Accepted Date: 14 December 2018

Please cite this article as: Dela Vega MSDC, Vasquez Jr. MR, Plasma-functionalized exfoliated
multilayered graphene as cement reinforcement, Composites Part B (2019), doi: https://doi.org/10.1016/
j.compositesb.2018.12.055.

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.
ACCEPTED MANUSCRIPT

Plasma-functionalized exfoliated multilayered graphene


as cement reinforcement

PT
Ma. Shanlene D.C. Dela Vega, Magdaleno R. Vasquez Jr.∗
Department of Mining, Metallurgical, and Materials Engineering, College of Engineering,

RI
University of the Philippines, Diliman, Quezon City 1101, Philippines

SC
Abstract

U
A fast, facile, nonhazardous, environment-friendly, and high yield process

AN
was developed for the plasma treatment of graphite particles and the pro-
duction of plasma-functionalized multilayered graphene (pf-MLG). Graphite
particles (<20µm) were functionalized using a subatmospheric 13.56 MHz
M
radio frequency-excited oxygen plasma followed by liquid-phase exfoliation
to produce pf-MLG with a high aspect ratio (>2585) with <20 graphene
D

layers. The exfoliated graphene also exhibited high dispersibility in water


TE

after plasma functionalization without the use of surfactants. The pf-MLG


nanoflakes were incorporated into a cement mixture with 0.1 and 0.5 wt% pf-
EP

MLG loading. A 56% increase in compressive strength of cement mortars was


achieved for the 0.5 wt% pf-MLG after 28 days curing. This is attributed
to the strong interfacial interaction between graphene and cement matrix
C

and the promotion of hydration. The highly scalable process of pf-MLG-


AC

reinforced cement will make a positive impact on the environment, especially


in the construction industry.


Corresponding author
Email address: mrvasquez2@up.edu.ph (Magdaleno R. Vasquez Jr.)

Preprint submitted to Composites Part B: Engineering December 17, 2018


ACCEPTED MANUSCRIPT

Keywords: plasma functionalization, multilayered graphene, graphene


exfoliation, cement composite, nanoreinforcement

PT
1 1. Introduction

RI
2 Nanoscience is revolutionizing the way researchers approach materials in-
3 novation and design venturing further on utilizing nanomaterials for macroap-

SC
4 plications. This promises limitless possibilities turning novel ideas into tan-
5 gible practical applications. The feasibility of using advanced materials as

U
6 reinforcements for composites cannot be imagined until decades ago with the
7

8
AN
discovery of high-performance structural materials such as graphene-based
materials. Utilizing the unconventional behavior of a single to few atomic
9 layer-thick hexagonal lattice carbon, widely known as graphene, to improve
M
10 the performance of products for a wide spectrum of applications. Since the
11 discovery of the atomically thin carbon layers [1], studies have presented
D

12 properties superior to bulk, micro, and even other nanomaterials. Such great
TE

13 potential paved the way in capturing the interest of researchers to further


14 explore the applications of carbon-based nanomaterials [2].
EP

15 Graphene is a monolayer of carbon (C) atoms tightly packed into a two-


16 dimensional (2D) honeycomb lattice [3]. It is the basic building block for
17 other carbon-based structures such as graphite, carbon nanotubes (CNTs),
C

18 and fullerenes [4, 5, 6]. Due to its structure, graphene exhibits a very large
AC

19 surface area (2630 m2 /g) among the carbon-based and other materials while
20 remaining to be lightweight [6, 7, 8, 9]. It was also reported that pristine
21 graphene has a Young's modulus of 1 TPa and a strength of 130 GPa which
22 leads to high strength and high stiffness [4, 10, 11, 12, 13] as well as high ther-

2
ACCEPTED MANUSCRIPT

23 mal conductivity (∼5000 W/mK) [14]. Moreover, nanostructured graphene


24 exhibits superb electrical properties [15] such as fast charge carrier mobility

PT
25 (∼200000 cm2 V−1 s−1 ) [14]. These extraordinary properties made graphene a
26 promising material for a wide variety of applications [2] ranging from energy

RI
27 storage [16, 17], electronics and semiconductor devices [3, 18], coatings and
28 barriers [19, 20, 21], fillers and reinforcements in metal-based [22], polymer-

SC
29 based [23, 24], and ceramic-based [25, 26, 27] composites.
30 Among the practical applications of graphene and other carbon-based

U
31 nanomaterials that is heavily explored is reinforcing composite materials.
For instance, CNTs have been used for strengthening metal matrix compos-
32

33
AN
ites [28]. Graphene and graphene oxide (GO) have been used to increase
34 mechanical properties of titanium [29] and aluminum matrices [30]. Tar-
M
35 geted properties of polymeric matrices were also enhanced upon the addition
36 of graphene [31, 32, 33]. Moreover, graphene and GO nanomaterials have
D

37 also been heavily studied for inclusion in cement reinforcement [34, 35, 36].
TE

38 To realize the applicability of graphene for cement reinforcement, the nano-


39 material should be produced in large quantities. Moreover, dispersion of
graphene in any media is critical in providing a large surface area for hydra-
EP

40

41 tion to maximize the interactions with the matrix [8].


42 While the quest for different practical applications of graphene continues,
C

43 progress in the production of graphene with specific properties has challenged


AC

44 researchers worldwide. Among the challenges is the realization of large-scale


45 production of graphene-based materials. Novoselov [2] and Raccichini [17]
46 illustrated synthesis routes of graphene that include exfoliation [1], reduction
47 of graphene oxide (rGO) [37, 38], and chemical vapor deposition (CVD) [39].

3
ACCEPTED MANUSCRIPT

48 To realize the industrial application of graphene, especially for the construc-


49 tion industry, synthesis methods must be cost-effective, scalable, and have

PT
50 high yield. rGO synthesis methods are capable of mass-producing graphene
51 but concerns on the use of harsh chemicals limit its industrial application

RI
52 due to health and environmental concerns [37]. CVD-grown graphene, on the
53 other hand, is of high quality but requires high temperature and high vacuum

SC
54 systems that limits its scalability for high-throughput applications. Thus,
55 among the three, the top-down approach through exfoliation, whether me-

U
56 chanical or solvent-based, is promising for large-scale synthesis of graphene.
Mechanical exfoliation of thin layers of carbon has been achieved through
57

58
AN
different means. Novoselov et al. [1] were able to extract stable graphene
59 films from graphite using adhesive tape. Jayasena and Subbiah [40] pro-
M
60 duced graphene via mechanical cleavage while Paton et al. [41] employed
61 shear forces to remove layers of graphene from graphite. Liquid-phase ex-
D

62 foliation (LPE) through sonication has also been proposed for large-scale
TE

63 synthesis of graphene. The addition of N-methyl-pyrrolidone [42, 43] al-


64 lowed the production of graphene flakes from graphite powders. Nuvoli et al.
[44] utilized an ionic liquid to obtain graphene in high concentrations that
EP

65

66 does not require chemical modification. Exfoliation-based synthesis route will


67 continue to be at the forefront of mass-producing graphene as long as the
C

68 use of harsh chemicals is minimized. Hence, alternative means are continu-


AC

69 ously being sought to reduce environmental and health hazards in large-scale


70 graphene synthesis.
71 Another challenge that hinders the applicability of graphene-based ma-
72 terials as fillers is their dispersion. Since the alkaline environment of cement

4
ACCEPTED MANUSCRIPT

73 restricts the effect of graphene reinforcement, an additional surface treat-


74 ment is necessary to improve its dispersibility especially in concrete compos-

PT
75 ites. For instance, polycarboxylate superplasticizer can be used to modify
76 GO and graphene nanoplatelets to enhance its dispersion and subsequently

RI
77 improve the mechanical behavior of cement composites [45, 46]. Sodium
78 cholate-modified graphene was also used to strengthen concrete composites

SC
79 [47]. While silica fume was used to improve graphene dispersion to en-
80 hance mechanical and electrical properties of graphene-cement composites

U
81 [48]. Clearly, surface modification of graphene-based filler materials is es-
sential for enhancing dispersability such as in cement reinforcement applica-
82

83
AN
tions. While most methods used wet-based approaches to modify graphene,
84 dry-based methods have not yet been fully explored in order to form a ho-
M
85 mogeneous and stable dispersion without the aid of surfactants and complex
86 solvents [49]. Covalently attaching oxygen-containing functional groups to
D

87 modify the surface properties of graphene can be done through different


TE

88 treatments such as chemical [50], thermal [51], ozone [52], or plasma oxida-
89 tion [49, 53, 54]. Among these approaches, plasma oxidation is a suitable
method for efficient functionalization wherein excited species, radicals, elec-
EP

90

91 trons, ions, and ultraviolet (UV) light contained in the plasma discharge
92 interact with the material surface. Specifically for graphene-based materials,
C

93 the C bonds are broken which creates active sites that interacts with reactive
AC

94 oxygen (O2 ) species from the discharge [53]. Subatmospheric plasma treat-
95 ment of graphene has been demonstrated in a related work of Sales et al.
96 [55] where nitrogen plasma was used to modulate the electrical properties of
97 spray-deposited exfoliated graphene. Indeed, the plasma-based modification

5
ACCEPTED MANUSCRIPT

98 of graphene is beneficial since it does not use harsh chemicals and has no
99 toxic byproducts.

PT
100 In this study, functionalization and production of multilayered graphene
101 (MLG) were performed via a two-step method: plasma functionalization and

RI
102 LPE. The plasma-functionalized MLG (pf-MLG) was incorporated in cement
103 composites and their effects on the compressive strength were evaluated. This

SC
104 paper aims to provide a preliminary understanding regarding the effect and
105 applicability of using pf-MLG as nanoreinforcement for cement composites.

U
106 This work will also pave the way for a facile and non-toxic approach for
large-scale production of MLG for macroapplications.
107
AN
108 2. Materials and Methods
M
109 2.1. Oxygen plasma functionalization

As-received graphite flakes (<20µm, C, Sigma-Aldrich 282863) were plas-


D

110

111 ma-treated using a 13.56 MHz radio-frequency (RF) capacitively-coupled


TE

112 plasma (CCP) system inside a glass bell jar chamber. The system can be op-
113 erated under subatmospheric conditions and described in detail by Cagomoc
EP

114 and Vasquez [56]. The functionalization of graphite was conducted using the
115 best settings predetermined from the results of previous studies [57]. In a
typical experiment, 0.35 g of graphite was treated under O2 plasma using 50
C

116

117 W RF power, 15 min exposure time, and at 100 Pa working pressure. For
AC

118 one batch, plasma treatment was conducted three (3) times interspersed with
119 homogenization of the graphite powders through mixing in between runs to
120 ensure even functionalization of the graphite particles.

6
ACCEPTED MANUSCRIPT

121 2.2. Liquid-phase exfoliation


122 The plasma-functionalized graphite (pf-graphite) particles were used as

PT
123 the precursor for the top-down approach for the graphene synthesis via LPE.
124 Time-resolved studies were done prior to LPE where at least 5 h is needed

RI
125 to produce a stable graphene dispersion. The dispersion was prepared with
126 initial graphite concentration of ∼0.25 mg/mL (250 ppm) in deionized wa-

SC
127 ter (DI H2 O) and sonicated (Branson M1800H-E) at 40 kHz for 5 h. Prior
128 to use, the resulting pf-graphene dispersion was further diluted in H2 O to

U
129 achieve the desired graphene reinforcement concentration to produce the ce-
ment composite.
130

131 2.3. Characterization of pf-MLG


AN
M
132 As discussed in the previous sections, LPE was the chosen method in pro-
133 ducing graphene from a graphite particle precursor. Initial characterizations
D

134 were performed to confirm the effectiveness of the top-down approach as well
as the process parameters used. Surface morphology and structure of the
TE

135

136 flakes before and after LPE were examined under a field-emission scanning
137 electron microscope (FESEM) (Helios Nanolab 600i). The FESEM samples
EP

138 were prepared by depositing a drop (5 µL) of graphite and the exfoliated
139 graphene dispersion on silicon (Si) wafer substrates. The lateral dimension
C

140 of the flakes were obtained via dynamic light scattering (DLS) (NanoPlus)
technique. Samples examined using the DLS were graphite flakes dispersed in
AC

141

142 pentanol and exfoliated MLG dispersed in ethanol. Ultraviolet-visible (UV-


143 vis) spectroscopy (Shimadzu UV-1800) in photometric mode was utilized to
144 confirm the presence of graphitic carbon and determine the graphene yield
145 after LPE. A calibration curve was obtained using graphene dispersions of

7
ACCEPTED MANUSCRIPT

146 known concentration (10 to 100 ppm, C, Sigma-Aldrich 799092) as standard


147 dispersions. Raman spectroscopy at 532 nm laser excitation with 1 µm di-

PT
148 ameter spot size (Horiba LabRAM HR Evolution) of the untreated MLG and
149 pf-MLG were obtained to evaluate the quality of the graphene produced and

RI
150 the effect of O2 plasma treatment on the graphene structure. Parametric
151 studies were done to determine the plasma parameters suitable to achieve

SC
152 functionalization.
153 Powder diffraction technique was performed on the MLG nanoflakes and

U
154 pf-MLG to analyze the phase composition using high-resolution x-ray diffrac-
tion (XRD) (Rigaku RINT2100 CMJ). Surface morphology and structure of
155

156
AN
the exfoliated graphene before and after plasma functionalization were ex-
157 amined under a high-resolution transmission electron microscope (HRTEM)
M
158 (JEM-2100F). Around 1 µL of the exfoliated graphene and pf-graphene were
159 deposited on a holey copper (Cu) TEM substrate.
D

160 2.4. Preparation of cement mortars


TE

161 The pf-MLG were used in the preparation of the MLG-cement compos-
162 ites as shown in Fig. 1. The MLG-cement composites were prepared in
EP

163 compliance with the standard preparation method for the compression test
164 specimens given by ASTM C109/109M-16a [58]. The MLG-cement ratio and
C

165 curing time were varied to determine the reinforcing effect of pf-MLG on the
compressive strength and the degree of hydration of the cement mortars.
AC

166

167 In a typical mixture, the mortars were composed of 1 part Portland ce-
168 ment per 2.75 parts sand, 0.68 v/w water-to-cement ratio, and varying con-
169 centrations of pf-MLG to realize a 0.1 and 0.5 wt% loading of graphene with
170 respect to cement. The dry components were mixed prior to adding the pf-

8
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 1: Preparation of graphene-cement composite. (a) 50-mm3 cement mortar, (b)
specimen mold, (c) pf-graphene in H2 O, (d) Portland cement, and (e) sand.
M
171 MLG in H2 O dispersion in the cement mixture. The components were mixed
thoroughly until a consistent mixture was achieved. Then, the mixture was
D

172

173 transferred to a custom-built 50-mm3 mold and was allowed to harden for
TE

174 24 h before demolding. To fully promote hydration in the cement mixture


175 and maintain the moisture content of the cement mortars, the samples were
EP

176 submerged in tap H2 O all throughout the curing period. Finally, the cement
177 mortars reinforced with pf-MLG were dried 24 h prior to compression test.
C

178 2.5. Mechanical testing of graphene-cement mortars


AC

179 Compressive strength of the MLG-cement composites were determined


180 conforming to specifications given by the Standard Test Method for Compres-
181 sive Strength of Hydraulic Cement Mortars [58] using the universal testing
182 machine (UTM) (NL Scientific CL-03). The effect of pf-MLG was evaluated

9
ACCEPTED MANUSCRIPT

183 by measuring the compressive loads and compressive stresses that can be ap-
184 plied on the cement mortars. The values obtained from the control specimen

PT
185 (0.0 wt% graphene) was used as basis of comparison with the results from
186 the 0.1 and 0.5 wt% loading of the cement mortars. The fractured surface

RI
187 of the cement mortars with 0.0 and 0.5 wt% reinforcement were also exam-
188 ined under the FESEM to investigate the morphology of the cracks along the

SC
189 fractured surface.

3. Results

U
190

3.1. Production of pf-MLG nanoflakes


191

192
AN
Pf-MLG nanoflakes were produced following a two-stage process – (1) O2
193 plasma functionalization succeeded by (2) LPE. The effects of the plasma
M
194 treatment and other process parameters are discussed in the succeeding sub-
195 sections.
D

196 3.1.1. Effect of O2 plasma functionalization


TE

197 As-received graphite flakes were subjected to O2 plasma treatment. The


198 pf-graphite was dispersed in DI H2 O prior to exfoliation of graphene via
EP

199 sonication. Figure 2 shows images of exfoliated graphene from as-received


200 graphite and O2 plasma-treated graphite prior to exfoliation. For the un-
C

201 treated sample, agglomeration occurred immediately after sonication. This


signifies the stronger affinity between graphene flakes than with the medium
AC

202

203 leading to instability of the dispersion. This effect is not desirable in the pro-
204 posed application of graphene as cement reinforcement. However, stability of
205 the graphene suspension was achieved when the graphite particles were pre-
206 treated with O2 discharge. The improved dispersability is attributed to the

10
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
Figure 2: Dispersion of exfoliated graphene after 5 h sonication: (a) graphene from as-
received graphite and (b) graphene from O2 plasma-treated graphite.
D

207 increase in affinity of graphene to H2 O due to the O2 functional groups at-


TE

208 tached to its structure which minimizes the agglomeration of the nanoflakes.
209 The graphite particles were spray-deposited on glass substrates and the Ses-
sile drop technique was used to determine the surface free energy (SFE).
EP

210

211 Calculations using the Owens-Wendt-Kaelble [59, 60] approach showed an


212 increase in SFE from 41 mJ m−2 for the untreated graphite to 76 mJ m−2
C

213 for the O2 plasma-treated graphite signifying an increase in hydrophilicity.


AC

214 After O2 plasma treatment a shift in the XRD peak was observed (Fig. 3)
215 signifying an increase in interlayer spacing of graphene from 0.345 nm to
216 0.347 nm for the untreated and pf-graphite, respectively. This increase in
217 interlayer spacing can be attributed to the oxidation of graphene [15].

11
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 3: Peak position shift of (002) of exfoliated graphite (a) before and (b) after O2
plasma treatment.
M
218 3.1.2. Liquid-phase exfoliation
219 The efficiency of the top-down approach was investigated by performing
D

220 preliminary characterization on the graphite precursor and exfoliated MLG.


TE

221 Surface morphology of the graphite precursor and exfoliated MLG after 5 h
222 of sonication were examined under FESEM as shown in Figs. 4 and 5. The
EP

223 graphite flakes are of different sizes as seen in the FESEM images (Fig. 4).
224 Based on the initial lateral measurements done using ImageJ [61], the size of
the flakes were confirmed to be less than 20 µm. Dynamic light scattering
C

225

226 (DLS) was employed to obtain a bulk distribution for the size of the graphite
AC

227 flakes. A bimodal distribution was observed and average particle sizes of
228 13.11 µm and 665.43 µm were obtained as shown in Fig. 6(a). The large
229 particle size distribution (665.43 µm) measured in the DLS data is due to the
230 agglomeration of the graphite flakes. The smaller particle size distribution

12
ACCEPTED MANUSCRIPT

231 can be attributed to the unexfoliated graphite particles. Moreover, for non-
232 spherical particles like MLG nanoflakes, the particle size measured is assumed

PT
233 to pertain to the lateral dimensions of the flakes [62]. Similarly, FESEM and
234 DLS performed on the exfoliated MLG as shown in Fig. 5 and Fig. 6(b).

RI
235 MLG is defined as “a sheet-like material consisting of a small number of
236 well-defined, countable, stacked graphene layers” [63]. In this paper, MLG

SC
237 refers to exfoliated stacked graphene less than 20 layers. Comparing the
238 lateral dimensions of the flakes before (Fig. 4) and after (Fig. 5) exfoliation,

U
239 sonication resulted to a decrease in lateral dimension. Although, sonication
can reduce the lateral dimension of the flakes, it was confirmed from the DLS
240

241
AN
distribution (Fig. 6(b)) the presence of flakes with large lateral dimensions
242 (15.69 µm) even after exfoliation. Moreover, from the tilted view (Fig. 5(c))
M
243 of the exfoliated MLG, the layer thickness was measured using ImageJ [61]
244 and an average thickness of 6.07 nm was obtained. Using the interlayer
D

245 spacing (0.345 nm) obtained from the XRD pattern reported in [57], we
TE

246 can estimate the MLG to have 17 layers. Chuah et al. [8] reported that
247 monolayer graphene has an aspect ratio (i.e. lateral dimension over layer
thickness) greater than 6000, GO has an aspect ratio greater than 1500,
EP

248

249 and CNTs have an aspect ratio greater than 1000. The aspect ratio of the
250 exfoliated MLG produced in this study was calculated to be greater than
C

251 2585 which is relatively higher than that of GO and CNTs. A high aspect
AC

252 ratio is desirable to achieve a large surface area that will act as hydration
253 sites when used as nanoreinforcement in cement [8, 64]. A calibration curve
254 was obtained by subjecting graphene dispersions of known concentration in
255 UV-vis spectroscopy via photometric method. Graphene is known to have an

13
ACCEPTED MANUSCRIPT

256 absorbance peak between 265 to 300 nm wavelength (λ). This corresponds
257 to the π-π∗ transitions in which the shift in wavelength accounts for the

PT
258 length of the C bonds present in the graphene structure [65]. In this study,
259 the graphene dispersion shows an absorbance peak at λ = 266.1 nm. Using

RI
260 Eq. 1, an exfoliation yield equation derived from the linear fit of the plot of
261 the graphene concentration against absorbance (Fig. 7). The graphene yield

SC
262 concentration, with A = 0.37486 at λ = 266.1 nm, is calculated to be 23.53
263 ppm. This represents 9.41% graphene yield after 5 h sonication.

U
A = 0.021c − 0.1192 (1)

264
AN
where A is the absorbance at λ = 266 nm and c is the graphene yield con-
265 centration.
M
266 The normalized Raman spectra of the MLG (Fig. 8(a)) and pf-MLG
267 (Fig. 8(b)) give an indirect correlation on the exfoliated MLG and effect
D

268 of O2 plasma functionalization to the structure of pf-MLG. Both Raman


TE

269 spectra exhibited graphene (G) and defect (D, 2D, D’, and D”) peaks typical
270 for pristine graphene. The D-, G- and 2D-peaks were observed at ∼1350,
EP

271 ∼1585, and ∼2700 cm−1 , respectively. The G-peak corresponds to the sp2
272 carbon structure while 2D peak is always present and no defects are needed
for its activation. The D-peak is due to breathing modes of six atom rings
C

273

274 and is activated with the presence of defects. Emergence of these defects
AC

275 depends on the edge orientation of the graphene structure, whether zig-zag
276 or armchair, structural damage from sp2 to sp3 hybridization, attachment of
277 functional groups, or oxidation. Lastly, presence of resonant peaks identified
278 as D’- and D+D”- peaks located at ∼1620 and ∼2450 cm−1 , respectively,

14
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 4: FESEM images of graphite particles prior to exfoliation at (a) 1kx, (b) 5kx, and
(c) 10kx magnification.

15
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 5: FESEM images of exfoliated graphene flakes at (a) 10kx, (b) 25kx, and (c) 100kx
magnification.

16
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 6: Intensity distribution from DLS results of (a) graphite and (b) exfoliated MLG.

17
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 7: Calibration curve of the graphene exfoliation yield.

279 were observed [66].


M
280 For the untreated MLG nanoflakes (Fig. 8(a)), the D-peak can be at-
281 tributed to the inherent edge structure of graphene and presence of some
D

282 hydroxyl groups due to the sp3 hybridized structure. Intercalation with H2 O
TE

283 can also be a source of the hydroxyl groups. Plasma oxidation introduces
284 disorder in the carbon backbone as observed in Fig. 8(b). The increase in
D-peak is due to damages on the sp2 carbon structure as a result of the at-
EP

285

286 tachment of O2 functional groups [66, 67]. Treating the MLG with O2 plasma
287 lead to an 8.63% increase in defect peak intensity. Higher defect peaks were
C

288 observed and resonant peaks, D’ and D+D”, were more pronounced after
AC

289 plasma functionalization [66, 68].


290 A summary of the intensity peak ratios of the exfoliated MLG and pf-
291 MLG, along with the estimated number of layers derived from the Raman
292 spectra, is presented in Table 1. The number of graphene layers were calcu-

18
ACCEPTED MANUSCRIPT

PT
RI
U SC
Figure 8: Raman spectra of the exfoliated (a) MLG and (b) pf-MLG.
AN
Table 1: Raman spectra peak ratios and the calculated number of stacked layers.

Graphene ID /IG IG /I2D n layers


M
(a) MLG 0.197 1.98 18
D

(b) pf-MLG 0.214 2.15 20


TE

293 lated according to the established formula given in Eq. 2 [68].


EP

IG n
= 0.14 + (2)
I2D 10
C

294 where IG and I2D are the peak intensities of the G- and 2D-peak, respectively,
and n is the number of graphene layers.
AC

295

296 The HRTEM images obtained (Figs. 9 and 10) shows the different lay-
297 ered graphene produced after exfoliation of graphite and pf-graphite flakes.
298 As described earlier in Fig. 2, using pure H2 O as graphene interacalation
299 compound (GIC) in the untreated graphite was not suitable to produce a

19
ACCEPTED MANUSCRIPT

300 stable dispersion. Thus, ethanol was added at 50% concentration. However,
301 H2 O was solely used during exfoliation of the pf-graphite. From the HRTEM

PT
302 images, unexfoliated graphite or sediments were observed after exfoliation as
303 shown in Fig. 9(a). Both exfoliation of graphite and pf-graphite resulted to

RI
304 producing few-layered graphene (FLG) (Fig. 9(c) and Fig. 10(b)) and MLG
305 (Fig. 9(b) and Fig. 10(a)) with MLG being more abundant in the yield.

SC
306 3.2. Pf-MLG-cement composite

307 The cement mortars were subjected to compression test to investigate the

U
308 effect of varying graphene concentration and curing time on its compression
309
AN
strength. The compressive load reported in this study was obtained from the
310 average compressive force that was applied on the cement specimens dur-
M
311 ing the compression test. Their corresponding average compressive strength
312 which is the average maximum allowable compressive stress experienced by
D

313 the cement mortars. The compressive strength (σ) of each sample was cal-
culated using basic concepts in engineering mechanics as expressed in Eq. 3.
TE

314

F
σ= (3)
Across-section
EP

315 where F is the applied load while Across-section is the cross-sectional area of
316 the specimen.
C

3.2.1. Effect of Pf-MLG concentration


AC

317

318 Previous studies have reported positive effects of carbon-based materials


319 as reinforcement such as the works of Gong et al. [4], Cwirzen et al. [69], Li
320 et al. [70], Kumar et al. [71], and Pan et al. [72]. Figure 11 shows the compar-
321 ison of the compressive load and compressive strength of the cement mortars

20
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 9: HRTEM images after LPE of graphite yielding (a) unexfoliated graphite, (b)
MLG, and (c) FLG.

21
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP

Figure 10: HRTEM images after LPE of pf-graphite yielding (a) pf-MLG and (b) pf-FLG.
AC

22
ACCEPTED MANUSCRIPT

PT
RI
U SC
Figure 11: Plot of (a) compressive load and (b) compressive strength of cement mortars
AN
with varying pf-MLG concentration after curing for 7 days.

with varying graphene loading after 7 days curing. A 7.06% increase in com-
M
322

323 pressive load which corresponds to 4.65% increase in compressive strength


was observed with the addition of 0.1 wt% pf-MLG. On the other hand, com-
D

324

325 pared to the control specimen, the compressive load increased by 52.33% and
TE

326 compressive strength by 56.07% upon the addition 0.5 wt% pf-MLG after 7
327 days of curing.
EP

328 Similarly, the comparison between the average compressive load and av-
329 erage compressive strength of cement mortars with varying pf-MLG concen-
tration after 28 days curing is presented in Fig. 12. A 21.88% increase in
C

330

331 compressive load and 18.85% increase in compressive strength observed with
AC

332 the addition of 0.1 wt% pf-MLG while the compressive load increased by
333 61.48% and compressive strength by 55.88% upon the addition of 0.5 wt%
334 pf-MLG.
335 Surface morphologies of the 0.0 and 0.5 wt% pf-MLG were examined

23
ACCEPTED MANUSCRIPT

PT
RI
U SC
Figure 12: Plot of (a) compressive load and (b) compressive strength of cement mortars
AN
with varying pf-MLG concentration after curing for 28 days.

under FESEM. Figure 14 shows the microstructure of unreinforced cement


M
336

337 mortar. Voids and microcracks were more evident in the unreinforced ce-
ment mortar (Fig. 14(a)) exhibiting a porous material. Formation of hy-
D

338

339 dration products such as Ettringite (a rod-like structure) and monosulfate


TE

340 hydrates (flower-like structure) were observed in Fig. 14(b) and (c). The
341 microstructure of cement mortar reinforced with 0.5 wt% pf-MLG is shown
EP

342 in Fig. 15. The distribution of pf-MLG nanoflakes (black specks) can be
343 seen in Fig. 15(a). Figure 15(b) shows a more compact structure with less
microcracks that has been achieved after addition of 0.5 wt% pf-MLG. An
C

344

345 isolated image of the pf-MLG reinforced in the cement matrix is shown in
AC

346 Fig. 15(c).

24
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 13: Plot of compressive load and compressive strength of cement mortars with (a,
b) 0.0 (c, d) 0.1 and (e, f) 0.5 wt% pf-MLG concentration at different curing time.

25
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 14: FESEM images of unreinforced cement mortar showing (a) the fracture surface,
(b) pores and voids, and (c) microstructure of the hydration products.

26
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 15: SEM images of cement mortar reinforced with 0.5 wt% pf-MLG showing (a) the
fracture surface, (b) pf-MLG in cement matrix, and (c) inset of the pf-MLG reinforcement.

27
ACCEPTED MANUSCRIPT

347 3.2.2. Effect of curing time


348 Cement specimens were subjected to compression test after 7 days of

PT
349 curing to investigate the effect of reinforcement during the early stages of
350 curing and after 28 days following the established minimum curing period of

RI
351 cement in actual structural applications. Comparison between the cement
352 mortar after 7 and 28 days, respectively, were done by determining the per-

SC
353 cent difference in terms of compressive load and compressive strength. Fig-
354 ure 13(a) shows that the control specimen resulted to a 15.11% and 16.51%

U
355 increase in compressive load and compressive strength, respectively. For the
0.1 wt% pf-MLG reinforced mortar shown in Fig. 13(b), the compressive load
356

357
AN
and compressive strength also increased to 31.01% and 32.37%, respectively.
358 Lastly, addition of 0.5 wt% pf-MLG on the cement mortar lead to an increase
M
359 of 22.02% and 16.39% in compressive load and compressive strength, respec-
360 tively, as shown in Fig. 13(c). Similar trends were observed for the control
D

361 and reinforced specimens, however, the cement mortar reinforced with 0.1
TE

362 wt% pf-MLG showed the largest percent increase in compressive load and
363 compressive strength during the span of 7 to 28 days curing time. These
large increase in values observed in the 0.1 wt% reinforced cement mortars
EP

364

365 can be correlated to a higher degree of hydration compared to the control


366 and 0.5 wt% reinforced specimens. Moreover, the percentages observed from
C

367 the 0.5 wt% reinforced cement mortars is relatively smaller than the 0.1 wt%
AC

368 reinforced specimens which suggests probable onset of saturation with the
369 addition of ≥0.5 wt% pf-MLG.

28
ACCEPTED MANUSCRIPT

PT
RI
Table 2: Comparison of the compressive strength in cement nanocomposites reinforced
with graphene-based materials.

Water- Compressive

SC
Reference Matrix cement Type of Concentration strength
ratio reinforcement (wt%) increase (%)

U
(w/c)

Li et al. Mortar AN
0.45 CNTs 0.50 19 (28 days)
(2005) [70]
M
Kumar et al. Paste 0.40 CNTs 0.50 15 (28 days)
(2012) [71]
D

Cwirzen et al. Paste 0.40 CNTs 0.45 50


(2008) [69]
TE

Gong et al. Paste 0.50 GO 0.03 46 (28 days)


(2015) [4]
EP

Pan et al. Mortar 0.50 GO 0.05 15-33 (28 days)


(2015) [72]
C

This work Mortar 0.68 pf-MLG 0.50 56 (28 days)


AC

29
ACCEPTED MANUSCRIPT

370 4. Discussion

Owing to the high total surface area of graphene, more area is exposed

PT
371

372 that can become active sites for chemical or physical interactions. Attaching
373 functional groups on the graphene surface alters the van der Waals forces

RI
374 present between graphene layers to improve bonding with the matrix and
its dispersibility in H2 O [72]. Chen et al. [53] proposed reaction mechanisms

SC
375

376 that can take place upon interaction of the plasma species and the surface
377 of graphene-based materials. Functionalization through O2 plasma can gen-

U
378 erate C – O, C – O, and O – C – O bonds. The active π bonds in C – C bonds
379

380
AN
dissociate upon interaction with the species in the plasma discharge gen-
erating radicals. These creates active sites for the functionalization where
reactive oxygen atoms attach to the structure of graphene. The breaking of
M
381

382 the C – C bonds was manifested as defects in the Raman spectra obtained
(Fig. 8) and Table 1 wherein an increase in the ID /IG ratio was observed.
D

383

384 Based on the studies reported by Chen et al. [53], a schematic of the func-
TE

385 tionalization mechanism is proposed as shown in Fig. 16(a).


386 Intercalation mechanism between the graphene layers and GICs have been
EP

387 discussed in the works of Stankovich et al. [15] and Dela Vega [57] among
388 others. In the case for pf-MLG (Fig. 16(b)), a stable dispersion can be
produced in pure H2 O without any aid of surfactants or dispersants (e.g.
C

389

390 alchohols, ketones, or organic solvents). The simplicity and non-toxicity of


AC

391 this process makes it convenient aside from the several benefits gained in
392 using pure H2 O. Sonication leads to the fracture of thick and large flakes
393 to smaller and thinner stacked layers of graphene. Due to the shear forces
394 and cavitation produced during exfoliation and the weak Van der Waals

30
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

31

Figure 16: Schematic diagram of proposed mechanism for (a) plasma functionalization,
(b) liquid-phase exfoliation, and (c) pf-MLG and cement hydrate reaction.
ACCEPTED MANUSCRIPT

395 forces between layers, the outer layers are displaced (i.e. tilts) creating small
396 gaps between layers. The H2 O molecule has a 0.28 nm diameter while the

PT
397 interlayer spacing between graphene layers is 0.34 nm. Thus, H2 O as an
398 intercalation compound can cleave graphene layers with the help of agitation

RI
399 from the sound waves propagated during exfoliation [15, 73, 74].
400 Two general mechanisms explain the observed enhancement in mechani-

SC
401 cal properties of the cement composite upon addition of nanoreinforcements
402 [72]. A schematic diagram of the reaction mechanism of pf-MLG and cement

U
403 hydrates is presented in Fig. 16(c). A material having high aspect ratio,
in this case pf-MLG (aspect ratio >2585) can act as nucleation site of the
404

405
AN
cement phase and aid in cement hydration. Moreover, with the comparable
406 size of graphene to that of the capillary and gel pores in the matrix, it can
M
407 act as nanofiller which densifies the microstructure. This is specially true for
408 the exfoliated pf-MLG nanoflakes with smaller lateral dimensions and fewer
D

409 layers. At lower pf-MLG concentration (0.1 wt%), there was a 31.01 and
TE

410 32.37% increase in compressive load and compressive strength, respectively,


411 from 7 to 28 days curing time suggesting the strengthening mechanism to
be mainly governed by the higher degree of hydration due to the increase in
EP

412

413 nucleation sites on the pf-MLG surface. At higher pf-MLG concentration,


414 (0.5 wt%), the increase in compressive load and compressive strength from
C

415 7 to 28 days curing time was relatively lower at 22.02 and 16.39%, respec-
AC

416 tively, suggesting possible onset of nucleation saturation. Nevertheless, the


417 samples with the higher pf-MLG concentration had the largest increase in
418 compressive load and compressive strength at 61.48 and 55.88%, respectively.
419 There can be a saturation in terms of hydration nuclei, however, the excess

32
ACCEPTED MANUSCRIPT

420 pf-MLG may have also acted as nanofillers which lead to the densification
421 of the microstructure of the cement and, in turn, resulted to a relatively

PT
422 higher compressive strength. These pf-MLG can also assist in crack arrest
423 up to microscale. Thus, for the 0.5 wt% reinforced cement composites, the

RI
424 strengthening mechanism can be both governed by hydration kinetics and
425 densification of the microstructure.

SC
426 Previous works on reinforcing cement using carbon-based nanomaterials
427 have been reported. For instance, inclusion of GO and CNT [4, 70, 71, 69, 72]

U
428 have demonstrated the increase in compressive strength of cement (Table 2).
However, macroapplications of these nanomaterials, such as cement reinforce-
429

430
AN
ment, require huge amounts of GOs and CNTs. In the case of GO, Hummers
431 method [37] is the conventional technique to produce large amounts of GO.
M
432 However, the process poses major challenges such as the use of toxic chemicals
433 and production of harmful byproducts. Cement reinforcement using CNTs,
D

434 on the other hand, also poses challenges in terms of compatibility with ce-
TE

435 ment hydrates. Thus, an additional step is needed where CNTs are modified
436 with surfactants to improve its compatibility [75]. Moreover, graphene has
the highest total surface area [8], thus pf-MLG is ideal for cement hydra-
EP

437

438 tion. Changes in the workability of the cement was also observed depending
439 on the water-cement ratios. It is important to account for the increase in
C

440 specific surface area of all the components in the cement dry mixture (i.e.
AC

441 incorporation of reinforcement or fillers) since water deficiency in the cement


442 mixture reduces workability and has a direct effect on the strength of the
443 cement structure [4, 8, 76].
444 The results of this work demonstrated the feasibility of engineering ce-

33
ACCEPTED MANUSCRIPT

445 ment with nanomaterials leading to increase in strength. The inclusion of


446 graphene will be of high interest in the construction industry since it will

PT
447 open up the development and use of the cement composites for different
448 applications. We have also demonstrated the use of facile techniques of pro-

RI
449 ducing functionalized graphene that is fast, safe, environment-friendly, non-
450 hazardous, and more importantly high scalability. Finally, The construction

SC
451 industry plays a critical role in the economic development of any nation.
452 However, the industry is also one of the most significant contributors to en-

U
453 vironmental degradation and carbon dioxide emission [77, 78, 79]. Hence,
the use of pf-MLG-reinforced cement will make a positive impact in the en-
454

455
AN
vironment especially in the construction industry.
M
456 5. Conclusion

457 A two-stage process for producing pf-MLG was developed intended for
D

458 large-scale applications such as in composites for structural and building


TE

459 materials. This work demonstrated the production of pf-MLG nanoflakes


460 with less than 20 graphene layers and high aspect ratio (>2585) and produced
EP

461 a surfactant-free graphene dispersion in water. This was achieved via pre-
462 treatment of graphite precursors using subatmospheric O2 plasma followed
by LPE. This work also demonstrated the feasibility of using the exfoliated
C

463

464 pf-MLG as cement reinforcement which increased its compressive strength


AC

465 by at least 56% at 0.5 wt% pf-MLG loading. The current process paves the
466 way of realizing a large-scale production of functionalized MLG that finds
467 applications in the building and construction industry. The process is fast,
468 efficient, environment-friendly, and scalable making nanomaterials applicable

34
ACCEPTED MANUSCRIPT

469 for macroapplications such as in the construction industry. Moreover, this


470 study can be extended to investigating the graphene-cement interaction as

PT
471 well as utilizing the unique properties of graphene in cement composites.

RI
472 Acknowledgement

473 The authors acknowledge the support from the University of the Philip-

SC
474 pines Diliman Office of the Vice-Chancellor for Research and Development
475 Outright Research Grant (OVCRD Project No. 141404 PNSE). M. Vasquez

U
476 is grateful for the support from the University of the Philippines Diliman
477

478
AN
Office of the Vice-President for Academic Affairs Balik-PhD Research Grant
(OVPAA-BPhD-2014-01) and the Jardiolin Family Professorial Chair. The
479 authors also acknowledge the nanoQuench project (CHED PCARI Project
M
480 No. IIID 2016-007) for the use of the Raman Spectrometer.
D

481 References
TE

482 [1] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V.


483 Dubonos, et al., Electric field effect in atomically thin carbon films,
EP

484 Science 306 (2004) 666–669.

485 [2] K. S. Novoselov, V. I. Fal’ko, L. Colombo, P. R. Gellert, M. G. Schwab,


C

486 K. Kim, A roadmap for graphene, Nature 490 (2012) 192–200.


AC

487 [3] A. K. Geim, K. S. Novoselov, The rise of graphene, Nat. Mater. 6 (2007)
488 183–191.

35
ACCEPTED MANUSCRIPT

489 [4] K. Gong, Z. Pan, A. H. Korayem, L. Qiu, D. Li, F. Collins, et al., Re-
490 inforcing effects of graphene oxide on portland cement paste, J. Mater.

PT
491 Civ. Eng. 27 (2015) 16.

[5] F. M. Koehler, W. J. Stark, Organic synthesis on graphene, Acc. Chem.

RI
492

493 Res. 46 (2013) 2297–2306.

SC
494 [6] L. Dai, Functionalization of graphene for efficient energy conversion and
495 storage, Acc. Chem. Res. 46 (2013) 31–42.

U
496 [7] M. D. Stoller, S. Park, Y. Zhu, J. An, R. S. Ruoff, Graphene-based
497
AN
ultracapacitors, Nano Lett. 8 (2008) 3498–3502.

498 [8] S. Chuah, Z. Pan, J. G. Sanjayan, C. M. Wang, W. H. Duan, Nano


M
499 reinforced cement and concrete composites and new perspective from
500 graphene oxide, Constr. Build. Mater. 73 (2014) 113–124.
D

501 [9] J. J. Yoo, K. Balakrishnan, J. Huang, V. Meunier, B. G. Sumpter,


TE

502 A. Srivastava, et al., Ultrathin planar graphene supercapacitors, Nano


503 Lett. 11 (2011) 1423–1427.
EP

504 [10] C. Lee, X. Wei, J. W. Kysar, J. Hone, Measurement of the elastic


505 properties and intrinsic strength of monolayer graphene, Science 321
C

506 (2008) 385–388.


AC

507 [11] W. C. Lee, C. H. Y. X. Lim, H. Shi, L. A. L. Tang, Y. Wang, C. T.


508 Lim, et al., Origin of enhanced stem cell growth and differentiation on
509 graphene and graphene oxide, ACS Nano 5 (2011) 7334–7341.

36
ACCEPTED MANUSCRIPT

510 [12] M. S. Mauter, M. Elimelech, Environmental applications of carbon-


511 based nanomaterials, Environ. Sci. Technol. 42 (2008) 5843–5859.

PT
512 [13] G. Wang, B. Wang, J. Park, J. Yang, X. Shen, J. Yao, Synthesis of
enhanced hydrophilic and hydrophobic graphene oxide nanosheets by a

RI
513

514 solvothermal method, Carbon 47 (2009) 68 – 72.

SC
515 [14] A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao,
516 et al., Superior thermal conductivity of single-layer graphene, Nano Lett.

U
517 8 (2008) 902–907.

518

519
AN
[15] S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Klein-
hammes, Y. Jia, et al., Synthesis of graphene-based nanosheets via
chemical reduction of exfoliated graphite oxide, Carbon 45 (2007) 1558
M
520

521 – 1565.
D

522 [16] Z.-S. Wu, G. Zhou, L.-C. Yin, W. Ren, F. Li, H.-M. Cheng,
Graphene/metal oxide composite electrode materials for energy storage,
TE

523

524 Nano Energy 1 (2012) 107 – 131.


EP

525 [17] R. Raccichini, A. Varzi, S. Passerini, B. Scrosati, The role of graphene


526 for electrochemical energy storage, Nat. Mater. 14 (2015) 271–279.
C

527 [18] F. Bonaccorso, Z. Sun, T. Hasan, A. C. Ferrari, Graphene photonics


AC

528 and optoelectronics, Nat. Photon. 4 (2010) 611–622.

529 [19] Y. Su, V. G. Kravets, S. L. Wong, J. Waters, A. K. Geim, R. R. Nair,


530 Impermeable barrier films and protective coatings based on reduced
531 graphene oxide, Nat. Comm. 5 (2014) 4843.

37
ACCEPTED MANUSCRIPT

532 [20] B. Wang, B. V. Cunning, S.-Y. Park, M. Huang, J.-Y. Kim, R. S. Ruoff,
533 Graphene coatings as barrier layers to prevent the water-induced corro-

PT
534 sion of silicate glass, ACS Nano 10 (2016) 9794–9800.

[21] B. Tan, N. Thomas, A review of the water barrier properties of poly-

RI
535

536 mer/clay and polymer/graphene nanocomposites, J. Memb. Sci. 514


(2016) 595 – 612.

SC
537

538 [22] X. Zhou, X. Huang, X. Qi, S. Wu, C. Xue, F. Y. C. Boey, et al., In


situ synthesis of metal nanoparticles on single-layer graphene oxide and

U
539

540 reduced graphene oxide surfaces, J. Phys. Chem. C 113 (2009) 10842–
541 10846.
AN
[23] D. Cai, M. Song, A simple route to enhance the interface between
M
542

543 graphite oxide nanoplatelets and a semi-crystalline polymer for stress


transfer, Nanotechnol. 20 (2009) 315708.
D

544

[24] I. Vlassiouk, G. Polizos, R. Cooper, I. Ivanov, J. K. Keum,


TE

545

546 F. Paulauskas, et al., Strong and electrically conductive graphene-based


547 composite fibers and laminates, ACS Appl. Mater. Interfaces 7 (2015)
EP

548 10702–10709.

549 [25] L. S. Walker, V. R. Marotto, M. A. Rafiee, N. Koratkar, E. L. Corral,


C

550 Toughening in graphene ceramic composites, ACS Nano 5 (2011) 3182–


AC

551 3190.

552 [26] K. Gong, Z. Pan, A. H. Korayem, L. Qiu, D. Li, F. Collins, et al., Re-
553 inforcing effects of graphene oxide on portland cement paste, J. Mater.
554 Civ. Eng. 27 (2014).

38
ACCEPTED MANUSCRIPT

555 [27] J. Liu, Y. Yang, H. Hassanin, N. Jumbu, S. Deng, Q. Zuo, et al.,


556 Graphenealumina nanocomposites with improved mechanical properties

PT
557 for biomedical applications, ACS Appl. Mater. Interfaces 8 (2016) 2607–
558 2616.

RI
559 [28] S. R. Bakshi, D. Lahiri, A. Agarwal, Carbon nanotube reinforced metal
matrix composites - a review, Int. Mater. Rev. 55 (2010) 41–64.

SC
560

561 [29] Z. Hu, G. Tong, Q. Nian, R. Xu, M. Saei, F. Chen, C. Chen, M. Zhang,
H. Guo, J. Xu, Laser sintered single layer graphene oxide reinforced

U
562

563 titanium matrix nanocomposites, Compos. Part B Eng. 93 (2016) 352


564 – 359.
AN
[30] W. Zhou, Y. Fan, X. Feng, K. Kikuchi, N. Nomura, A. Kawasaki, Cre-
M
565

566 ation of individual few-layer graphene incorporated in an aluminum ma-


trix, Compos. Part A Appl. Sci. Manuf. 112 (2018) 168 – 177.
D

567

[31] H. Luo, J. Dong, X. Xu, J. Wang, Z. Yang, Y. Wan, Exploring excellent


TE

568

569 dispersion of graphene nanosheets in three-dimensional bacterial cellu-


570 lose for ultra-strong nanocomposite hydrogels, Compos. Part A Appl.
EP

571 Sci. Manuf. 109 (2018) 290 – 297.

572 [32] L. Zhang, H. Li, X. Lai, X. Liao, J. Wang, X. Su, H. Liu, W. Wu,
C

573 X. Zeng, Functionalized graphene as an effective antioxidant in natural


AC

574 rubber, Compos. Part A Appl. Sci. Manuf. 107 (2018) 47 – 54.

575 [33] V. B. Mohan, R. Brown, K. Jayaraman, D. Bhattacharyya, Optimisa-


576 tion of hybridisation effect in graphene reinforced polymer nanocompos-
577 ites, Adv. Compos. Mater. 27 (2018) 349–365.

39
ACCEPTED MANUSCRIPT

578 [34] E. Shamsaei, F. B. de Souza, X. Yao, E. Benhelal, A. Akbari, W. Duan,


579 Graphene-based nanosheets for stronger and more durable concrete: A

PT
580 review, Constr. Build. Mater. 183 (2018) 642 – 660.

581 [35] H. Yang, H. Cui, W. Tang, Z. Li, N. Han, F. Xing, A critical review on

RI
582 research progress of graphene/cement based composites, Compos. Part
A Appl. Sci. Manuf. 102 (2017) 273 – 296.

SC
583

584 [36] A. Gholampour, M. Valizadeh Kiamahalleh, D. N. H. Tran,


T. Ozbakkaloglu, D. Losic, From graphene oxide to reduced graphene

U
585

586 oxide: Impact on the physiochemical and mechanical properties of


587 graphenecement composites, AN ACS Appl. Mater. Interfaces 9 (2017)
588 43275–43286.
M
589 [37] W. Hummers, R. Offeman, Preparation of graphitic oxide, J. Am.
590 Chem. Soc. 80 (1958) 1339–1339.
D

591 [38] M. Hirata, T. Gotou, S. Horiuchi, M. Fujiwara, M. Ohba, Thin-film


TE

592 particles of graphite oxide 1:: High-yield synthesis and flexibility of the
593 particles, Carbon 42 (2004) 2929–2937.
EP

594 [39] Y. Zhang, L. Zhang, C. Zhou, Review of chemical vapor deposition of


595 graphene and related applications, Acc. Chem. Res. 46 (2013) 2329–
C

596 2339.
AC

597 [40] B. Jayasena, S. Subbiah, A novel mechanical cleavage method for syn-
598 thesizing few-layer graphenes, Nanoscale Res. Lett. 6 (2011) 95.

599 [41] K. R. Paton, E. Varrla, C. Backes, R. J. Smith, U. Khan, A. O’Neill,


600 et al., Nat. Mater. 13 (2014) 624–630.

40
ACCEPTED MANUSCRIPT

601 [42] Y. Hernandez, V. Nicolosi, M. Lotya, F. M. Blighe, Z. Sun, S. De,


602 et al., High-yield production of graphene by liquid-phase exfoliation of

PT
603 graphite, Nat. Nanotechnol. 3 (2008) 563–568.

[43] V. Alzari, D. Nuvoli, S. Scognamillo, M. Piccinini, E. Gioffredi,

RI
604

605 et al., Graphene-containing thermoresponsive nanocomposite hydrogels


of poly(n-isopropylacrylamide) prepared by frontal polymerization, J.

SC
606

607 Mater. Chem. 21 (2011) 8727–8733.

U
608 [44] D. Nuvoli, L. Valentini, V. Alzari, S. Scognamillo, S. B. Bon, M. Pic-
609 cinini, et al., High concentration few-layer graphene sheets obtained by
610
AN
liquid phase exfoliation of graphite in ionic liquid, J. Mater. Chem. 21
611 (2011) 3428–3431.
M
612 [45] L. Zhao, X. Guo, C. Ge, Q. Li, L. Guo, X. Shu, J. Liu, Mechanical
behavior and toughening mechanism of polycarboxylate superplasticizer
D

613

614 modified graphene oxide reinforced cement composites, Composites Part


TE

615 B: Engineering 113 (2017) 308 – 316.

616 [46] H. Du, S. D. Pang, Dispersion and stability of graphene nanoplatelet


EP

617 in water and its influence on cement composites, Constr. Build. Mater.
618 167 (2018) 403 – 413.
C

619 [47] D. Dimov, I. Amit, O. Gorrie, M. D. Barnes, N. J. Townsend, A. I. S.


AC

620 Neves, F. Withers, S. Russo, M. F. Craciun, Ultrahigh performance


621 nanoengineered grapheneconcrete composites for multifunctional appli-
622 cations, Adv. Funct. Mater. 28 (2018) 1705183.

41
ACCEPTED MANUSCRIPT

623 [48] S. Bai, L. Jiang, N. Xu, M. Jin, S. Jiang, Enhancement of mechanical


624 and electrical properties of graphene/cement composite due to improved

PT
625 dispersion of graphene by addition of silica fume, Constr. Build. Mater.
626 164 (2018) 433 – 441.

RI
627 [49] C. Chen, A. Ogino, X. Wang, M. Nagatsu, Oxygen functionalization
of multiwall carbon nanotubes by ar/h2o plasma treatment, Diamond

SC
628

629 Relat. Mater. 20 (2011) 153 – 156.

U
630 [50] S. Gilje, S. Han, M. Wang, K. L. Wang, R. B. Kaner, A chemical route
631 to graphene for device applications, Nano Lett. 7 (2007) 3394–3398.

632
AN
[51] S. P. Surwade, Z. Li, H. Liu, Thermal oxidation and unwrinkling of
chemical vapor deposition-grown graphene, J. Phys. Chem. C 116 (2012)
M
633

634 20600–20606.
D

635 [52] J. M. Simmons, B. M. Nichols, S. E. Baker, M. S. Marcus, O. M.


Castellini, C.-S. Lee, et al., Effect of ozone oxidation on single-walled
TE

636

637 carbon nanotubes, J. Phys. Chem. B 110 (2006) 7113–7118.


EP

638 [53] C. Chen, B. Liang, A. Ogino, X. Wang, M. Nagatsu, Oxygen function-


639 alization of multiwall carbon nanotubes by microwave-excited surface-
wave plasma treatment, J. Phys. Chem. C 113 (2009) 7659–7665.
C

640
AC

641 [54] B. Zhao, L. Zhang, X. Wang, J. Yang, Surface functionalization


642 of vertically-aligned carbon nanotube forests by radio-frequency ar/o2
643 plasma, Carbon 50 (2012) 2710–2716.

644 [55] M. G. C. Sales, M. S. D. C. D. Vega, M. R. Vasquez, Properties of

42
ACCEPTED MANUSCRIPT

645 spray-deposited liquid-phase exfoliated graphene films, Jpn. J. Appl.


646 Phys. 57 (2018) 01AF06.

PT
647 [56] C. M. D. Cagomoc, M. R. Vasquez, Enhanced chromium adsorption
capacity via plasma modification of natural zeolites, Jpn. J. Appl. Phys.

RI
648

649 56 (2017) 01AF02.

SC
650 [57] M. S. D. C. Dela Vega, RF Plasma Functionalization of Exfoliated Multi-
651 layered Graphene as Cement Reinforcement, Master’s thesis, University

U
652 of the Philippines, Diliman, Quezon City, 2017.

653

654
AN
[58] ASTM C109/C109M-16a, Standard Test Method for Compressive
Strength of Hydraulic Cement Mortars (Using 2-in. or [50-mm] Cube
Specimens), Standard, ASTM International, Conshohocken, PA, USA,
M
655

656 2016.
D

657 [59] D. Owens, R. Wendt, Estimation of free energy of polymers, J. Appl.


Polymer Sci. 13 (1969) 1741–1747.
TE

658

659 [60] D. H. Kaelble, Dispersion-polar surface tension properties of organic


EP

660 solids, J. Adhesion 2 (1970) 66–81.

661 [61] C. Schneider, W. Rasband, K. Eliceiri, Nih image to imagej: 25 years


C

662 of image analysis, Nat. Methods 9 (2012) 671–675.


AC

663 [62] J. N. Coleman, Liquid exfoliation of defect-free graphene, Acc. Chem.


664 Res. 46 (2013) 14–22.

665 [63] A. Bianco, H.-M. Cheng, T. Enoki, Y. Gogotsi, R. H. Hurt, N. Koratkar,

43
ACCEPTED MANUSCRIPT

666 et al., All in the graphene family a recommended nomenclature for two-
667 dimensional carbon materials, Carbon 65 (2013) 1 – 6.

PT
668 [64] A. Romani, Graphene oxide as a cement reinforcing additive, 2015. Un-
published thesis.

RI
669

670 [65] D. G. Goodwin, A. S. Adeleye, L. Sung, K. T. Ho, R. M. Burgess, E. J.

SC
671 Petersen, Detection and quantification of graphene-family nanomaterials
672 in the environment, Environmental Science & Technology 52 (2018)

U
673 4491–4513. PMID: 29505723.

674

675
AN
[66] A. C. Ferrari, D. M. Basko, Raman spectroscopy as a versatile tool
for studying the properties of graphene, Nat. Nanotechnol. 8 (2013)
235–246.
M
676

677 [67] A. Dey, A. Chroneos, N. S. J. Braithwaite, R. P. Gandhiraman, S. Krish-


D

678 namurthy, Plasma engineering of graphene, Appl. Phys. Rev. 3 (2016)


021301.
TE

679

680 [68] A. Das, B. Chakraborty, A. K. Sood, Raman spectroscopy of graphene


EP

681 on different substrates and influence of defects, Bull. Mater. Sci. 31


682 (2008) 579–584.
C

683 [69] A. Cwirzen, K. Habermehl-Cwirzen, A. Nasibulin, E. Kaupinen,


AC

684 P. Mudimela, V. Penttala, Sem/afm studies of cementitious binder mod-


685 ified by mwcnt and nano-sized fe needles, Mater. Charact. 60 (2009) 735
686 – 740. 11th Euroseminar on Microscopy Applied to Building Materials
687 (EMABM).

44
ACCEPTED MANUSCRIPT

688 [70] G. Y. Li, P. M. Wang, X. Zhao, Mechanical behavior and microstructure


689 of cement composites incorporating surface-treated multi-walled carbon

PT
690 nanotubes, Carbon 43 (2005) 1239 – 1245.

[71] S. Kumar, P. Kolay, S. Malla, S. Mishra, Effect of multiwalled carbon

RI
691

692 nanotubes on mechanical strength of cement paste, J. Mater. Civ. Eng.


24 (2012) 84–91.

SC
693

694 [72] Z. Pan, L. He, L. Qiu, A. H. Korayem, G. Li, J. W. Zhu, et al., Mechani-

U
695 cal properties and microstructure of a graphene oxidecement composite,
696 Cem. Concr. Compos. 58 (2015) 140 – 147.

697
AN
[73] A. Hadi, J. Karimi-Sabet, S. M. A. Moosavian, S. Ghorbanian, Opti-
mization of graphene production by exfoliation of graphite in supercrit-
M
698

699 ical ethanol: A response surface methodology approach, J. Supercrit.


Fluids 107 (2016) 92 – 105.
D

700

[74] S. Park, R. Ruoff, Chemical methods for the production of graphenes 5


TE

701

702 (2010) 309.


EP

703 [75] M. S. Konsta-Gdoutos, Z. S. Metaxa, S. P. Shah, Highly dispersed


704 carbon nanotube reinforced cement based materials, Cem. Concr. Res.
40 (2010) 1052 – 1059.
C

705
AC

706 [76] W. Li, Z. Huang, T. Zu, C. Shi, W. H. Duan, S. P. Shah, Influence of


707 nanolimestone on the hydration, mechanical strength, and autogenous
708 shrinkage of ultrahigh-performance concrete, J. Mater. Civ. Eng. 28
709 (2016) 04015068.

45
ACCEPTED MANUSCRIPT

710 [77] Y. Dutil, D. Rousse, G. Quesada, Sustainable buildings: An ever evolv-


711 ing target, Sustainability 3 (2011) 443464.

PT
712 [78] W.-H. Tsai, S.-J. Lin, J.-Y. Liu, W.-R. Lin, K.-C. Lee, Incorporating
life cycle assessments into building project decision-making: An energy

RI
713

714 consumption and co2 emission perspective, Energy 36 (2011) 3022 –


3029.

SC
715

716 [79] J. Nässén, J. Holmberg, A. Wadeskog, M. Nyman, Direct and indirect

U
717 energy use and carbon emissions in the production phase of buildings:
718 An inputoutput analysis, Energy 32 (2007) 1593 – 1602.
AN
M
D
TE
C EP
AC

46

S-ar putea să vă placă și