You are on page 1of 45

Subscriber access provided by University of Rochester | River Campus & Miner Libraries

Article
Electrochemical exfoliation of graphite in aqueous
sodium halide electrolytes towards low oxygen content
graphene for energy and environmental applications
Jose Maria Munuera, Juan Ignacio Paredes, Marina Enterria, Ana Pagán, Silvia
Villar-Rodil, Manuel Fernando R. Pereira, J. I. Martins, José L. Figueiredo,
Jose Luis Cenis, Amelia Martinez-Alonso, and Juan Manuel D. Tascon
ACS Appl. Mater. Interfaces, Just Accepted Manuscript • Publication Date (Web): 23 Jun 2017
Downloaded from http://pubs.acs.org on June 23, 2017

Just Accepted

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

ACS Applied Materials & Interfaces is published by the American Chemical Society.
1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 44 ACS Applied Materials & Interfaces

1
2
3 Electrochemical exfoliation of graphite in aqueous sodium halide electrolytes towards
4
5
low oxygen content graphene for energy and environmental applications
6
7 J.M. Munueraa*, J.I. Paredesa, M. Enterríab, A. Pagánc, S. Villar-Rodila, M.F.R. Pereirab, J.I.
8 Martinsd,e, J.L. Figueiredob, J.L. Cenisc, A. Martínez-Alonsoa, J.M.D. Tascóna
9
10 a
Instituto Nacional del Carbón, INCAR-CSIC, Apartado 73, 33080 Oviedo, Spain
11 b
12 Laboratório de Processos de Separação e Reacção – Laboratório de Catálise e Materiais
13 (LSRE-LCM), Departamento de Engenharia Química, Faculdade de Engenharia,
14 Universidade do Porto, R. Dr. Roberto Frias s/n, 4200-465 Porto, Portugal
15 c
Instituto Murciano de Investigación y Desarrollo Agrario y Alimentario (IMIDA), Calle
16 Mayor 1, 30150 La Alberca, Spain
17 d
18
Departamento de Engenharia Química, Faculdade de Engenharia, Universidade do
19 Porto, R. Dr. Roberto Frias s/n, 4200-465 Porto, Portugal
e
20 Universidade do Minho, LAB2PT- Laboratório de Paisagens, Património e Território,
21 Portugal
22
23
24
25 Abstract
26
27
28 Graphene and graphene-based materials have shown great promise in many
29
30 technological applications, but their large-scale production and processing by simple and
31
32 cost-effective means still constitute significant issues in the path of their widespread
33
34
implementation. Here, we investigate a straightforward method for the preparation of a
35 ready-to-use and low oxygen content graphene material that is based on electrochemical
36
37 (anodic) delamination of graphite in aqueous medium with sodium halides as the
38
39 electrolyte. Contrary to previous conflicting reports on the ability of halide anions to act as
40
41 efficient exfoliating electrolytes in electrochemical graphene exfoliation, we show that
42 proper choice of both graphite electrode (e.g., graphite foil) and sodium halide
43
44 concentration readily leads to the generation of large quantities of single-/few-layer
45
46 graphene nanosheets possessing a degree of oxidation (O/C ratio down to ~0.06) lower than
47
48 that typical of anodically exfoliated graphenes obtained with commonly used electrolytes.
49
50
The halide anions are thought to play a role in mitigating the oxidation of the graphene
51 lattice during exfoliation, which is also discussed and rationalized. The as-exfoliated
52
53 graphene materials exhibited a three-dimensional morphology that was suitable for their
54
55
56
57 Keywords: graphene, electrochemical exfoliation, supercapacitors, environmental applications, sodium
58 halides
59
60 1
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 2 of 44

1
2
3 practical use without the need to resort to any kind of post-production processing. When
4
5 tested as dye adsorbents, they outperformed many previously reported graphene-based
6
7 materials (e.g., they adsorbed ~920 mg g-1 for methyl orange) and were useful sorbents for
8
9 oils and non-polar organic solvents. Supercapacitor cells assembled directly from the as-
10
exfoliated products delivered energy and power density values (up to 15.3 Wh kg-1 and
11
12 3220 W kg-1, respectively) competitive with those of many other graphene-based devices,
13
14 but with the additional advantage of extreme simplicity of preparation.
15
16
17
18
19 1. Introduction
20
21 Since its first isolation in 2004, graphene has spurred enormous interest in the scientific
22
23 community due to its excellent electronic, mechanical, thermal and optical properties, as
24
25 well as to its wide range of potential applications in many fields, including electronics,
26
photonics, energy conversion/storage, catalysis and biomedicine.1 However, one of the
27
28 most pressing issues that hinder the actual implementation and widespread use of graphene
29
30 is that related to its large-scale production. At present, there is no single method that can
31
32 simultaneously provide high amounts of material of good quality in a reasonable amount of
33
34
time at a low cost. The existing production methods can be classified into two types:
35 bottom-up and top-down. The former relies on synthesizing graphene sheets from small-
36
37 molecule precursors, usually following chemical vapor deposition (CVD) methodologies
38
39 and giving rise to large-area sheets.2 Top-down methods, on the other hand, rely on the
40
41 exfoliation of graphitic materials into single- and/or few-layer graphene flakes. Even
42 though there exists a number of exfoliation strategies, the most relevant ones are arguably
43
44 (i) the graphite oxide route, (ii) the direct liquid-phase exfoliation method and (iii) the
45
46 electrochemical exfoliation approach.3,4 The graphite oxide strategy involves harsh
47
48 oxidation of graphite, causing the oxidized basal planes to be easily exfoliated in polar
49
50
solvents (including water) into single-layer graphene oxide by sonication or other means.5
51 This oxidation can be partially reversed through a reduction process, but the resulting
52
53 product (reduced graphene oxide) always retains significant amounts of oxygen and
54
55 structural defects.6 Direct liquid-phase exfoliation generally circumvents oxidation issues
56
57 because delamination is directly promoted via, e.g., ultrasound waves or high-shear forces
58
59
60 2
ACS Paragon Plus Environment
Page 3 of 44 ACS Applied Materials & Interfaces

1
2
3 in a medium that can colloidally stabilize the graphene flakes (organic solvents, aqueous
4
5 surfactant solutions, etc).7 The main drawbacks of this method are the predominance of
6
7 several-layer or even multilayer graphene flakes rather than single- or few-layer (<4)
8
9 objects, as well as its low production yield (typically below 5 wt%).
10
On the other hand, electrochemical (or electrolytic) exfoliation methods are based on
11
12 the intercalation of ions between the layers of a graphite electrode due to the flow of
13
8,9
14 electrical current in an electrolytic cell. These ions induce expansion of the interlayer
15
16 space in graphite and thus facilitate the exfoliation process. There are essentially two
17
18
approaches to electrochemical exfoliation, namely, the cathodic and the anodic approach. In
19 the former case, graphite acts as the cathode in the electrochemical cell and intercalates
20
21 cations from the electrolyte. Studies in this area have been typically carried out in organic
22
23 solvents with lithium or alkylammonium salts,10,11 as well as molten salts12,13 and ionic
24
25 liquids as the electrolyte,14 and the obtained products are usually few-layer to several-layer
26
graphene flakes of a high structural quality. Anodic exfoliation, on the other hand, takes
27
28 place with the graphite material as the anode and is typically accomplished in aqueous
29
30 electrolytes, which is advantageous from an environmental and practical standpoint.15,16,17,18
31
32 Some studies have also reported the possibility of simultaneous anodic and cathodic
33
34
exfoliation.19,20,21 One of the main issues regarding aqueous anodic exfoliation, though, is
35 related to the fact that graphite is subjected to a large positive voltage (typically between a
36
37 few and a few tens of volts).8 This voltage induces the formation of highly reactive oxygen
38
39 species (e.g., hydroxyl radicals) from the oxidation of water molecules, which in turn
40
41 trigger a significant oxidation and structural degradation the carbon lattice.8,15
42 Due to the well-known ability of the sulfate anion to intercalate graphite, sulfuric acid
43
44 is widely employed as an electrolyte in the preparation of graphene flakes by anodic
45
46 exfoliation.16,17,22,23,24 However, the highly acidic nature of this electrolyte is especially
47
48 conducive to obtaining graphene flakes with a considerable extent of oxidation (typical O/C
49
50
atomic ratios in the ~0.10-0.25 range). Other protic acids, such as phosphoric, oxalic or
51 perchloric acids have also been explored, with similar results. 23,25 Use of the corresponding
52
53 inorganic and organic salts of these acids (e.g., sodium sulfate, citrate or perchlorate) as
54
55 electrolytes has been proposed as a way to mitigate oxidation of the exfoliated graphenes,
56
57 but these efforts have met with varied success.17,22,26 More recently, novel strategies to
58
59
60 3
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 4 of 44

1
2
3 prevent graphene oxidation during anodic exfoliation have been explored.27 For example,
4
5 exfoliation has been accomplished with standard electrolytes (e.g., sulfuric acid) in the
6
7 presence of certain additives that inhibit graphene oxidation, such as melamine,28 or the
8
9 antioxidants sodium borohydride, ascorbic acid and (2,2,6,6-tetramethylpiperidin-1-yl)oxyl
10
(TEMPO).29 Likewise, some sulfonated aromatic compounds have been simultaneously
11
12 used as electrolytes and oxidation-preventing agents for the production of low oxygen
13
14 content graphene flakes.30 In these cases, graphenes with O/C atomic ratios down to ~0.02-
15
16 0.04 were attained. However, such methods suffer from some drawbacks, including the use
17
18
of relatively expensive additives/electrolytes and the difficulty in removing adsorbed
19 additives/electrolytes from the exfoliated products to obtain neat graphenes (the additives
20
21 can also induce functionalization of the flakes). Thus, the availability of anodic exfoliation
22
23 methods that make use of very simple and inexpensive electrolytes and at the same time
24
25 afford low oxygen content and additive-free graphenes would be highly desirable, but for
26
the most part such methods are currently lacking.
27
28 We hypothesized that some very simple salts such as sodium halides (NaCl, NaBr,
29
30 NaI) could be effective, additive-free electrolytes for the anodic exfoliation of graphite into
31
32 graphene flakes with a limited extent of oxidation. This idea was supported by the fact that
33
34
graphitic materials can be intercalated with halogens/halides through electrochemical and
35 other means,31 so that anodic exfoliation of a graphite electrode could in principle be
36
37 possible in the presence of halide-based electrolytes. Although alkali halides have been
38
39 previously used as additives in the preparation of graphene via direct ultrasound-induced
40
41 exfoliation,32 there are conflicting reports as to the ability of halide anions to act as efficient
42 electrolytes in the aqueous anodic exfoliation of graphite.16,17,33 We also note that because
43
44 halide anions are relatively prone to oxidation, these intercalated species could act as
45
46 sacrificial agents and neutralize the oxygen radicals generated under anodic potential, thus
47
48 mitigating the oxidation of the graphene flakes in comparison with other commonly used
49
50
electrolytes, but this point has not yet been demonstrated.
51 We have thus investigated the use of sodium halides as electrolytes for the aqueous
52
53 anodic exfoliation of graphite, the results of which are reported here. Significantly, the
54
55 production of single- and few-layer graphene flakes with a limited degree of oxidation, and
56
57 thus of a high structural quality, was made possible by chosing an appropiate graphite
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 44 ACS Applied Materials & Interfaces

1
2
3 material as the anode and carefully testing different electrolyte concentrations in order to
4
5 find the intermediate range where exfoliation is succesful. Considering the potential
6
7 scalability of this method together with the use of sustainable and widely available
8
9 electrolytes, and that no additives are required to limit oxidation, the present work should
10
broaden the prospects of anodic exfoliation as a competitive method for the large-scale
11
12 production of single-/few-layer graphene flakes towards different applications, particularly
13
14 in energy- and environment-related fields. Indeed, we demonstrate the use of the halide-
15
16 derived, as-exfoliated foam-like graphene products as efficient adsorbents for dyes and oils
17
18
as well as electrodes for supercapacitors. Also, the biocompatibility of these graphene
19 nanosheets is attested.
20
21
22
23 2. Results and discussion
24
25 2.1. Key aspects of the electrochemical exfoliation of graphite in aqueous halide-based
26
electrolytes
27
28 We aimed at determining whether some very simple and widely available salts, such as
29
30 the sodium halides NaCl, NaBr and NaI, could be effectively used as electrolytes towards
31
32 the anodic exfoliation of graphite in aqueous medium to give graphene flakes with a limited
33
34
extent of oxidation. Some previous experiments had concluded that the anodic exfoliation
35 of graphite using halide-based electrolytes was not even possible.16,17 On the other hand, a
36
37 recent work by Munaiah et al has suggested that exfoliated products can be obtained,
38
39 although they possess a considerable degree of oxidation (O/C atomic ratios of ~0.10-20).33
40
41 On the basis of a suite of tests with the aforementioned sodium halides, we found that
42 single-/few-layer graphene flakes could indeed be produced in considerable quantities. It is
43
44 important to note, though, that effective exfoliation was found to be critically dependent on
45
46 two main factors, namely, the type of graphite used as the anode and the concentration of
47
48 sodium halide used as electrolytic medium. To understand the influence of these two
49
50
factors, as well as some characteristics of the resulting graphene materials that will be
51 discussed below, the basic mechanism underlying the anodic exfoliation process has to be
52
53 borne in mind. Such a mechanism has been discussed elsewhere17 and will just be briefly
54
55 outlined here: first of all, exfoliation occurring at a graphite anode in the presence of
56
57 common aqueous electrolytes (e.g., H2SO4) under typical bias voltage conditions (5-15 V)
58
59
60 5
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 6 of 44

1
2
3 is believed to start with the oxidation of grain boundaries and edges of the graphite crystals
4
5 due to the presence of oxygen radicals (e.g., hydroxyl species) generated through oxidation
6
7 of water molecules at the anode. This initial oxidation induces an expansion of the
8
9 interlayer space at the edges, which in turn facilitates the intercalation of the graphite layers
10
with anions from the electrolyte together with water molecules. Subsequent redox processes
11
12 involving the intercalated species are thought to give rise to gaseous products that exert
13
14 strong forces on the layers, leading to the expansion and swelling of the graphite anode. As
15
16 a result, an expanded graphitic material composed of very weakly attached layers is finally
17
18
formed, which can then be readily separated into individual single- or few-layer graphene
19 flakes in a proper solvent medium by means of low-power ultrasound or shear forces.17,18
20
21 In anodic exfoliation tests carried out with sodium halides using three different types of
22
23 graphite, namely, graphite foil, highly oriented pyrolytic graphite (HOPG) and graphite
24
25 flakes (+10 V applied to graphite anode for 60 min with platinum foil as a counter
26
electrode; see Experimental section), we observed that well-exfoliated graphitic materials
27
28 colloidally dispersed either in the organic solvent N,N-dimethylformamide (DMF) or in
29
30 water using, e.g., the biomolecule flavin mononucleotide (FMN) as a stabilizer18,34 could
31
32 only be obtained with the former graphite type. By contrast, just slightly expanded
33
34
materials that could not be stably suspended in DMF or water/FMN were attained with both
35 HOPG and graphite flakes. This result suggested that the halide anions failed to extensively
36
37 intercalate and/or expand the HOPG and graphite flake materials. We believe that the
38
39 distinct expansion behaviour of graphite foil is a consequence of its particular
40
41 morphological features and packing configuration. Specifically, the morphology of graphite
42 foil is typically made up of disordered stacks of micrometer-sized graphitic platelets that
43
44 exhibit many folds, wrinkles and voids in their structure, as noticed from field-emission
45
46 scanning electron microscopy (FE-SEM) images (Fig. 1a). On the other hand, the HOPG
47
48 crystals and graphite flakes are comprised of very tightly packed layers with no or very
49
50
little voids in-between such layers (Fig. 1b and c). The peculiar morphology of graphite foil
51 is a direct result of its production process, which is based on roll compaction of thermally
52
53 expanded graphite particles and leads to a relatively low density product (~0.7-1.3 g cm-3
54
55 according to the specifications of the manufacturer, to be compared with a value of ~2.25 g
56
57 cm-3 for HOPG). It is then reasonable to assume that the presence of a large number of
58
59
60 6
ACS Paragon Plus Environment
Page 7 of 44 ACS Applied Materials & Interfaces

1
2
3 openings and gaps within the layers facilitates the intercalation of chemical species (e.g.,
4
5 halide anions and water molecules) in graphite foil and its subsequent expansion during the
6
7 anodic treatment, whereas such a relative advantage would be essentially absent from
8
9 graphites with a more compact morphology (HOPG and graphite flakes). We also note that
10
explorative exfoliation experiments at different bias voltages, both higher and lower than
11
12 10 V, were carried out. We observed that the exfoliation process was only successful when
13
14 using bias voltages above a threshold value of ~4 V, but at such low voltages the flowing
15
16 current and thus the yield of exfoliated product were lower, whereas the quality of the
17
18
resulting graphene flakes did not change compared to exfoliation at 10 V. For bias voltages
19 above 10 V, neither the amount of exfoliated product nor its quality changed significantly.
20
21 Concerning the current profile during exfoliation, generally speaking it was observed to
22
23 increase at first, probably as a result of the expansion of the graphite anode and the ensuing
24
25 increase in surface area exposed to the aqueous electrolyte. However, after some time the
26
current tended to reach a plateau. The latter was probably due to the fact that further
27
28 increases in surface area resulting from exfoliation were counteracted by the loss of
29
30 expanded graphite fragments that were seen to detach from the anode.
31
32 The distinct fate of graphite foil was apparent from the FE-SEM images of the
33
34
anodically treated materials, as exemplified in Fig. 1d, e, and f using 0.05 M NaCl as the
35 electrolyte. While graphite foil (Fig. 1d) was seen to expand considerably as a result of the
36
37 anodic process, yielding very thin and well separated layers, a rather limited extent of
38
39 cleavage was noticed for both HOPG (Fig. 1e) and graphite flakes (Fig. 1f), which
40
41 exhibited just partially detached graphite platelets of a relatively large thickness. It is thus
42 not surprising that the subsequent mild sonication step was able to complete delamination
43
44 of the anodically treated graphites into graphene flakes only for the case of graphite foil,
45
46 but not for the poorly expanded HOPG and graphite flake materials. In line with this result,
47
48 the effect of compact vs. loose configuration of the graphite anode has been recently
49
50
highlighted by Yang et al, who reported that a modest pre-expansion of graphite foil prior
51 to the anodic exfoliation step with a sulfate-based electrolyte brings about a considerable
52
53 increase in the amount of generated single-/few-layer graphene flakes.29 In the present case,
54
55 we surmise that the intercalation and/or expansion of graphites using halide-based
56
57 electrolytes are relatively inefficient processes compared to those associated to the use of
58
59
60 7
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 8 of 44

1
2
3 commonly employed electrolytes (e.g., sulfate-based ones). Thus, the deployment of
4
5 graphite types for which the intercalation/expansion process is facilitated due to a suitable
6
7 morphology (e.g., having a large amount of interlayer openings, voids and gaps) could
8
9 make a critical difference in attaining a well-expanded product that can then be separated
10
into single-/few-layer graphene flakes, as was actually observed here with graphite foil
11
12 compared to HOPG and graphite flakes. Consequently, the former graphite type was used
13
14 exclusively for the remainder of this work. This forced choice of graphite turns out to be
15
16 quite advantageous with a view to up-scaling the graphene production process, taking into
17
18
account that graphite foil is a widely available and modestly priced commodity (less than
19 US$0.1 per gram, compared to ~US$100 per gram for HOPG). The fact that virtually no
20
21 graphene can be obtained with HOPG and graphite flakes does not necessarily mean that
22
23 the halide anions do not intercalate such graphites at all, but rather that if intercalation takes
24
25 place, it is too slow and/or limited in extent so as not to allow an efficient exfoliation of
26
these graphite types into graphene. Indeed, the latter are known to be readily intercalated
27
28 and delaminated by other commonly used anions, such as sulfates.
29
30 The second decisive factor for successful anodic exfoliation using sodium halides was
31
32 the electrolyte concentration. In general terms, for a given applied bias voltage, the anodic
33
34
process is only able to afford graphene products within a certain range of electrolyte
35 concentrations.17 We tested the range of concentrations for each electrolyte, and found a
36
37 range where the exfoliation process is effective for each sodium halide. If the electrolyte
38
39 concentration is too low, the graphite anode will be poorly intercalated and no or very little
40
41 expanded product will be obtained. If the concentration is too high, a very vigorous and fast
42 intercalation process will usually take place, so that graphite fragments detach too quickly
43
44 from the anode without having completed their layer-by-layer expansion. As a result, a
45
46 large amount of incompletely expanded material will be attained, which cannot be
47
48 subsequently separated into single-/few-layer graphene flakes via sonication, but gives rise
49
50
to multilayer graphitic platelets instead, which are unstable in colloidal dispersion and
51 appear as thick (>10 nm) objects in AFM images. However, if an intermediate electrolyte
52
53 concentration is used, the process is slow enough for the detaching graphite fragments to
54
55 expand and delaminate to an extent large enough so as to yield graphene nanosheets after
56
57 sonication. As a result of this behavior, the final concentration of graphene dispersed in
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 44 ACS Applied Materials & Interfaces

1
2
3 solvents as a function of electrolyte concentration is seen to go through a maximum at
4
5 intermediate electrolyte concentrations (see Table 1 and discussion below). Such a
6
7 constraint in the useful concentration range was seen to be particularly stringent in the case
8
9 of sodium halides. To determine the useful concentration ranges for NaCl, NaBr and NaI,
10
pieces of graphite foil of fixed dimensions were subjected to a positive voltage (+10 V, 60
11
12 min) in aqueous solutions of the aforementioned salts at different concentrations (see
13
14 Experimental section for details). The resulting expanded products (if any) were thoroughly
15
16 rinsed with water and ethanol to remove remnants of salts and other reaction products, and
17
18
were subsequently dispersed via sonication either in DMF or in water/FMN. This cleaning
19 process was seen to be a vital step for the colloidal stability of the final graphene product,
20
21 as very small amounts of residual salts will drastically reduce the colloidal stability of the
22
23 dispersions. After a centrifugation step intended to sediment the non-fully exfoliated
24
25 fraction of the material, the concentration of the supernatant, which was the final graphene
26
dispersion, was measured using UV-vis absorption spectroscopy. We carried out a control
27
28 experiment under the same conditions in the absence of any electrolyte (i.e., only in pure
29
30 water), in order to test the possible influence of the hydroxide anions present in water. In
31
32 this case, no change was seen in the graphite anode and no expanded material was obtained.
33
34
Table 1 shows the concentration of graphene suspensions in DMF obtained with
35 different concentrations of the sodium halides in the anodic expansion process. The use of
36
37 water/FMN as a dispersing medium instead of DMF yielded similar results. A graphene
38
39 concentration value of zero indicates that no significant amount of material could be
40
41 achieved in the final suspension, either because no anodically expanded product could be
42 generated (at low electrolyte concentrations) or because virtually all the expanded product
43
44 sedimented after the sonication and centrifugation steps (at high electrolyte concentrations).
45
46 It can be noticed that NaCl afforded the narrowest useful range of concentrations. Indeed,
47
48 dispersions of graphitic material were only attained at the tested NaCl concentration of 0.05
49
50
M, whereas concentrations of 0.1 M and above as well as 0.01 M and below failed to yield
51 any suspended material. In the case of NaBr, the useful range was seen to be slightly
52
53 broader, with significant amounts of dispersed graphitic material being obtained at
54
55 electrolyte concentrations of both 0.05 and 0.1 M. Finally, NaI was determined to perform
56
57 effectively as an electrolyte in a relatively broad concentration range (between 0.01 and 0.2
58
59
60 9
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 10 of 44

1
2
3 M), leading in all cases to very stable, homogeneous dispersions (see inset to Fig. 2 for a
4
5 digital photograph of representative dispersions in water/FMN obtained with the three
6
7 sodium halides). As a tentative explanation for this behaviour, we note that the reduction
8
9 potential of the iodide anion is lower than that of its bromide and chloride counterparts (see
10
below), so that the better exfoliation behavior with iodide can be ascribed to a more
11
12 extensive formation of the resulting intercalated molecular species (i.e., I2 in the case of
13
14 iodide) that can therefore promote layer cleavage to a larger extent than that attained with
15
16 bromide and chloride. We also note that aqueous chloride anions are not known from the
17
18
literature to intercalate graphite. This could be an indication that such anions are indeed
19 more difficult to intercalate compared to the case of bromide and iodide anions. We believe
20
21 that the fact that chloride intercalation of graphite is not generally thought to be possible is
22
23 due to the very narrow range of concentrations that make the intercalation (and exfoliation)
24
25 possible, as well as the critical role of the type of graphite material used. As we have
26
demonstrated here, the use of a graphite anode with many interlayer openings and voids in
27
28 its structure (i.e., graphite foil) is critical to make intercalation and exfoliation possible.
29
30 This result suggests that intercalation of chloride anions into graphite is possible provided
31
32 that it is facilitated through the use of graphites with somewhat increased interlayer
33
34
spacings. Consequently, we do not believe that the mechanism of chloride intercalation is
35 fundamentally different to that bromide, iodide or other electrolytes; it only needs larger
36
37 graphite interlayer spacings to proceed.
38
39 The graphitic nature of the suspended material was made apparent by UV-vis
40
41 absorption spectroscopy, as demonstrated in Fig. 2 for dispersions in water/FMN solutions.
42 The spectra were dominated by an absorption peak located at ~265-270 nm together with
43
44 strong absorbance in the whole wavelength range above 270 nm. These features are
45
46 characteristic of sp2-based carbon systems having extended electronically conjugated
47
48 domains and have been observed, e.g., both on pristine graphene and on well-reduced
49
50
graphene oxide nanosheets.35 The very weak, shoulder-like absorption bands noticed at
51 about 250, 375 and 450 nm correspond to absorbance from the FMN stabilizer.34 As
52
53 evaluated by this technique, significant amounts of suspended graphitic material (up to
54
55 ~0.75 mg mL-1) were attained for the as-prepared suspensions following the protocol
56
57 described in the Experimental section. However, they could be further increased to the
58
59
60 10
ACS Paragon Plus Environment
Page 11 of 44 ACS Applied Materials & Interfaces

1
2
3 range of several milligrams per milliliter through different post-processing concentration
4
5 strategies, including partial solvent removal using a rotary evaporator or sedimentation of
6
7 the suspension via high-speed centrifugation and re-dispersion in a smaller solvent
8
9 volume.34 The energy consumption in the anodic exfoliation process is largely due to the
10
current flowing through the cell, and is estimated to be ~0.05 kW h for a usual experiment
11
12 carried out in our laboratory considering the power consumption of the power supply.
13
14
15
16 2.2. Characteristics of the exfoliated products and role of the halide anions in the
17
18
exfoliation process
19 Having demonstrated that sodium halides can indeed be used in aqueous anodic
20
21 exfoliation processes to give stable dispersions of graphitic materials, the next step of the
22
23 investigation focused on determining their physicochemical characteristics. First, the lateral
24
25 size and exfoliation degree of the dispersed graphitic materials were investigated by means
26
of scanning transmission electron microscopy (STEM) and atomic force microscopy
27
28 (AFM). Fig. 3 shows typical STEM (a,b,c) and AFM (d,e,f) images recorded from
29
30 water/FMN dispersions obtained with the three different electrolytes at a given
31
32 concentration within their respective useful ranges: 0.05 M NaCl (a,d), 0.1 M NaBr (b,e)
33
34
and 0.1 M NaI (c,f). All the dispersions were comprised of thin nanosheets with mean
35 lateral dimensions of 560±210, 655±390 and 610±260 nm for 0.05 M NaCl, 0.1 M NaBr
36
37 and 0.1 M NaI respectively, and apparent thickness (derived from AFM height profiles)
38
39 between ~2 and 3 nm. The apparent thickness values for nanosheets from dispersions
40
41 obtained at other sodium halide concentrations were very similar and are given in Table 1.
42 The lateral size of the graphene flakes is mainly limited by the sonication step applied to
43
44 the electrolytically expanded graphite material and it is in the same range as that measured
45
46 for other electrolytes we have reported before. We also note that in the present work we
47
48 searched for conditions that allow an efficient delamination of the graphite anode so as to
49
50
attain very thin graphene nanosheets. In addition to the electrolytic treatment, sonication
51 and centrifugation steps were applied to the electrochemically delaminated material. The
52
53 main role of these steps, particularly centrifugation, is to select very thin nanosheets
54
55 generated in the process (if any are generated at all). Therefore, it is not surprising that we
56
57 obtained similar thickness values for all the samples, because the applied processes were
58
59
60 11
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 12 of 44

1
2
3 indeed meant to select the thinnest nanosheets. The fact that different halides lead to
4
5 different amounts of nanosheets is most probably a result of the different efficiency of the
6
7 electrolytes in the delamination of the graphite anode. Considering that AFM thickness
8
9 measurements of graphene nanosheets on SiO2/Si and other substrates are affected by a
10
positive offset of ~0.5-1 nm36 and that FMN molecules adsorbed on both sides of the
11
12 nanosheets could also contribute to the measured thickness, we conclude the actual
13
14 thickness of these nanosheets to be not larger than ~0.5-1.5 nm, implying that they are
15
16 typically comprised between one and four monolayers. Such a result indicates that a high
17
18
degree of exfoliation could be attained with these electrolytes. For electrolyte
19 concentrations above the useful range, virtually no suspended graphitic material was
20
21 attained after the sonication and centrifugation steps, as noted above. However, AFM
22
23 inspection of a sonicated suspension that was allowed to stand undisturbed overnight (no
24
25 centrifugation) revealed that it was comprised of relatively thick objects (typical apparent
26
thickness above ~10 nm). This observation corroborated the idea that high electrolyte
27
28 concentrations lead to an inefficient exfoliation of the graphite anode.
29
30 Chemical analysis of the dispersed materials was carried out by X-ray photoelectron
31
32 spectroscopy (XPS). Fig. 4 shows survey spectra (a,c,e) as well as high resolution C 1s core
33
34
level spectra (b,d,f) for graphene dispersions obtained with 0.05 M NaCl (a,b), 0.1 M NaBr
35 (c,d) and 0.1 M NaI (e,f). In all cases carbon was the dominant element, although oxygen
36
37 was also present to a significant but variable extent depending on the specific electrolyte
38
39 concentration used for exfoliation. Signals corresponding to the halogen used in the anodic
40
41 process were not observed, i.e., the amount of halogen present in the graphene samples was
42 below the detection limit of this technique (~0.1 at%), implying that the washing procedure
43
44 applied to the anodically expanded products was rather effective. This is an advantage over
45
46 some previously reported electrolytes that cannot be readily removed due to their strong
47
48 adsorption onto the graphene nanosheets.30 Another advantage of the present electrolytes is
49
50
related to their cost: sodium halides are simple, inexpensive and widely available
51 chemicals. The O/C atomic ratios for the different tested electrolyte concentrations are
52
53 given in Table 1, which were seen to range between 0.06 and 0.11. The high resolution C
54
55 1s spectra provided information on the nature of the oxidized carbon species. These spectra
56
57 comprised a majority component located at ~284.6 eV that could be ascribed to carbon
58
59
60 12
ACS Paragon Plus Environment
Page 13 of 44 ACS Applied Materials & Interfaces

1
2
3 atoms in unoxidized graphitic environments (C=C species) and a minor component at about
4
5 286.7 eV, which was attributed to carbon atoms single-bonded to oxygen (C-O species), for
6
7 example in hydroxyl, ether or epoxy groups.24 The XPS results thus indicated that the
8
9 exfoliated graphene nanosheets were oxidized only to a limited extent. Specifically, the
10
O/C atomic ratio of samples prepared with 0.05 M NaCl (0.06) and 0.1 M NaBr (0.07) lay
11
12 at the lower end of the range of values that have been reported for most graphenes
13
14 anodically exfoliated using a number of inorganic or organic electrolytes, such as sulfuric
15
16 acid (O/C ratios ~0.08-0.3),17,18,22,30,37 sulfate salts (~0.06-0.19),17,18,22,30,38 phosphoric acid
17
18
(~0.3),23 lithium perchlorate (~0.25),22 sodium methanesulfonate (~0.15)39 or sodium citrate
19 (~0.13).25
20
21 To understand the origin of the limited oxidation degree of the graphene nanosheets
22
23 exfoliated here with the sodium halides, the possible redox processes involving these
24
25 electrolytes and the evolution of oxygen species as a result of water oxidation at the
26
graphite anode must be considered. First of all, as outlined above, it is believed that anionic
27
28 intercalation in general strongly relies on the oxidation of edges and defects of the graphite
29
30 anode by water molecules from the aqueous electrolyte. This oxidation induces a
31
32 significant increase in the interlayer spacing of graphite in the local area around the edges.
33
34
In turn, such a local increase in the interlayer spacing allows even relatively large anions to
35 penetrate the graphitic material at the edges, which thus act as a wedge facilitating the
36
37 subsequent separation of the graphite layers at regions increasingly farther from the edges.
38
39 On the basis of this general mechanism, it follows that having a graphite anode with many
40
41 pre-existing interlayer voids and openings, as it is the case with graphite foil, should clearly
42 facilitate the intercalation and exfoliation of the anode, especially in cases where anion
43
44 intercalation appears to be generally difficult (e.g., the present halides). Second, we note
45
46 that halide intercalation in the interlayer spaces at the positively biased graphite electrode
47
48 should take place most likely in the form of hydrated anions, i.e., halide anions surrounded
49
50
by a bound shell of water molecules40,41,42 (see step A in the schematic process chart of Fig.
51 5). At the high anodic potential used (10 V), oxygen evolution reactions take place and
52
53 these water molecules are expected to readily oxidize (step B1 in Fig. 5), giving first
54
55 hydroxyl radicals (·OH) and then oxygen molecules (O2) through further oxidation43 (step
56
57 C1). The build-up of gaseous O2 in the interlayer spaces should exert a considerable
58
59
60 13
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 14 of 44

1
2
3 pressure on the graphitic layers and hence contribute to their separation from one another.
4
5 Additionally, some of the generated ·OH radicals can attack the exfoliated layers (step C2),
6
7 so that the resulting graphene nanosheets will bear a certain degree of oxidation. Third, it is
8
9 reasonable to assume that the intercalated halide anion itself also participates in redox
10
processes. In particular, it can be oxidized at the graphite electrode (step B2) to give the
11
12 corresponding diatomic molecule (e.g., I2 will be obtained from the anodic oxidation of I-).
13
14 As a matter of fact, when using NaBr and NaI as the electrolyte we observed that the
15
16 aqueous solution progressively turned brown or red-brown around the graphite anode, such
17
18
a color being characteristic of Br2 and I2 dissolved in water. For NaCl no color change was
19 observed: Cl2 is readily hydrolyzed to give colorless HCl and HClO (or their corresponding
20
21 sodium salt forms). In any case, accumulation of the halide oxidation products in the
22
23 interlayer spaces should also contribute to separate the graphene sheets and therefore to the
24
25 overall exfoliation process, in a similar way as it has reported to happen when using sulfate
26
salts for the same process.17
27
28 However, besides reacting directly with the graphite anode, the halide anions can also
29
30 be oxidized by the generated ·OH radicals (step C3 in Fig. 5). This reaction is
31
32 thermodynamically feasible for the three halides contemplated here. Indeed, the standard
33
34
reduction potential of the different halogen/halide pairs in aqueous solution are +1.37 V for
35 Cl2/Cl-, +1.09 V for Br2/Br- and +0.54 V for I2/I-,,44 compared with a value of +1.9 V for the
36
37 ·OH/OH- pair.43 Furthermore, the oxidation of aqueous Cl-, Br- and I- ions by ·OH radicals
38
39 has been well investigated in the past,45,46,47 and such a process is known to involve
40
41 conversion of the highly oxidizing ·OH radicals to hydroxide anions (OH-) or water
42 molecules. The reaction of ·OH radicals with halide anions would thus be expected to
43
44 prevent, at least to some extent, the attack of the graphite electrode by the former species,
45
46 thus providing a mechanism to mitigate the oxidation of the graphene nanosheets that result
47
48 from the anodic treatment. On the other hand, this type of reaction is not expected to occur
49
50
when the electrolyte is comprised of a non-oxidizable anion, such as the sulfate, perchlorate
51 or phosphate anion (for these anions, S, Cl and P are in their corresponding highest possible
52
53 oxidation state). Hence, under such circumstances the intercalating anions would not be
54
55 able to curtail the oxidation of the graphene lattice. Based on these considerations, we
56
57 conclude that the halide anions play a dual role in the anodic exfoliation of graphite to give
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 44 ACS Applied Materials & Interfaces

1
2
3 graphene nanosheets: (1) they act as intercalating electrolyte to promote the exfoliation of
4
5 graphite, and (2) they mitigate the oxidation of the exfoliated graphene nanosheets by
6
7 acting as scavengers of the highly reactive oxygen radicals generated during the anodic
8
9 process. We also note that the presence of voids and openings in the starting graphite foil
10
electrode could contribute to mitigate the oxidation of the resulting graphene nanosheets, as
11
12 it is expected to make the exfoliation process faster and less reliant on the oxidative
13
14 opening of the interlayer spaces compared with the use of more compact types of graphite
15
16 (e.g., graphite rods).18,33
17
18
In addition to the morphological and chemical analysis, the structural quality of the
19 anodically exfoliated graphenes was probed by Raman spectroscopy, as illustrated in Fig. 6
20
21 for samples prepared with 0.05 M NaCl (a), 0.1 M NaBr (b) and 0.1 M NaI (c). In all
22
23 cases, the well-known G and D bands associated to graphite/graphene-based materials and
24
25 located at about 1582 and 1350 cm−1, respectively, were seen to dominate the first-order
26
region of the spectra (i.e., the 1100–1700 cm−1 region), whereas the main feature observed
27
28 in the second-order region (2300–3300 cm−1) was the so-called G´ or 2D band (∼2700
29
30 cm−1). As noticed from Table 1, the integrated intensity ratio of the D and G bands (ID/IG
31
32 ratio), which is usually taken as a quantitative measure of the structural disorder present in
33
34
graphite/graphene materials,48 was seen to differ somewhat between the different graphene
35 samples. Indeed, the ID/IG ratio tended to positively correlate with the O/C atomic ratio.
36
37 Such a result was reasonable considering that anodic oxidation of the graphene nanosheets
38
39 should be associated to the covalent grafting of oxygen functional groups on the carbon
40
41 lattice (e.g., hydroxyl groups; see XPS results from Fig. 4), which in turn should convert a
42 fraction of the originally sp2-hybridized carbon atoms to defect-like sp3-hybridized species.
43
44 We note that, in the calculation of the IG value (integrated intensity of the G band), we
45
46 explicitly exclude the contribution of the D´ band, which appears as a shoulder on the high
47
48 wavenumber side of the G band (~1620 cm-1). Because we are calculating the ID/IG ratio
49
50
and not the ID/(IG+ID´) ratio, we first have to determine the areas associated to the G and D´
51 bands through a peak-fitting procedure. This implies that the actual area of the G band is
52
53 obviously smaller than the area of the composite (G+D´) envelope. As a result, the ID/IG
54
55 ratios given in Table 1 might appear to be somewhat larger than the values an observer
56
57 would guess through direct visual inspection of the spectra shown in Fig. 6. We also
58
59
60 15
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 16 of 44

1
2
3 measured the electrical conductivity of free-standing, paper-like graphene films obtained
4
5 through vacuum filtration of the corresponding dispersions in water/FMN. For example, the
6
7 conductivities of films made up of nanosheets obtained via exfoliation in 0.1 M NaBr (~10
8
9 µm thick, as determined by FE-SEM) and 0.1 M NaI (~12 µm thick) were 15220 and 6750
10
S m-1, respectively. These values were lower by a factor of ~3-6 than those previously
11
12 reported for films derived from completely defect-free, pristine graphene flakes produced
13
14 by direct ultrasound-induced exfoliation of graphite in water/FMN solutions,34 which can
15
16 be attributed to the somewhat defected nature of the present anodically exfoliated materials.
17
18
On the other hand, they were similar to the conductivity values measured beforehand for
19 films made up of pristine graphene flakes prepared by ultrasound-exfoliation in water with
20
21 such stabilizers as sodium dodecylbenzenesulfonate (~1500 S m-1),49 sodium cholate
22
23 (~15000 S m-1),50 sodium taurodeoxycholate (~13000 S m-1)51 or gum Arabic (~10000 S m-
24 1 52
25 ) . We note that these prior values could only be obtained after removing at least part of
26
the stabilizer from the as-filtered films through an appropriate treatment (e.g., high
27
28 temperature annealing under inert or reducing atmosphere), whereas no post-treatment of
29
30 the films was required in the present case to achieve similar conductivity results.
31
32
33
34
2.3. Use of halide-based anodically exfoliated graphene in energy- and environment-
35 related applications
36
37 Owing to their high surface area, good electrical conductivity as well as prominent
38
39 chemical, mechanical and thermal stability, graphene materials are widely investigated
40
41 towards their implementation in a number of environment- and energy-related applications,
42 including their use as sorbents for oils and water-soluble organic compounds (e.g., dyes),53
43
44 or as electrodes for supercapacitors and Li-ion batteries.54 Central to a successful
45
46 performance in many of these applications is the ability to assemble the graphene sheets
47
48 into three-dimensional, macroscopic and porous structures in which the accessible area of
49
50
each individual sheet is retained to the largest possible extent. This has been typically
51 accomplished via appropriate processing of graphene oxide nanosheets or through CVD
52
53 protocols that make use of metal foams as templates,55 but such strategies are generally
54
55 time-consuming, elaborate and/or costly. As an alternative that could be competitive in
56
57 terms of both performance and cost, we propose to use the as-exfoliated graphene materials
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 44 ACS Applied Materials & Interfaces

1
2
3 developed here for some of the mentioned applications. Before their dispersion in any
4
5 solvent via sonication, these materials exhibit a foam-like appearance (Fig. 7a), with an
6
7 estimated density of ~40-45 mg cm-3, and are comprised of expanded and loosely
8
9 connected graphene layers having micrometer- to nanometer-sized gaps in-between (Fig.
10
1d and inset to Fig. 7a). The extremely simple and expeditious nature of their preparation
11
12 process (just anodic treatment of graphite foil in sodium halide electrolyte) could facilitate
13
14 the large-scale deployment of such graphene products in many practical uses. Thus, to
15
16 probe their potential, we examined their performance as adsorbents for dyes and non-polar
17
18
organic solvents/oils, as well as electrodes for supercapacitors. The material exfoliated in
19 0.1 M NaI was selected for these studies mainly because it affords the largest amount of
20
21 graphene products among all the tested electrolytes (see Table 1), which is advantageous
22
23 from the viewpoint of applications.
24
25 Fig. 7b shows the adsorption capacity of the as-exfoliated graphene product towards
26
several dyes in aqueous solution, namely, rhodamine B, basic fuchsin, methylene blue and
27
28 methyl orange, measured at room temperature (295 K) and pH ~6. For the three cationic
29
30 dyes (i.e., rhodamine B, basic fuchsin and methylene blue), values between 300 and 400
31
32 mg g-1 were obtained, being comparable to or even higher than those of many graphene-
33
34
based adsorbent materials that have been previously reported.7,53,56,57,58 In most prior work,
35 however, the adsorbents typically comprised graphene oxide or reduced graphene oxide
36
37 nanosheets, which favored adsorption through electrostatic attraction between the positively
38
39 charged dyes and the negatively charged (ionizable) oxygen functional groups present in
40
41 the nanosheets (e.g., carboxylates), with π-π stacking probably also playing some role.53,57
42
43
On the other hand, although an electrostatic component could certainly be in place for the
44 present anodically exfoliated materials by virtue of their non-negligible oxidation, we
45
46 believe dye adsorption to be more reliant on π-π interactions in this case. This is because
47
48 anodically exfoliated graphenes tend to possess a more graphitic (electronically conjugated)
49
50 character than that of graphene oxides/reduced graphene oxides with a similar or even
51
52
smaller extent of oxidation.59 Furthermore, the adsorption capacity measured for the
53 anionic dye methyl orange was substantially higher than that of its cationic counterparts
54
55 (~920 mg g-1), a result that is not consistent with a predominance of negatively charged
56
57 adsorption sites on the graphene sheets. To the best of our knowledge, the adsorbed amount
58
59
60 17
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 18 of 44

1
2
3 of methyl orange was the highest ever reported for a graphene-based adsorbent. The
4
5 different adsorption performance of the anodically exfoliated material towards methyl
6
7 orange and the three other dyes is currently not understood, but we hypothesize that it is
8
9 related to differences in packing densities of the molecules on the graphene surface. A
10
comparison of dye adsorption capacities of different graphene-based materials from the
11
12 literature is provided in Table S1 of the Electronic Supplementary Information. Likewise,
13
14 due to its primarily hydrophobic nature, the anodically exfoliated material was a good
15
16 sorbent for non-polar organic solvents and oils, such as toluene, hexane, dodecane,
17
18
tetrahydrofuran, acetone, chloroform, ethylenglycol, ethanol, olive oil or pump oil, with the
19 measured sorption capacities ranging between ~12 and 24 g g-1 (Fig. 7c). Although such
20
21 values are lower than those of lighter graphene and other carbon-based sorbents reported in
22
23 the literature (e.g., lightweight graphene aerogels), they are comparable to those of sorbents
24
25 with a similar density.7,56,60 Nonetheless, we expect that the oil sorption performance of the
26
anodically exfoliated materials can be significantly improved on the basis of simple
27
28 strategies, such as working with starting graphite foils of a lower density (densities down to
29
30 ~0.05 g cm-3 have been reported, compared to a value of 0.7-1.3 g cm-3 for the foils used
31
32 here). Because the use of a graphene or any other material in environmental applications
33
34
could require such material to be non-cytotoxic, we also explored the biocompatibility of
35 the present graphenes. In particular, we investigated the proliferation of murine fibroblasts
36
37 onto halide-based graphene film substrates. The results (see SI) suggested these materials to
38
39 be biocompatible.
40
41 Finally, the as-exfoliated graphene material was explored as a supercapacitor electrode
42 in aqueous 1 M H2SO4 electrolyte (two-electrode configuration). The recorded cyclic
43
44 voltammetry (CV) curves (Fig. 8a) exhibited an almost rectangular shape with negligible
45
46 deviation at the limit potentials, even for a voltage window as wide as 1.4 V (CV curves
47
48 were also obtained in a three-electrode configuration, and are presented and discussed in
49
50
the Electronic Supplementary Information). This rather unusual, wide operation voltage
51 was further confirmed by charging-discharging the supercapacitor cell at 0.1 A g-1. As
52
53 noticed from Fig. 8b, the galvanostatic charge-discharge profiles kept a triangular form
54
55 when the voltage was gradually increased from 1.1 to 1.4 V. We note that the working
56
57 voltage window of two-electrode symmetric cells using aqueous electrolytes is generally
58
59
60 18
ACS Paragon Plus Environment
Page 19 of 44 ACS Applied Materials & Interfaces

1
2
3 limited by the electrolytic decomposition of water to values ~1 V.61 In particular, a voltage
4
5 window of 1.4 V is considerably larger than that usually observed for porous carbon-based
6
7 aqueous supercapacitors. For conventional porous carbons (e.g., activated carbons),
8
9 although physisorption can furnish the water molecules with some degree of stability
10
against electrolytic decomposition, the extensive presence of amorphous and/or highly
11
12 defective structures in their lattice, which act as reactive chemisorption sites, tends to
13
14 promote water electrolysis. On the other hand, such reactive sites are expected to be much
15
16 less abundant in the as-exfoliated graphene materials investigated here due to their well-
17
18
preserved graphitic character (Fig. 6). This feature, together with the particular morphology
19 of voids and pores observed for this material (inset to Fig. 7a), is believed to contribute to a
20
21 broader voltage window.
22
23 The evolution of electrode performance with current density was also investigated on
24
25 the basis of charge-discharge tests at 1.4 V, and the results were used for the construction of
26
the Ragone plot shown in Fig. 8c. Even though the cells yielded modest values of specific
27
28 capacitance (e.g., 50 F g-1 at 0.1 A g-1), the fact that the device could be operated at a
29
30 relatively high voltage led to remarkable energy and power density characteristics, with
31
32 values up to 15.3 Wh kg-1 and 3220 W kg-1, respectively. Such a good capacitive behavior
33
34
was made apparent by comparison to other similar graphene-based materials that have been
35 previously reported (also given in Fig. 8c). For example, the power and energy densities
36
37 delivered by the present graphene material were comparable or even higher than those of
38
39 many three-dimensional graphene structures (such as graphene aerogels),62,63,64,65,66
40
41 graphene films,67 graphene compounded with electroactive materials (e.g., polyaniline),68
42 or graphene/porous carbon hybrids.69,70 It is important to note that in addition to
43
44 competitive performance, a critical advantage of the present graphene materials lies in the
45
46 extreme simplicity and environmental friendliness of their preparation compared with that
47
48 of other graphene materials that rely on harsh chemical treatments (e.g, those based on
49
50
graphene oxide), high temperatures and/or complex processing strategies (e.g., supercritical
51 or freeze drying). Additional evidence of the cell stability was demonstrated on the basis of
52
53 cycling tests (Fig. 8d), which revealed that ~98% of the initial capacitance was retained
54
55 after 5000 cycles at 1 A g-1. The plots in Fig. 8d suggest a significant internal resistance
56
57 (IR) drop in the cell, which is discussed in the Supporting Information. In any case, we
58
59
60 19
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 20 of 44

1
2
3 believe that the approach described here could greatly facilitate the practical
4
5 implementation of graphene materials towards supercapacitor applications.
6
7
8
9
10
3. Conclusions
11
12 We have demonstrated that sodium halides (NaCl, NaBr and NaI) are efficient
13
14 electrolytes in the aqueous anodic exfoliation of graphite, affording single-/few-layer
15
16 graphene nanosheets of a high structural quality and limited oxygen content. The successful
17
18
role of these salts as exfoliating electrolytes was seen to rely on two key aspects. First, a
19 proper selection of the graphite source material was required, in particular graphites having
20
21 many openings, gaps and inner voids in their structure. In this regard, graphite foil (a
22
23 modestly priced commodity) turned out to be an appropriate choice. Second, identification
24
25 of an operative range of salt concentrations was also critical towards successful exfoliation.
26
Such a range was seen to be particularly tight in the case of NaCl and NaBr. In addition to
27
28 their role as exfoliating electrolyte, the halide anions were concluded to act as sacrificial
29
30 agents that prevented the extensive oxidation of the graphene lattice during anodic
31
32 exfoliation. Such an oxidation-preventing ability was ascribed to the relatively low redox
33
34
potentials of these anions. Significantly, the as-exfoliated graphene materials exhibited a
35 three-dimensional morphology that allowed their direct use in a number of energy- and
36
37 environment-related applications, thus avoiding the complex processing steps that are
38
39 usually required for many graphene-based materials. Testing of the halide-derived graphene
40
41 products as dye adsorbents, as sorbents for oils and non-polar organic solvents, and as
42 electrodes for supercapacitors revealed a performance comparable to or even higher than
43
44 that frequently reported for different types of graphene (e.g., those obtained from graphite
45
46 oxide). Cell proliferation tests also indicated a good biocompatibility of the halide-derived
47
48 graphenes. Overall, we believe that the extremely simple and environmentally friendly
49
50
nature of the production/processing method of graphene described here, combined with a
51 competitive performance of the material in some relevant applications, constitutes a
52
53 significant step forward towards the practical implementation of graphene in many real-life
54
55 applications. Further efforts to improve the performance of the graphene materials in these
56
57 and other applications along the lines proposed in this work are currently under way.
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5 4. Experimental
6
7 4.1. Materials and reagents
8
9 Graphite anodes used throughout the study consisted of high purity graphite foil
10
(Papyex I980, acquired from Mersen). For comparison purposes, highly oriented pyrolytic
11
12 graphite (HOPG; grade ZYH, obtained from Advanced Ceramics) and natural graphite
13
14 flakes (Sigma-Aldrich, ref. 332461) were also employed. NaCl, NaBr and NaI solutions in
15
16 Milli-Q water were used as electrolytes. N,N-dimethylformamide (DMF) was used as a
17
18
dispersing solvent for graphene, and flavin mononucleotide (FMN, riboflavin 5′-
19 monophosphate sodium salt) was used as a stabilizer to colloidally disperse the graphene
20
21 flakes in aqueous medium. All these chemicals were purchased from Sigma-Aldrich and
22
23 used as received.
24
25
26
4.2. Aqueous anodic exfoliation of graphite using sodium halides as the electrolyte
27
28 The anodic exfoliation process was carried out in a two-electrode configuration, using
29
30 a piece of graphite foil (dimensions: 40×25×0.5 mm3) as the anode and a platinum foil
31
32 (dimensions: 25×25×0.025 mm3) as the cathode. HOPG pieces and graphite flake pellets
33
34
were also used as anodes. In all cases both anode and cathode were held by spring clips
35 connected to the current source. Both electrodes were immersed in an aqueous solution (20
36
37 mL) of a sodium halide (NaCl, NaBr or NaI) at a certain concentration (specified below). A
38
39 positive voltage (10 V) was then applied to the graphite anode for 60 min using an Agilent
40
41 6614C DC power supply. The platinum foil cathode was placed parallel to the graphite foil
42 surface at a distance of about 2 cm. During this process, gas bubbles formed in both
43
44 electrodes, with the graphite anode being typically seen to expand and release graphitic
45
46 fragments from its surface. After the 60 min electrolysis period, the resulting expanded
47
48 graphitic product was gently rubbed off from the graphite anode using a spatula, thoroughly
49
50
rinsed with copious amounts of Milli-Q water and ethanol to remove residual salts as well
51 as other products of the electrochemical reaction and then dried overnight at room
52
53 temperature under reduced pressure. Subsequently, this expanded material was sonicated in
54
55 a given volume (20 mL) of either DMF or aqueous FMN solution (1 mg mL-1) for 3 hours
56
57 (Selecta Ultrasons system, 40 kHz), and then the resulting dispersion was centrifuged at
58
59
60 21
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 22 of 44

1
2
3 200 g for 20 min (Eppendorf 5424 microcentrifuge) to sediment insufficiently exfoliated
4
5 material. The obtained supernatant was collected and stored for further use. In the case of
6
7 the FMN-stabilized aqueous dispersions, they were further purified (to remove the fraction
8
9 of free, non-adsorbed FMN from the medium) by two cycles of sedimentation via
10
centrifugation (20000 g, 20 min) and re-suspension in Milli-Q water with a brief sonication
11
12 step.
13
14
15
16 4.3. Characterization techniques
17
18
Characterization of the graphene samples was carried out by means of UV-vis
19 absorption spectroscopy, Raman spectroscopy, X-ray photoelectron spectroscopy (XPS),
20
21 field emission scanning electron microscopy (FE-SEM), scanning transmission electron
22
23 microscopy (STEM), atomic force microscopy (AFM) and measurement of electrical
24
25 conductivity. UV-vis absorption spectroscopy was carried out with a double-bean Heλios α
26
spectrophotometer, from Thermo Spectronic. Following previous reports,18,34 the
27
28 suspension concentration was estimated by means of UV-vis absorption spectroscopy on
29
30 the basis of the Lambert-Beer law and using a value of 2440 mL mg-1 m-1 for the extinction
31
32 coefficient at the measured wavelength of 660 nm. Raman spectra were acquired with a
33
34
Horiba Jobin-Yvon LabRam instrument at an incident power of 2.5 mW and using a laser
35 excitation wavelength of 532 nm. XPS measurements were performed on a SPECS
36
37 apparatus at a pressure of 10-7 Pa and using a non-monochromatic Mg Kα X-ray source
38
39 (11.81 kV, 100 W). Samples for both Raman spectroscopy and XPS were prepared by
40
41 drop-casting the corresponding aqueous, FMN-stabilized graphene dispersion onto a pre-
42
43
heated (~50-60 ºC) circular stainless steel sample-holder 12 mm in diameter to give thin,
44 macroscopic graphene films. XPS survey spectra were used to determine the overall
45
46 elemental composition of the graphene samples in such films. Although XPS is a surface-
47
48 sensitive technique that only probes the outermost surface layer of materials (a few to
49
50 several nanometers deep), in the case of graphene samples prepared from colloidal
51 dispersions it can be inferred that the elemental compositions derived from XPS are
52
53 representative of the whole sample. Because the colloidal graphene suspensions used to
54
55 prepare the thin films are expected to be spatially homogeneous, there is no reason to
56
57 believe that the graphene nanosheets constituting the other surface of the formed films are
58
59
60 22
ACS Paragon Plus Environment
Page 23 of 44 ACS Applied Materials & Interfaces

1
2
3 not representative of the whole graphene sample, i.e., the graphene sheets probed in the
4
5 XPS measurement must be representative of the sample as a whole. Also, because the
6
7 thickness of the graphene nanosheets is generally comparable to the probing depth of XPS,
8
9 this technique must probe individual graphene nanosheets in its entirety, and not just the
10
outer surface of nanosheets. Having these two factors in mind, we have to conclude that
11
12 survey XPS spectra must provide a reasonably accurate picture of the elemental
13
14 composition of this type of graphene samples. FE-SEM and STEM imaging was
15
16 accomplished with a Quanta FEG 650 system (FEI Company) operated at 20-30 kV.
17
18
Specimens for STEM were prepared by mixing a given volume of aqueous graphene
19 suspension with an equal volume of ethanol, and then 40 µL of the resulting water-ethanol
20
21 dispersion were drop-cast onto copper grids (200 mesh) covered with a thin continuous film
22
23 of amorphous carbon. AFM images were recorded on a Nanoscope IIIa Multimode
24
25 apparatus (Veeco Instruments) in the tapping mode of operation under ambient conditions.
26
27
Rectangular silicon cantilevers with nominal spring constant and resonance frequency
28 values of ~40 N m-1 and 250-300 kHz, respectively, were employed. HOPG pieces or
29
30 SiO2(300 nm)/Si wafers were used as substrates in the preparation of specimens for AFM.
31
32 To this end, a small volume (20-40 µL) of the graphene dispersion at a concentration of
33
34 ~0.05 mg mL-1 was drop-cast onto the HOPG or SiO2/Si substrate pre-heated at ~50-60 ºC
35
36
and then allowed to dry. For the measurement of electrical conductivity, free-standing
37 paper-like films were prepared from the aqueous graphene dispersions by vacuum filtration
38
39 through silver membrane filters 25 mm in diameter and 0.2 µm in pore size (Sterlitech
40
41 Corporation). Conductivity was determined through the van der Pauw method by means of
42
43 a homemade setup (Agilent 6614C potentiostat and Fluke 45 digital multimeter) with
44
45
12×12 mm2 square pieces that were cut from the free-standing films. The thickness of the
46 films was estimated by FE-SEM.
47
48
49
50 4.4. Testing of as-prepared anodically exfoliated graphene in energy and environmental
51
52 applications
53 The graphene material obtained by anodic exfoliation in 0.1 M NaI was tested as an
54
55 adsorbent for water-soluble dyes as well as for non-polar organic solvents and oils, and as
56
57 an electrode for supercapacitors. To this end, the as-exfoliated product, i.e. the material
58
59
60 23
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 24 of 44

1
2
3 directly generated by the anodic treatment without subsequent sonication (only washed to
4
5 remove remnants of the electrolyte), was used. For the dye adsorption experiments, ~10 mg
6
7 of the graphene material was added to 10 mL of an aqueous solution of the dye (rhodamine
8
9 B, basic fuchsin, methylene blue and methyl orange) at room temperature and pH ~6,
10
which was then gently stirred with a magnetic bar for 48 h to ensure that the adsorption
11
12 equilibrium was attained. After removal of the graphene component via centrifugation
13
14 (10000 g, 10 min), the concentration of dye remaining in the aqueous solution was
15
16 determined through UV-vis absorption spectroscopy on the basis of absorption peaks
17
18
characteristic of each dye (located at ~555 nm for rhodamine B, 540 nm for basic fuchsin,
19 660 nm for methylene blue and 460 nm for methyl orange), which in turn was used to infer
20
21 the amount of dye adsorbed on the graphene sample. To estimate maximum adsorption
22
23 capacities, different starting concentrations of the dye were probed (typically up to 1-2 mg
24
25 mL-1). For the oil/non-polar organic solvent sorption experiments, a certain amount of the
26
graphene material (~5 mg) was placed in a 1.5 mL centrifuge tube, and then small known
27
28 volumes of the oil/organic solvent (typically ~80-140 µL) were added in succession. These
29
30 volumes were quickly absorbed by the graphene sample until a saturation point was
31
32 reached, which was used to estimate the amount retained by the graphene sorbent. Sorption
33
34 assays were carried out with the following compounds: toluene, hexane, dodecane,
35 tetrahydrofuran, acetone, chloroform, ethylenglycol, ethanol, olive oil, new and used pump
36
37 oil.
38
39 For the supercapacitor electrode tests, the as-exfoliated graphene material was hand-
40
41 ground to obtain a homogeneous powder that was subsequently mixed with
42
43
polytetrafluoroethylene (PTFE) in absolute ethanol to obtain a carbon paste
44 (graphene:PTFE weight ratio of 8:1). This paste (~7 mg) was pressed in a mould under 1
45
46 ton for 15 s to obtain two circular pellets (~0.78 cm2) with similar mass and thickness. The
47
48 pellets were degassed at 100 ºC under vacuum for 4h, soaked into the electrolyte (1 M
49
50 H2SO4) for 24 h, and then used as symmetric working electrodes in a two-electrode
51 Swagelok-type with glass microfiber filter (grade 934-AH, from Whatman) acting as the
52
53 separator. The supercapacitor performance was tested by means of a computer-controlled
54
55 potentiostat (PGZ 402 Voltalab, from Radiometer Analytical). Cyclic voltammetry (CV)
56
57 tests were performed at 10 mV s-1 and different voltage windows. Galvanostatic charge-
58
59
60 24
ACS Paragon Plus Environment
Page 25 of 44 ACS Applied Materials & Interfaces

1
2
3 discharge profiles were recorded up to different voltages (1.1-1.4 V) and at different current
4
5 loads (0.1-6 A g-1). Long-term stability tests were carried out by cycling the cell at a
6
7 constant current density of 1 A g-1 for 5000 cycles. The gravimetric capacitance, CS (F g-1)
8
9 of a single electrode was calculated on the basis of the following formula:
10
11
12 2  ∆

13  = (1)
14 ∆

15
16
17 , where I is the current (A), ∆td is the time of discharge at the fixed potential window (s), m
18
19 is the mass of the active material within the working electrode (g) and ∆Vd is the potential
20
21 difference (V) between start and finish of the discharge process after omitting the internal
22
resistance (IR) drop. The gravimetric energy density of the electrode, E (Wh kg-1), was
23
24 calculated through the formula:
25
26
27
28 1
 =  ∆  (2)
29 2 

30
31
32
, where CCELL is the gravimetric capacitance of the total cell (F g-1). Finally, the gravimetric
33
34 power density, P (W kg-1), was calculated by applying the following equation:
35
36
37
38 
= (3)
39 ∆

40
41
42
43
44
Supporting Information
45 Cell proliferation tests of murine fibroblasts grown onto halide-based graphene substrates.
46
47 Characterization of as-exfoliated graphene material by N2 physisorption. Additional XPS
48
49 data of the exfoliated graphenes. Three electrode characterization of as-prepared anodically
50
51 exfoliated graphene. Discussion of IR drop contribution to the supercapacitor cell
52 performance. Table comparing adsorption capacities of dyes with different graphene-based
53
54 materials.
55
56
57
58 Corresponding Author
59
60 25
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 26 of 44

1
2
3 *Telephone: (+34) 985 11 90 90. Fax: (+34) 985 29 76 62. Email:
4
5 j.munuera@incar.csic.es.
6
7
8
9 Acknowledgements
10
Financial support from both the Spanish Ministerio de Economía y Competitividad
11
12 (MINECO) and the European Regional Development Fund (ERDF) through grant
13
14 MAT2015-69844-R is gratefully acknowledged. We also acknowledge partial funding by
15
16 both Plan de Ciencia, Tecnología e Innovación 2013-2017 del Principado de Asturias and
17
18
the ERDF through Project GRUPIN14-056. This work was also partially funded by project
19 “AIProcMat@N2020 - Advanced Industrial Processes and Materials for a Sustainable
20
21 Northern Region of Portugal 2020” (NORTE-01-0145-FEDER-000006), supported by
22
23 Norte Portugal Regional Operational Programme (NORTE 2020), under the Portugal 2020
24
25 Partnership Agreement, through the European Regional Development Fund (ERDF) and of
26
Project POCI-01-0145-FEDER-006984 – Associate Laboratory LSRE-LCM funded by
27
28 ERDF through COMPETE2020 - Programa Operacional Competitividade e
29
30 Internacionalização (POCI) – and by national funds through FCT - Fundação para a Ciência
31
32 e a Tecnologia. A.P. thanks her research contract partially supported (80%) by the
33
34
ERDF/FEDER Operative Programme of the Region of Murcia (Project No. 14-20-01).
35 J.M.M. is grateful to the Spanish Ministerio de Educación, Cultura y Deporte (MECD) for
36
37 his pre-doctoral contract (FPU14/00792).
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 26
ACS Paragon Plus Environment
Page 27 of 44 ACS Applied Materials & Interfaces

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

1
2
3 Table 1. Characteristics of the graphene materials obtained by aqueous anodic exfoliation
4
5
of graphite foil in different concentrations of the sodium halides NaCl, NaBr and NaI. [X-]:
6 electrolyte concentration in the aqueous anodic system. [Graphene]: concentration of
7 graphene dispersed in DMF after anodic exfoliation, sonication and centrifugation. O/C
8 atomic ratio derived from XPS survey spectra of the samples. Apparent thickness of the
9 nanosheets derived from AFM height data.
10
11
12 Electrolyte [X-] (M) [Graphene] O/C ratio ID/IG Apparent
13 (mg/mL) thickness (nm)
14 NaCl 0.01 0 - - -
15 0.05 0.21 0.06 0.8 2-3
16
0.10 0 - - -
17
18 0.20 0 - - -
19 0.50 0 - - -
20 1.00 0 - - -
21 NaBr 0.01 0 - - -
22 0.05 0.05 0.08 0.9 2-3
23
24 0.10 0.26 0.07 1.1 2-3
25 0.20 0 - - -
26 NaI 0.01 0.11 0.08 1.3 2-3
27 0.05 0.57 0.11 1.4 2-3
28 0.10 0.75 0.11 1.4 2-3
29
30
0.20 0.24 0.08 1.2 2-3
31 0.50 0 - - -
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 28
ACS Paragon Plus Environment
Page 29 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
a b c
12
13
14
15
16
17
18
19
20
21 d e f
22
23
24
25
26
27
28
29
30
31
Figure 1. Representative FE-SEM images of graphite foil (a,d), HOPG (b,e) and graphite
32 flakes (c,f) before (a,b,c) and after (d,e, f) electrochemical treatment in 0.05 M NaCl (+10
33 V, 60 min).
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 29
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 30 of 44

1
2
3
4 1.8
5
6 1.6
7 1.4

Absorbance
8 1.2
9 1
10
11 0.8
12 0.6
13 0.4
14
0.2
15
16 0
17 200 400 600 800 1000
18 Wavelength (nm)
19
20 Figure 2. UV-visible absorption spectra of aqueous FMN-stabilized graphene dispersions
21 prepared through anodic exfoliation in 0.05 M NaCl (black trace), 0.1 M NaBr (green trace)
22 and 0.1 M NaI (red trace). Inset: digital photograph of aqueous FMN-stabilized graphene
23 dispersions prepared with 0.05 M NaCl (right), 0.1 M NaBr (middle) and 0.1 M NaI (left).
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 30
ACS Paragon Plus Environment
Page 31 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7 a b c
8
9
10
11
12
13
14
15
16
17
18 d e f
19
20
21
22
23
24
25
26
27
28
29 Figure 3. Representative STEM (a,b,c) and AFM (d,e,f) images of graphene nanosheets
30 obtained by aqueous anodic exfoliation of graphite foil in 0.05 M NaCl (a,d), 0.1 M NaBr
31
32 (b,e) and 0.1 M NaI (c,f). In d, e and f, line profiles (black traces) taken along the marked
33 white lines are shown overlaid on the AFM images.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 31
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 32 of 44

1
2
3
4
5
6
7
8
9
10
11 a C 1s b
12
13
14
15
O 1s
16
17
18
19 c d
Intensity (a. u.)

20
21
22
23
24
25
26 e f
27
28
29
30
31
32
33 600 500 400 300 200 290 288 286 284 282

34 Binding energy (eV) Binding energy (eV)


35
36
37 Figure 4. XPS survey spectra (a,c,e) and high resolution C 1s core level spectra (b,d,f) for
38 graphene samples obtained by aqueous anodic exfoliation of graphite foil in 0.05 M NaCl
39 (a,b), 0.1 M NaBr (c,d) and 0.1 M NaI (e,f).
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 32
ACS Paragon Plus Environment
Page 33 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Figure 5. Schematic chart depicting the possible redox processes that take place when
44
45 halide anions (X- ≡ Cl-, Br- or I-) are used as electrolyte in the aqueous anodic exfoliation of
46 graphite. Step A: intercalation of hydrated halide [X-(aq)] anions in the interlayer space of
47
graphite. B1: anodic oxidation of intercalated water molecules to give ·OH radicals. B2:
48
49 anodic oxidation of intercalated halide anions to give X2 molecules. C1: further oxidation
50 of ·OH radicals to give O2 molecules. C2: attack of the graphitic lattice by ·OH radicals to
51
52
give oxidized graphene materials. C3: neutralization of the highly reactive ·OH radicals by
53 reaction with halide anions.
54
55
56
57
58
59
60 33
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 34 of 44

1
2
3
4
5
6
7
8 a G
9
10
11 2D
12
D
13
14
15 b
Intensity (a. u.)

16
17
18
19
20
21
22
23 c
24
25
26
27
28
29 1500 2000 2500 3000
30 -1
Raman shift (cm )
31
32 Figure 6. Raman spectra of graphene samples prepared by aqueous anodic exfoliation of
33 graphite foil in 0.05 M NaCl (a), 0.1 M NaBr (b) and 0.1 M NaI (c).
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 34
ACS Paragon Plus Environment
Page 35 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
6
7
8
9
10
11 1000
b 30 c

Adsorption capacity (mg g-1)


12 a
13 800 25

Weight gain (g g-1)


14
600 20
15 300 µm
16 15
17 400
10
18
19 200
5
20
0 0
21
Methylene Blue

Methyl Orange
Basic Fuchsin

Rhodamin B

Hexane

Acetone
Chloroform
Ethylenglycol
Ethanol
Olive Oil
Toluene

Dodecane

Pump Oil (new)


Pump Oil (used)
Tetrahydrofuran
22
23
24
25
26
27
28
29
30 Figure 7. Digital photograph of as-prepared anodically exfoliated graphene using 0.1M NaI
31 as electrolyte (a), SEM image of the same material (inset), adsorption capacities of this
32 material for different dyes (b) and weight gains of the material for different solvents and
33
34 oils (c).
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 35
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 36 of 44

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 150 102
17 a 1.1 V 1.3 V C
Capacitance (F/ g)

100
18
Expanded graphite
Halide-based anodic graphene (this work)
50
19 RGORuO2/ PANI
20 0

21
ref [7]
Ref. 68
-50
22 1.2 V 1.4 V
-100
23 GH-Hz8

Energy Density (Wh kg-1)


24 -150 BN-GAs Ref. 62
ref [2]
0.0 0.4 0.8 1.2
25 Voltage (V)
Ref. 65
ref [5]
26 3D graphene
27
1.4 Ref. 66
ref [6]
b U-GAs 101
28 1.2
1.4 V
Ref. 65
Voltage (V)

29 1.0 ref [5]


30 1.3 V
0.8
31
0.6 1.2 V
32
HPGF 600
33 0.4 1.1 V
34 ref [3]
Ref. 63
0.2
35 e-CMG
36 -600 -400 -200 0 200 400 600
Ref. 64
ref [4]
37 time (sec) RGO/ CMK5
38 1.4
ref [9]
Ref. 69
39 d 100 cycles RGO 100
40 1.2
5000 cycles K1T8
Ref. 68
ref [7]
Voltage (V)

41 1.0 ref [10]


Ref. 70
42 0.8 SSG
43 Ref. 67
ref [8]
0.6
44
45 0.4
46 0.2 Power Density (W Kg-1)
47
48 -20 -10 0 10 20
49 time (sec)
101 102 103 104 105
50
51 Figure 8. Supercapacitor performance tests for the halide-exfoliated graphite foil (0.1M
52
53 NaI); a) cyclic voltammetry curves and b)galvanostatic charge-discharge profiles at
54 different potential windows, c) Ragone plot for the exfoliated graphite and for some
55 representative graphene-derived porous structures and d) cycling stability of the symmetric
56
57 supercapacitor at 1 A/g over 2500 cycles.
58
59
60 36
ACS Paragon Plus Environment
Page 37 of 44 ACS Applied Materials & Interfaces

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

1
2
3 Table of contents graphic
4
5
6
7
1000

Adsorption capacity
8 Halide-based graphene
9 800 3D GRAPHENE
101

Specific Energy (Wh kg )


10 600

-1
11 400
12 U-GAs
200
13
14 0

MB
BF
MO
RhB
HPGF 600
100
15
-1
16 Specific Power (W Kg )

17 10
2
10
3
10
4

18
19
20
21
22
23
References
24
25
26
27 1 Ferrari, A. C.; Bonaccorso, F.; Fal'ko, V.; Novoselov, K. S.; Roche, S.; Boggild, P.;
28 Borini, S.; Koppens, F. H. L.; Palermo, V.; Pugno, N.; Garrido, J. A.; Sordan, R.; Bianco,
29
30 A.; Ballerini, L.; Prato, M.; Lidorikis, E.; Kivioja, J.; Marinelli, C.; Ryhanen, T.; Morpurgo,
31 A.; Coleman, J. N.; Nicolosi, V.; Colombo, L.; Fert, A.; Garcia-Hernandez, M.; Bachtold,
32 A.; Schneider, G. F.; Guinea, F.; Dekker, C.; Barbone, M.; Sun, Z.; Galiotis, C.;
33 Grigorenko, A. N.; Konstantatos, G.; Kis, A.; Katsnelson, M.; Vandersypen, L.; Loiseau,
34 A.; Morandi, V.; Neumaier, D.; Treossi, E.; Pellegrini, V.; Polini, M.; Tredicucci, A.;
35
36
Williams, G. M.; Hee Hong, B.; Ahn, J.-H.; Min Kim, J.; Zirath, H.; van Wees, B. J.; van
37 der Zant, H.; Occhipinti, L.; Di Matteo, A.; Kinloch, I. A.; Seyller, T.; Quesnel, E.; Feng,
38 X.; Teo, K.; Rupesinghe, N.; Hakonen, P.; Neil, S. R. T.; Tannock, Q.; Lofwander, T.;
39 Kinaret, J. Science and Technology Roadmap for Graphene, Related Two-Dimensional
40 Crystals, and Hybrid Systems. Nanoscale 2015, 7, 4598-4810.
41
42
43 2 Zhang, Y.; Zhang, L.; Zhou, C. Review of Chemical Vapor Deposition of Graphene and
44 Related Applications. Acc. Chem. Res. 2013, 46, 2329-2339.
45
46 3 Zhong, Y. L.; Tian, Z.; Simon, G. P.; Li, D. Scalable Production of Graphene Via Wet
47 Chemistry: Progress and Challenges. Mater. Today 2015, 18, 73-78.
48
49
50 4 Du, W.; Jiang, X.; Zhu, L. From Graphite to Graphene: Direct Liquid-Phase Exfoliation
51 of Graphite to Produce Single- and Few-Layered Pristine Graphene. J. Mater. Chem. A
52 2013, 1, 10592-10606.
53
54 5 Chen, L.; Xu, Z.; Li, j.; Li, Y.; Shan, M.; Wang, C.; Wang, Z.; Guo, Q.; Liu, L.; Chen,
55
56 G.; Qian, X. A Facile Strategy to Prepare Functionalized Graphene Via Intercalation,
57 Grafting and Self-Exfoliation of Graphite Oxide. J. Mater. Chem. 2012, 22, 13460-13463
58
59
60 38
ACS Paragon Plus Environment
Page 39 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
6 6 Pei, S.; Cheng, H.-M. The Reduction of Graphene Oxide. Carbon 2012, 50, 3210-3228.
7
8 7 Niu, L.; Coleman, J. N.; Zhang, H.; Shin, H.; Chhowalla, M.; Zheng, Z. Production of
9 Two-Dimensional Nanomaterials Via Liquid-Based Direct Exfoliation. Small 2016, 12,
10 272-293
11
12
13 8 Abdelkader, A. M.; Cooper, A. J.; Dryfe, R. A. W.; Kinloch, I. A. How to Get Between
14 the Sheets: a Review of Recent Works on the Electrochemical Exfoliation of Graphene
15 Materials from Bulk Graphite. Nanoscale 2015, 7, 6944-6956.
16
17
18
9 Yang, S.; Lohe, M. R.; Müllen, K.; Feng, X. New-Generation Graphene from
19 Electrochemical Approaches: Production and Applications. Adv. Mater. 2016, 28, 6213-
20 6221.
21
22 10 Abdelkader, A. M.; Kinloch, I. A.; Dryfe, R. A. W. Continuous Electrochemical
23 Exfoliation of Micrometer-Sized Graphene Using Synergistic Ion Intercalations and
24
25 Organic Solvents. ACS Appl. Mater. Interfaces 2014, 6, 1632-1639.
26
27 11 Cooper, A. J.; Wilson, N. R.; Kinloch, I. A.; Dryfe, R. A. W. Single Stage
28 Electrochemical Exfoliation Method for the Production of Few-Layer Graphene Via
29 Intercalation of Tetraalkylammonium Cations. Carbon 2014, 66, 340-350.
30
31
32 12 Ali Reza Kamali, A. R., Fraya, D. J. Large-Scale Preparation of Graphene by High
33 Temperature Insertion of Hydrogen into Graphite. Nanoscale, 2015, 7, 11310-11320
34
35 13 Kim, H. K.; Kamali, A. R.; Roh, K. C.; Kim, K. B.; Fray, D. J. Dual Coexisting
36
Interconnected Graphene Nanostructures for High Performance Supercapacitor
37
38 Applications. Energy Environ. Sci. 2016, 9, 2249-2256
39
40 14 Abdelkader, A. M.; Patten, H. V.; Li, Z.; Chen, Y.; Kinloch, I. A. Electrochemical
41 Exfoliation of Graphite in Quaternary Ammonium-Based Deep Eutectic Solvents: a Route
42 for the Mass Production of Graphane. Nanoscale, 2015, 7, 11386-11392
43
44
45 15 Lu, J.; Yang, J.-x.; Wang, J.; Lim, A.; Wang, S.; Loh, K. P. One-Pot Synthesis of
46 Fluorescent Carbon Nanoribbons, Nanoparticles, and Graphene by the Exfoliation of
47 Graphite in Ionic Liquids. ACS Nano 2009, 3, 2367-2375.
48
49
16 Su, C.-Y.; Lu, A.-Y.; Xu, Y.; Chen, F.-R.; Khlobystov, A. N.; Li, L.-J. High-Quality
50
51 Thin Graphene Films from Fast Electrochemical Exfoliation. ACS Nano 2011, 5, 2332-
52 2339.
53
54 17 Parvez, K.; Wu, Z.-S.; Li, R.; Liu, X.; Graf, R.; Feng, X.; Müllen, K. Exfoliation of
55 Graphite into Graphene in Aqueous Solutions of Inorganic Salts. J. Am. Chem. Soc. 2014,
56
57
136, 6083-6091.
58
59
60 39
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 40 of 44

1
2
3
4
5
18 Munuera, J. M.; Paredes, J. I.; Villar-Rodil, S.; Ayán-Varela, M.; Pagán, A.; Aznar-
6 Cervantes, S. D.; Cenis, J. L.; Martínez-Alonso, A.; Tascón, J. M. D. High Quality, Low
7 Oxygen Content and Biocompatible Graphene Nanosheets Obtained by Anodic Exfoliation
8 of Different Graphite Types. Carbon 2015, 94, 729-739.
9
10 19 Najafabadi, A. T.; Gyenge, E. Synergistic Production of Graphene Microsheets by
11
12 Simultaneous Anodic and Cathodic Electro-Exfoliation of Graphitic Electrodes in Aprotic
13 Ionic Liquids. Carbon, 2015, 84, 449-459
14
15 20 Najafabadi, A. T.; Ng, N.; Gyenge, E. High-Yield Graphene Production by
16 Electrochemical Exfoliation of Graphite: Novel Ionic liquid (IL)–Acetonitrile Electrolyte
17
18
With Low IL Content. Carbon, 2014, 71, 58-69
19
20 21 Najafabadi, A. T.; Gyenge, E. Electrochemically Exfoliated Graphene Anodes with
21 Enhanced Biocurrent Production in Single-Chamber Air-Breathing Microbial Fuel Cells.
22 Biosens. Bioelectron. 2016, 81, 103-110
23
24
25 22 Ambrosi, A.; Pumera, M. Electrochemically Exfoliated Graphene and Graphene Oxide
26 for Energy Storage and Electrochemistry Applications. Chem. - Eur. J. 2016, 22, 153-159.
27
28 23 Liu, J.; Poh, C. K.; Zhan, D.; Lai, L.; Lim, S. H.; Wang, L.; Liu, X.; Gopal Sahoo, N.;
29 Li, C.; Shen, Z.; Lin, J. Improved Synthesis of Graphene Flakes From the Multiple
30
31
Electrochemical Exfoliation of Graphite Rod. Nano Energy 2013, 2, 377-386.
32
33 24 Wu, L.; Li, W.; Li, P.; Liao, S.; Qiu, S.; Chen, M.; Guo, Y.; Li, Q.; Zhu, C.; Liu, L.
34 Powder, Paper and Foam of Few-Layer Graphene Prepared in High Yield by
35 Electrochemical Intercalation Exfoliation of Expanded Graphite. Small 2014, 10, 1421-
36
1429.
37
38
39 25 Abdelkader, A. M.; Kinloch, I. A.; Dryfe, R. A. W. High-Yield Electro-Oxidative
40 Preparation of Graphene Oxide. Chem. Commun. 2014, 50, 8402-8404.
41
42 26 Chuang, C.-H.; Su, C.-Y.; Hsu, K.-T.; Chen, C.-H.; Huang, C.-H.; Chu, C.-W.; Liu, W.-
43
44
R. A Green, Simple and Cost-Effective Approach to Synthesize High Quality Graphene by
45 Electrochemical Exfoliation Via Process Optimization. RSC Adv. 2015, 5, 54762-54768.
46
47 27 Paredes, J. I.; Munuera, J. M. Recent Advances and Energy-Related Applications of
48 High Quality/Chemically Doped Graphenes Obtained by Electrochemical Exfoliation
49
Methods. J. Mater. Chem. A 2017, 5, 7228-7242.
50
51
52 28 Chen, C.-H.; Yang, S.-W.; Chuang, M.-C.; Woon, W.-Y.; Su, C.-Y. Towards the
53 Continuous Production of High Crystallinity Graphene Via Electrochemical Exfoliation
54 with Molecular In Situ Encapsulation. Nanoscale 2015, 7, 15362-15373.
55
56
57
29 Yang, S.; Brüller, S.; Wu, Z.-S.; Liu, Z.; Parvez, K.; Dong, R.; Richard, F.; Samorì, P.;
58 Feng, X.; Müllen, K. Organic Radical-Assisted Electrochemical Exfoliation for the
59
60 40
ACS Paragon Plus Environment
Page 41 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
Scalable Production of High-Quality Graphene. J. Am. Chem. Soc. 2015, 137, 13927-
6 13932.
7
8 30 Munuera, J. M.; Paredes, J. I.; Villar-Rodil, S.; Ayán-Varela, M.; Martínez-Alonso, A.;
9 Tascón, J. M. D. Electrolytic Exfoliation of Graphite in Water with Multifunctional
10 Electrolytes: En Route Towards High Quality, Oxide-Free Graphene Flakes. Nanoscale
11
12 2016, 8, 2982-2998.
13
14 31 Gaier, J. R.; Ditmars, N. F.; Dillon, A. R. Aqueous Electrochemical Intercalation of
15 Bromine into Graphite Fibers. Carbon 2005, 43, 189-193.
16
17
18
32 Niu, L.; Li, M.; Tao, X.; Xie, Z.; Zhou, X.; Raju, A. P. A.; Young, R. J.; Zheng, Z. Salt-
19 Assisted Direct Exfoliation of Graphite Into High-Quality, Large-Size, Few-Layer
20 Graphene Sheets. Nanoscale 2013, 5, 7202-7208.
21
22 33 Munaiah, Y.; Ragupathy, P.; Pillai, V. K. Single-Step Synthesis of Halogenated
23 Graphene through Electrochemical Exfoliation and Its Utilization as Electrodes for Zinc
24
25 Bromine Redox Flow Battery. J. Electrochem. Soc. 2016, 163, A2899-A2910.
26
27 34 Ayán-Varela, M.; Paredes, J. I.; Guardia, L.; Villar-Rodil, S.; Munuera, J. M.; Díaz-
28 González, M.; Fernández-Sánchez, C.; Martínez-Alonso, A.; Tascón, J. M. D. Achieving
29 Extremely Concentrated Aqueous Dispersions of Graphene Flakes and Catalytically
30
31
Efficient Graphene-Metal Nanoparticle Hybrids with Flavin Mononucleotide as a High-
32 Performance Stabilizer. ACS Appl. Mater. Interfaces 2015, 7, 10293-10307.
33
34 35 Guardia, L.; Fernández-Merino, M. J.; Paredes, J. I.; Solís-Fernández, P.; Villar-Rodil,
35 S.; Martínez-Alonso, A.; Tascón, J. M. D. High-Throughput Production of Pristine
36
Graphene in an Aqueous Dispersion Assisted by Non-Ionic Surfactants. Carbon 2011, 49,
37
38 1653-1662.
39
40 36 Nemes-Incze, P.; Osváth, Z.; Kamarás, K.; Biró, L. P. Anomalies in Thickness
41 Measurements of Graphene and Few Layer Graphite Crystals by Tapping Mode Atomic
42 Force Microscopy. Carbon 2008, 46, 1435-1442.
43
44
45 37 Xuhua, H.; Senlin, L.; Zhiqiang, Q.; Wei, Z.; Wei, Y.; Yanyan, F. Low Defect
46 Concentration Few-Layer Graphene Using a Two-Step Electrochemical Exfoliation.
47 Nanotechnology 2015, 26, 105602-105608
48
49
38 Lou, F.; Buan, M. E. M.; Muthuswamy, N.; Walmsley, J. C.; Ronning, M.; Chen, D.
50
51 One-Step Electrochemical Synthesis of Tunable Nitrogen-Doped Graphene. J. Mater.
52 Chem. A 2016, 4, 1233-1243.
53
54 39 Wei, W.; Wang, G.; Yang, S.; Feng, X.; Müllen, K. Efficient Coupling of Nanoparticles
55 to Electrochemically Exfoliated Graphene. J. Am. Chem. Soc. 2015, 137, 5576-5581.
56
57
58
59
60 41
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 42 of 44

1
2
3
4
5
40 Mancinelli, R.; Botti, A.; Bruni, F.; Ricci, M. A.; Soper, A. K. Hydration of Sodium,
6 Potassium, and Chloride Ions in Solution and the Concept of Structure Maker/Breaker. J.
7 Phys. Chem. B 2007, 111, 13570-13577.
8
9 41 D’Angelo, P.; Migliorati, V.; Guidoni, L. Hydration Properties of the Bromide Aqua
10 Ion: the Interplay of First Principle and Classical Molecular Dynamics, and X-ray
11
12 Absorption Spectroscopy. Inorg. Chem. 2010, 49, 4224-4231.
13
14 42 Fulton, J. L.; Schenter, G. K.; Baer, M. D.; Mundy, C. J.; Dang, L. X.;
15 Balasubramanian, M. Probing the Hydration Structure of Polarizable Halides: A Multiedge
16 XAFS and Molecular Dynamics Study of the Iodide Anion. J. Phys. Chem. B 2010, 114,
17
18
12926-12937.
19
20 43 Foote, S. C.; Valentine, J. S.; Greenberg, A.; Liebman, J. F. Active Oxygen in
21 Chemistry, 1st ed.; Chapman & Hall: Glasgow, 1995.
22
23 44 Vanýsek, P., CRC Handbook of Chemistry and Physics, 93rd ed.; CRC Press: Boca
24
25 Raton, 2012.
26
27 45 Jayson, G. G.; Parsons, B. J.; Swallow, A. J. Some Simple, Highly Reactive, Inorganic
28 Chlorine Derivatives in Aqueous Solution. Their Formation Using Pulses of Radiation and
29 Their Role in the Mechanism of the Fricke Dosimeter. J. Chem. Soc., Faraday Trans.1
30
31
1973, 69, 1597-1607.
32
33 46 Zehavi, D.; Rabani, J. Oxidation of Aqueous Bromide Ions by Hydroxyl Radicals. Pulse
34 Radiolytic Investigation. J. Phys. Chem. 1972, 76, 312-319.
35
36
47 Thomas, J. K. Rates of Reaction of the Hydroxyl Radical. Trans. Faraday Soc. 1965,
37
38 61, 702-707.
39
40 48 Ferrari, A. C.; Basko, D. M. Raman Spectroscopy as a Versatile Tool for Studying the
41 Properties of Graphene. Nat. Nanotechnol. 2013, 8, 235-246.
42
43
44
49 Lotya, M.; Hernandez, Y.; King, P. J.; Smith, R. J.; Nicolosi, V.; Karlsson, L. S.;
45 Blighe, F. M.; De, S.; Wang, Z.; McGovern, I. T.; Duesberg, G. S.; Coleman, J. N. Liquid
46 Phase Production of Graphene by Exfoliation of Graphite in Surfactant/Water Solutions. J.
47 Am. Chem. Soc. 2009, 131, 3611-3620.
48
49
50 De, S.; King, P. J.; Lotya, M.; O'Neill, A.; Doherty, E. M.; Hernandez, Y.; Duesberg, G.
50
51 S.; Coleman, J. N. Flexible, Transparent, Conducting Films of Randomly Stacked
52 Graphene from Surfactant-Stabilized, Oxide-Free Graphene Dispersions. Small 2010, 6,
53 458-464.
54
55 51 Sun, Z.; Masa, J.; Liu, Z.; Schuhmann, W.; Muhler, M. Highly Concentrated Aqueous
56
57
Dispersions of Graphene Exfoliated by Sodium Taurodeoxycholate: Dispersion Behavior
58
59
60 42
ACS Paragon Plus Environment
Page 43 of 44 ACS Applied Materials & Interfaces

1
2
3
4
5
and Potential Application as a Catalyst Support for the Oxygen-Reduction Reaction. Chem.
6 - Eur. J. 2012, 18, 6972-6978.
7
8 52 Chabot, V.; Kim, B.; Sloper, B.; Tzoganakis, C.; Yu, A. High Yield Production and
9 Purification of Few Layer Graphene by Gum Arabic Assisted Physical Sonication. Sci. Rep.
10 2013, 3, 1378-1385
11
12
13 53 Perreault, F.; Fonseca de Faria, A.; Elimelech, M. Environmental Applications of
14 Graphene-Based Nanomaterials. Chem. Soc. Rev. 2015, 44, 5861-5896.
15
16 54 Raccichini, R.; Varzi, A.; Passerini, S.; Scrosati, B. The Role of Graphene for
17
18
Electrochemical Energy Storage. Nat. Mater. 2015, 14, 271-279.
19
20 55 Sherrell, P. C.; Mattevi, C. Mesoscale Design of Multifunctional 3D Graphene
21 Networks. Mater. Today 2016, 19, 428-436.
22
23 56 Chen, B.; Ma, Q.; Tan, C.; Lim, T.-T.; Huang, L.; Zhang, H. Carbon-Based Sorbents
24
25 with Three-Dimensional Architectures for Water Remediation. Small 2015, 11, 3319-3336.
26
27 57 Li, F.; Jiang, X.; Zhao, J.; Zhang, S. Graphene Oxide: A Promising Nanomaterial for
28 Energy and Environmental Applications. Nano Energy 2015, 16, 488-515.
29
30
31
58 Wang, H.; Yuan, X.; Zeng, G.; Wu, Y.; Liu, Y.; Jiang, Q.; Gu, S. Three Dimensional
32 Graphene Based Materials: Synthesis and Applications from Energy Storage and
33 Conversion to Electrochemical Sensor and Environmental Remediation. Adv. Colloid
34 Interface Sci. 2015, 221, 41-59.
35
36
59 Munuera, J. M.; Paredes, J. I.; Villar-Rodil, S.; Martínez-Alonso, A.; Tascón, J. M. D. A
37
38 Simple Strategy to Improve the Yield of Graphene Nanosheets in the Anodic Exfoliation of
39 Graphite Foil. Carbon 2017, 115, 625-628.
40
41 60 Wan, W.; Lin, Y.; Prakash, A.; Zhou, Y. Three-Dimensional Carbon-Based
42 Architectures for Oil Remediation: From Synthesis and Modification to Functionalization.
43
44
J. Mater. Chem. A 2016, 4, 18687-18705.
45
46 61 Dai, Z.; Peng, C.; Chae, J. H.; Ng, K. C.; Chen, G. Z. Cell Voltage Versus Electrode
47 Potential Range in Aqueous Supercapacitors. Sci. Rep. 2015, 5, 9854-9862.
48
49
62 Zhang, L.; Shi, G. Preparation of Highly Conductive Graphene Hydrogels for
50
51 Fabricating Supercapacitors with High Rate Capability. J. Phys. Chem. C 2011, 115,
52 17206-17212.
53
54 63 Zuo, Z.; Kim, T. Y.; Kholmanov, I.; Li, H.; Chou, H.; Li, Y. Ultra-light Hierarchical
55 Graphene Electrode for Binder-Free Supercapacitors and Lithium-Ion Battery Anodes.
56
57
Small 2015, 11, 4922-4930.
58
59
60 43
ACS Paragon Plus Environment
ACS Applied Materials & Interfaces Page 44 of 44

1
2
3
4
5
64 Choi, B. G.; Yang, M.; Hong, W. H.; Choi, J. W.; Huh, Y. S. 3D Macroporous
6 Graphene Frameworks for Supercapacitors with High Energy and Power Densities. ACS
7 Nano 2012, 6, 4020-4028.
8
9 65 Wu, Z.-S.; Winter, A.; Chen, L.; Sun, Y.; Turchanin, A.; Feng, X.; Müllen, K. Three-
10 Dimensional Nitrogen and Boron Co-doped Graphene for High-Performance All-Solid-
11
12 State Supercapacitors. Adv. Mater. 2012, 24, 5130-5135.
13
14 66 Ramadoss, A.; Yoon, K.-Y.; Kwak, M.-J.; Kim, S.-I.; Ryu, S.-T.; Jang, J.-H. Fully
15 Flexible, Lightweight, High Performance All-Solid-State Supercapacitor Based on 3-
16 Dimensional-Graphene/Graphite-Paper. J. Power Sources 2017, 337, 159-165.
17
18
19 67 Yang, X.; Zhu, J.; Qiu, L.; Li, D. Bioinspired Effective Prevention of Restacking in
20 Multilayered Graphene Films: Towards the Next Generation of High-Performance
21 Supercapacitors. Adv. Mater. 2011, 23, 2833-2838.
22
23 68 Zhang, J.; Jiang, J.; Li, H.; Zhao, X. S. A High-Performance Asymmetric
24
25 Supercapacitor Fabricated with Graphene-Based Electrodes. Energy Environ. Sci. 2011, 4,
26 4009-4015.
27
28 69 Lei, Z.; Liu, Z.; Wang, H.; Sun, X.; Lu, L.; Zhao, X. S. A High-Energy-Density
29 Supercapacitor with Graphene-CMK-5 as the Electrode and Ionic Liquid as the Electrolyte.
30
31
J. Mater. Chem. A 2013, 1, 2313-2321.
32
33 70 Enterría, M.; Martín-Jimeno, F. J.; Suárez-García, F.; Paredes, J. I.; Pereira, M. F. R.;
34 Martins, J. I.; Martínez-Alonso, A.; Tascón, J. M. D.; Figueiredo, J. L. Effect of
35 Nanostructure on the Supercapacitor Performance of Activated Carbon Xerogels Obtained
36
from Hydrothermally Carbonized Glucose-Graphene Oxide Hybrids. Carbon 2016, 105,
37
38 474-483.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 44
ACS Paragon Plus Environment

You might also like