doi: 10.1002/ ((please add manuscript number)) full paper

41
1 DOI: 10.1002/ ((please add manuscript number)) Full Paper Advanced composite twodimensional energy materials by simultaneous anodic and cathodic exfoliation Fengwang Li, Mianqi Xue, Xiaolong Zhang, Lu Chen, Gregory P. Knowles, Douglas R. MacFarlane* and Jie Zhang* F. Li, X. Zhang, Dr. L. Chen, Dr. G. P. Knowles, Prof. D. R. MacFarlane, Dr. J. Zhang School of Chemistry, Monash University, Clayton, Victoria 3800, Australia E-mail: [email protected]; [email protected] F. Li, Prof. D. R. MacFarlane, Dr. J. Zhang ARC Centre of Excellence for Electromaterials Science, School of Chemistry, Monash University, Clayton, Victoria 3800, Australia Dr. M. Xue Institute of Physics and Beijing National Laboratory for Condensed Matter Physics, Chinese Academy of Sciences, Beijing 100190, China Keywords: electrochemical exfoliation, 2D materials, composite materials, energy storage, energy conversion

Upload: others

Post on 10-May-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: DOI: 10.1002/ ((please add manuscript number)) Full Paper

1

DOI: 10.1002/ ((please add manuscript number))

Full Paper

Advanced composite two–dimensional energy materials by simultaneous anodic and

cathodic exfoliation

Fengwang Li, Mianqi Xue, Xiaolong Zhang, Lu Chen, Gregory P. Knowles, Douglas R.

MacFarlane* and Jie Zhang*

F. Li, X. Zhang, Dr. L. Chen, Dr. G. P. Knowles, Prof. D. R. MacFarlane, Dr. J. Zhang

School of Chemistry, Monash University, Clayton, Victoria 3800, Australia

E-mail: [email protected]; [email protected]

F. Li, Prof. D. R. MacFarlane, Dr. J. Zhang

ARC Centre of Excellence for Electromaterials Science, School of Chemistry, Monash

University, Clayton, Victoria 3800, Australia

Dr. M. Xue

Institute of Physics and Beijing National Laboratory for Condensed Matter Physics,

Chinese Academy of Sciences, Beijing 100190, China

Keywords: electrochemical exfoliation, 2D materials, composite materials, energy

storage, energy conversion

Page 2: DOI: 10.1002/ ((please add manuscript number)) Full Paper

2

Abstract: Composite materials based on graphene and other two–dimensional (2D)

materials are of considerable interest in the fields of catalysis, electronics and energy

conversion and storage, because of the unique structural features and electronic properties

of each component and the synergetic effects brought about by the compositing.

Approaches to the mass production of 2D materials and their composites in a facile and

affordable way are urgently needed to enable their implementation in practical applications.

Here a novel electrochemical exfoliation approach to prepare 2D composites is proposed,

which combines simultaneous anodic exfoliation of graphite and cathodic exfoliation of

other 2D materials (namely MoS2, MnO2 and graphitic carbon nitride). The synthesis is

carried out in a single–compartment electrochemical cell to in situ produce functional 2D

composite materials. Applications of the as-prepared 2D composites are demonstrated as

(i) effective hydrogen evolution catalysts and (ii) supercapacitor electrode materials. The

method enables the compositing of semiconductive, or even insulating, 2D materials with

conductive graphene in an easy, cheap, ecofriendly, yet efficient way, liberating the

intrinsic functions of 2D materials, which are usually hindered by their poor conductivity.

The method is believed to be widely applicable to the family of 2D materials.

Page 3: DOI: 10.1002/ ((please add manuscript number)) Full Paper

3

1. Introduction

Combination of different materials strategically to prepare new functional composite

materials with unique properties that are absent from each component, is central to

advanced science and technology.[1] Thus, since its discovery in 2004,[2] graphene with

advanced properties including unique structural and electronic nature, high surface area,

excellent mechanical strength, high thermal and electric conductivity, etc. has been

intensively investigated to exploit its functions as a promising building block for composite

materials.[3] Such interest has also extended to other two–dimensional (2D) materials, such

as transition metal dichalcogenides (e.g., MoS2, WSe2),[4] transition metal oxides (e.g.,

MnO2, MoO3),[5] graphitic carbon nitride (g-C3N4),

[6] hexagonal boron nitride (h–BN),[7]

transition metal carbides, carbonitrides and nitrides (MXenes),[8] elemental nanosheets

(e.g., phosphorene, silicene),[9] and metal–organic frameworks (MOFs) nanosheets,[10]

which have recently emerged as appealing functional materials for various applications in

electronics, energy and catalysis, due to their unique physical and chemical properties

compared to their three–dimensional (3D) counterparts.

However, a large proportion of these materials are typically semiconductive or even

nonconductive,[4] which hinders the migration of charge carriers in applications such as

electrochemical energy storage, electrocatalysis and photocatalysis. Therefore,

compositing these materials with conductive graphene emerges as a crucial strategy to

unlock their capabilities as functional active sites for these key applications of interest

while maintaining their 2D form. Moreover, recent advances in creating heterostructures

utilizing these 2D materials also offer researchers the ability to finely engineer the

Page 4: DOI: 10.1002/ ((please add manuscript number)) Full Paper

4

physicochemical properties of the components, stimulating synergetic effects, and, hence,

boosting the performance of the composites.[11]

Such graphene based composites are generally prepared either by blend of separate

2D materials, or in situ formation of the other composition in the presence of graphene.

Several examples reported for the latter include solvothermal reaction, electrodeposition

and chemical vapor deposition (CVD) in the presence of graphene and the precursor for

the other 2D material to be composited with the graphene.[3d,12] The former is generally

regarded as inferior since the surface of 2D materials can be easily contaminated and the

chemical and/or electronic interactions in between are weak, diminishing their reactivity

and affecting the performance of the resulting composites.[12a,13] Therefore, ideally a 2D

composite should be synthesized in situ upon the formation of its constituent 2D precursors

to minimize the possible contaminations while maximize their chemical and electronic

interaction.

Recently, electrochemical exfoliation of graphite has emerged as a new avenue to

produce high–quality graphene in a facile and environmentally friendly way.[14] The

working principle of this approach is that the intercalation of ions or charged molecules

into the interlayer space of graphite, driven by a strong local electrical field, causes

structural expansion and eventually exfoliation of the graphite.[14a] A few other 2D

materials with layered structures, such as MoS2,[15] Bi2Se3, Bi2Te3,

[16] and phosphorene,[17]

have also been obtained via this method. However, it is challenging to use this method for

the exfoliation of 2D materials with poor conductivity, since most of the applied potential

will be used to overcome the iR drop across the electrode material. This drawback severely

limits the general applicability of the method to other materials in 2D family, since a

Page 5: DOI: 10.1002/ ((please add manuscript number)) Full Paper

5

majority of 2D materials are semiconductive or nonconductive. Furthermore, all the

configurations reported so far only produced a single material in the exfoliation process

(Table S1 in the Supporting Information, SI), regardless of the potential waveform applied

(either alternating current (AC)[14j] or direct current (DC)) or nature of electrode process

(anodic, cathodic or combined[18]), rendering its powerless for the in situ production of a

2D composite.

Thus, the new concept allowing the efficient production of composite materials in a

“one–pot” process by simultaneously exfoliating different 2D material “precursors” on

each electrode was proposed in this research. This design creates the possibility for in situ

formation of novel 2D composites via the strong interactions between the simultaneously

exfoliated materials from the cathode and anode (Scheme 1). Moreover, by inclusion of

the conducting additive in the preparation of the electrode used for exfoliation, a number

of 2D materials can be exfoliated regardless of the conductivity of the bulk materials. The

method is capable, for example, of preparing 2D materials with different conductivities

(MoS2, MnO2 or g-C3N4) at the cathode, while graphene is produced at the anode and their

composites form directly in the electrolyte. The as–made composite materials demonstrate

much enhanced performances in electrocatalytic hydrogen evolution reaction (HER) when

formulated as a MoS2 nanosheet–graphene (MoS2–G) composite. Other examples

demonstrated include MnO2 nanosheet–graphene (MnO2–G) composites as excellent

supercapacitor materials and g-C3N4 nanosheet–graphene (g-C3N4–G) composites for

photocatalytic HER.

2. Results and discussion

2.1 Cathodic exfoliation of graphite

Page 6: DOI: 10.1002/ ((please add manuscript number)) Full Paper

6

Exfoliation of 2D materials at anode was the most frequently reported electrochemical

exfoliation method, which is beneficial to the dispersion of graphene in polar solvents due

to the presence of oxygen–containing groups.[14g-i] However, it unavoidably introduces

surface oxidation because of the large positive potential applied, diminishing the quality of

the materials. For example, the (partial) oxidation of MoS2 can affect the band gap[19] and

the ability to catalyze HER in acidic media.[20] Therefore, we primarily focus on accessing

the possibility of generating 2D materials from cathodic exfoliation, in which oxidation is

avoided due to the highly negative applied potential. Interestingly, despite this obvious

merit, only a few groups have reported the successful production of 2D materials via this

route.[14b-d]

To gain insights into the key parameters that govern the effective exfoliation of 2D

materials at cathode, graphite, as a model layered material, was initially investigated in

acetonitrile as solvent and with a range of salts as supporting electrolyte (0.1 M). Compared

to the aqueous electrolyte solutions that are usually used in electrochemical exfoliation in

previous reports,[14h-j] non–aqueous solvent/electrolyte solutions have much wider

electrochemical potential windows, and a greater fraction of the current is attributed to ion

intercalation, instead of water splitting as in aqueous media. The electrochemical cell is

illustrated in Figure S1 in the SI, where a glass frit is used to separate two compartments

of the cell to avoid possible influences/contaminations of the products generated from the

anode.[14f] Two identical graphite rods (Figure S2) were used as cathode and anode,

respectively, and a DC potential was applied on the electrodes to drive the exfoliation

process (see Experimental Section in the SI for detail).

Page 7: DOI: 10.1002/ ((please add manuscript number)) Full Paper

7

Graphite has been used as the negative electrode in lithium ion batteries to host

lithium between the graphene layers, and a lithium-ion battery setup was used for the

exfoliation of various 2D materials by Zhang et al.[14b] Inspired by this, LiBF4 was initially

used as electrolyte to exfoliate graphite, but only trace amounts of graphene could be

obtained (Figure 1a), which is attributed to the smaller ion diameter of Li+ (0.146 nm)[14c]

compared to the interplanar spacing of graphite. A prolonged sonication was necessary to

exfoliate the as–formed graphite intercalation complex (GIC) into graphene, as reported

previously.[14d]

When cations with larger diameters (Figure S3) were utilized, much enhanced

exfoliation yields (arbitrarily defined as the weight ratio between the ultimate nanosheet

materials after 1–hour exfoliation and the pristine material immersed into the electrolyte

and used solely as a performance indicator for experimental parameter optimization) were

achieved and Py1,4+ (1–butyl–1–methylpyrrolidinium) demonstrated the highest exfoliation

yield (~30%). A lower exfoliation efficiency was seen when either TBA+

(tetrabutylammonium) or THA+ (tetrahexylammonium) was used, which is ascribed to the

diameters being too large for facile intercalation. It should be noted that Py1,4+ is ~3 times

more efficient for exfoliation than BMIM+ (1–butyl–3–methylimidazolium), although they

have the same crystallographic diameter. Actually, the exfoliation ability of BMIM+ is even

weaker than that of TMA+ (tetramethylammonium), which has a smaller crystallographic

diameter. The strong π–π interaction between BMIM+ and graphite basal plane enables the

face–to–face configuration between them, and thus may weaken the exfoliation ability of

BMIM+.[14f,21]

Page 8: DOI: 10.1002/ ((please add manuscript number)) Full Paper

8

The fact that the rate of intercalation is correlated to the diameter of cation offers the

possibility to exfoliate 2D materials with different interplanar spacing by choosing

electrolytes with optimal sized cations. Furthermore, as expected, graphene generated via

cathodic exfoliation has much fewer defects (the ID/IG in the Raman spectra is less than

0.25, Figure 1b). In addition, increasing the applied potential also speeds up the exfoliation

rate (Figure 1b).

2.2 Cathodic exfoliation of MoS2 crystal

To probe the feasibility of cathodic exfoliation process for other materials, a piece of

natural MoS2 crystal (Figure 2a and S4) was directly mounted on the clamp of the cathode

and immersed in acetonitrile (0.1 M [Py1,4][BF4]). No obvious exfoliation of the crystal

was observed upon a DC potential of –10 V, even after 1 hour. This is not surprising given

that the interplanar spacing in MoS2 crystal structure (0.609 nm) is much larger than that

of graphite, hence needing a larger cation to drive exfoliation. By replacing Py1,4+ with

TBA+, which has larger diameter of 0.826 nm, the crystal could be expanded within 10

minutes, clearly visualized by eyes and revealed from scanning electron microscopic

(SEM) images (Figure 2b,c and S4). In the XRD patterns, the (002) diffraction peak of the

expanded MoS2 is significantly smaller than that of the pristine crystal (Figure 2d),

indicating that a great portion of the crystal has been successfully expanded.[15] Some flakes

were observed during the exfoliation process in 30 minutes but the majority of the

expanded crystal was still physically connected, which is different from the case of graphite

rod. This phenomenon is similar to the exfoliation of highly oriented pyrolytic graphite

(HOPG) under similar conditions[14c] and is probably due to the strong bonding interaction

among the basal planes of MoS2 crystal. Sonicating the expanded crystal for 2 hours

Page 9: DOI: 10.1002/ ((please add manuscript number)) Full Paper

9

facilitated the final exfoliation. The dispersion of the as–exfoliated MoS2 crystal displayed

the Tyndall effect even after 3 months (Figure 2e), indicating the colloidal state and good

stability of as–prepared MoS2 nanosheets.

Figure 2f depicts a transmission electron microscopic (TEM) image of a typical thin

MoS2 nanosheet of about 5 μm × 3 μm in lateral size. It is interesting to note that the lateral

size of the exfoliated MoS2 could reach tens of micrometers (Figure S4) if exfoliation

parameters are further optimized. The reduction of size under present conditions could be

caused by the long periods of sonication. High–resolution TEM (HRTEM) image (Figure

2g) reveals the hexagonal lattice structure of the MoS2 nanosheet and the corresponding

fast Fourier transform (FFT) pattern shows the hexagonal symmetry characteristic of 2H–

MoS2.[15] The Raman spectrum (Figure 2h) presents two peaks, attributed to the in–plane

E12g mode (383 cm–1) and out–of–plane A1g mode (409 cm–1), which is consistent with

mechanically exfoliated 2H–MoS2 nanosheets.[4] The HRTEM and Raman spectrum

results indicate that the MoS2 nanosheets obtained via cathodic exfoliation exhibit high

crystallinity with no detectable defects. This approach of production of MoS2 represents a

novel method of production not otherwise previously described, and it is advantageous in

that high quality MoS2 nanosheets can be obtained in an efficient way with fewer defects

than typically caused by anodic oxidation.[15]

2.3 Electrochemical exfoliation of MoS2 and graphene composite

MoS2 powder with size of 1–5 μm (Figure S5) was further used as the cathode material for

electrochemical exfoliation. A consideration in using the powder instead of the single

crystal is that the electrocatalytic activity of MoS2 in the HER is correlated to the total

length of the exposed edges, rather than the total surface area.[22] Therefore, smaller size of

Page 10: DOI: 10.1002/ ((please add manuscript number)) Full Paper

10

MoS2 is expected to result in higher performance in the composite material. Since the

electrical conductivity of bulk MoS2 is poor (band gap of 1.2 eV), to fabricate MoS2

cathode, 25 wt% of graphite powder was added to MoS2 powder to enhance the

conductivity (Figure S6). This mixed powder was then compressed into a tablet (13 mm in

diameter, ~0.5 mm in thickness) to serve as the electrode. It should be emphasized that the

addition of conducting material is essential and important. A great number of 2D materials

are semiconductive or even nonconductive, which hinders the general application of

electrochemical exfoliation method, because a large applied potential is needed to

overcome the huge resistance. Inclusion of the conducting additive makes the exfoliation

of an extensive range of 2D materials possible regardless of the conductivity of the bulk

materials (vide infra).

The exfoliation efficiency of a MoS2 powder cathode is also dependent on the size

of electrolyte cation, as expected (Figure 3a). TBA+ has an appropriate size compared to

the interplanar spacing of MoS2 (0.618 nm, Figure S5), giving the best exfoliation rate

(more than 60 %) at an applied potential of –10 V. The exfoliation efficiency can also be

modulated by the potential applied (Figure 3b). TEM images (Figure 3c and S7)

demonstrate the nanosheet morphology of the as–exfoliated MoS2 with an average lateral

size of ~235 nm. The atomic force microscopic (AFM) image (Figure 3d) of a typical piece

of the MoS2 nanosheets shows a similar shape to the TEM image, with a thickness of ~ 2

nm (Figure S7). More than 75% of the nanosheets are less than 2.3 nm in thickness (Figure

S7), consistent with there being around 2 layers and implying a successful exfoliation.

A MoS2–G composite was readily produced in situ when the two–compartment

electrochemical cell was replaced by the single–compartment cell illustrated in Scheme 1.

Page 11: DOI: 10.1002/ ((please add manuscript number)) Full Paper

11

Graphene sheets were generated at the anode (refer to Supplementary Note in the SI for the

choice of electrolyte and applied potential to achieve optimal exfoliation yield and quality)

and can be expected to carry a positive charge, while MoS2 is exfoliated at the cathode and

is expected to be negatively charged. Hence, electrostatic interaction provides the major

driving force for the formation of a composite material, which will precipitate in the

electrolyte and can be re–dispersed by mild sonication after washing. Not only the yield,

but also the ratio between MoS2 and graphene can be tuned by the potential applied (Figure

3e). This is of importance to the application of the composite materials, since the ratio

between each functional component plays a key role in optimizing the performance of the

eventual material.[1,12b] The TEM image in Figure 3f shows the MoS2 basal plane

overlapping and stacking with the basal plane of graphene and no aggregation was

observed. Such a configuration provides bridging graphene sheets to promote fast electron

communication between the MoS2 nanosheets (both the basal plane and the edges) and the

substrate electrode.[12,23]

The electrocatalytic HER activity of the MoS2–G composites was investigated in 0.5

M H2SO4 solution by drop-casting the composite material onto a glassy carbon electrode.

Pt was used as a benchmark catalyst which exhibits a near zero overpotential for HER in

the strong acidic medium. The polarization curves (where current density is calculated

based on the geometric area of the electrode) recorded are shown in Figure 3g. The MoS2–

G composite material with the optimum composition (graphene 38.3 wt%, obtained by

exfoliation at 10 V, see Figure S8a for the HER performance of composites with different

MoS2 and graphene ratio) requires an HER overpotential of only 125 mV to reach a current

density of 10 mA cm–2. This places the material among the best MoS2–carbon composite

Page 12: DOI: 10.1002/ ((please add manuscript number)) Full Paper

12

based catalysts for HER (Table S2). By contrast, bare electrochemical exfoliated MoS2

nanosheets exhibited performance inferior to that of MoS2–G, despite the improvement

compared to that of bulk MoS2 powder. MoS2 nanosheets physically mixed with exfoliated

graphene also showed inferior performance compared to the in situ formed MoS2–G

composite (Figure S8b), highlighting the advantage of the in situ formation method in

achieving unique composite structures and superior properties.

The Tafel slopes for bulk MoS2 powder, MoS2 nanosheets and MoS2–G composite

are 128, 64 and 41 mV dec–1, respectively (Figure 3h). The value of 41 mV dec–1 suggests

a Volmer–Heyrovsky HER mechanism for MoS2–G.[12a] Smaller Tafel slopes are

beneficial to practical applications, since they lead to higher HER rates at any given

overpotential.[24] The high activity of MoS2–G can be ascribed to the face–to–face layering

of the MoS2 nanosheets and graphene (Figure S9), which leads to the electronic interaction

between the basal plane of MoS2 and graphene. A redshift of the E12g and A1g bands in the

Raman spectrum was observed in the MoS2–G sample, which was assigned to be the n–

doping effect of carbon materials to the MoS2 nanosheets.[25] This electronic interaction

between graphene and MoS2 could tune the adsorption properties of the catalyst and

therefore improve catalytic kinetics, as have been reported for enhanced HER, CO2

reduction and CO oxidation activities from graphene supported nanoparticle

catalysts.[12a,26] Additionally, the composite catalyst also presents high durability, with

negligible activity loss after 24 hours (Figure S10), resulting from the interaction between

MoS2 nanosheet and graphene substrate, which can minimize the mechanical loss of the

active materials.[27]

2.4 Electrochemical exfoliation of MnO2 and graphene composite

Page 13: DOI: 10.1002/ ((please add manuscript number)) Full Paper

13

The proposed method herein possesses the capability of producing other 2D composite

materials. MnO2 is one of the most promising electroactive materials as it is cheap, Earth

abundant, environmental friendly, and most importantly, has high theoretical specific

capacitance (~1300 F g–1).[28] However, the pseudo–capacitive reaction of MnO2 only

occurs on the surface or a very thin surface layer of the oxide due to the poor electron and

ion mobility in MnO2.[29] Therefore, generating MnO2 with nanoscopic dimensions and

integrating it with a high–surface–area conductive support is highly desirable to optimize

the electrochemical performance of MnO2. Birnessite–type δ-MnO2 has layered structure

and can be delaminated by the intercalation of quaternary ammonium ions.[30] Herein, we

show that δ-MnO2 can be effectively exfoliated by cathodic intercalation and in situ

formation of MnO2–G composites is demonstrated with graphene exfoliated from graphite

at the anode. δ-MnO2 has similar interplanar spacing (0.694 nm, Figure S11) to MoS2;

therefore, the TBA+ cation once again displays the most powerful exfoliation ability

compared to other cations (Figure 4a). The TEM (Figure 4b) image indicates the nanosheet

morphology of the as–exfoliated MnO2 material with an average lateral size of 205 nm

(Figure S12a); and the average height is 5 nm, as revealed by AFM height profile (Figure

S12b).

When coupled with the anodic exfoliation of graphene from a graphite anode, a

MnO2–G composite was obtained. The ratio between graphene and MnO2 can be tuned by

the applied potential (Figure 4c), similar to the case of MoS2–G. The composite can be

easily dispersed in water and forms a uniform film via simple vacuum filtration (Figure

4d), which can be easily peeled off or transferred to other substrates (such as silicon wafer,

polymer film, etc.). SEM and energy–dispersive X–ray spectroscopy (EDS) mapping

Page 14: DOI: 10.1002/ ((please add manuscript number)) Full Paper

14

results (Figure S13) demonstrate the uniform distribution of MnO2 on graphene. The

thickness can be adjusted by the volume of the dispersion used for filtration. Figure 4e and

f depict the cross–sectional SEM images of a typical film with a thickness of ~4 μm. The

wavy graphene in each layer enlarges the interlayer spacing and provides numerous

accessible pathways for ions.

The film was transferred to a flexible polyethylene terephthalate (PET) film and

assembled to a planar, symmetric all–solid–state supercapacitor (Figure S14) using

polyvinyl alcohol (PVA)/H3PO4 gel electrolyte. The performance of the device was

evaluated by galvanostatic charge–discharge cycling at different current densities (Figure

4g). The specific capacitance of the device derived from the discharge curves is 265 F g–1

at a current density of 2.5 A g–1 and drops to 212 F g–1 at 50 A g–1, indicating only 20%

capacitance loss when the current density increases by a factor of 20 (Figure S15). As

shown in the Ragone plot (Figure S15), the device delivers a high energy density of 20.7

Wh kg–1 at a power density of 2570 W kg–1, which are comparable to, or even better than,

previously reported state–of–the–art supercapacitors based on MnO2–carbon composites

(Table S3). The cycling test of the device shows capacitance retention of ~94% over 5,000

cycles (Figure S15). The good rate capability and cycling stability can be attributed to a

synergistic effect of the material composition and device configuration: the tight interaction

between 2D MnO2 nanosheets and graphene can significantly promote the electrical

communication between the electroactive MnO2 and the conductive substrate,[29] while the

planar 2D device enables fast ion transport to potentially all of the active sites in the

horizontal plane.[28] The larger interplanar spacing within the composite film also facilitates

the ion transport.[31] In addition, thanks to the solid–state electrolyte, the as–fabricated

Page 15: DOI: 10.1002/ ((please add manuscript number)) Full Paper

15

supercapacitor presents only slight specific capacitance change under folded or rolled

conditions (Figure 4h), indicating the flexibility of the device.

2.5 Electrochemical exfoliation of g-C3N4 and graphene composite

To further illustrate the generality of the approach, we demonstrate that g-C3N4, with

layered structure (interplanar spacing of 0.322 nm, Figure S16), can be exfoliated at the

cathode using Py1,4+ as the intercalation cation (average lateral size of 137 nm and ~60%

of 4–6 nm in height, Figure 5a-e, Figure S17), and g-C3N4 and graphene composite (g-

C3N4–G) can be in situ formed when combined with anodic graphite exfoliation (Figure

5f,g). The as–prepared g-C3N4 nanosheets showed enhanced performance towards

photocatalytic HER under visible light irradiation compared to the bulk g-C3N4 (Figure

5h). As expected, the in situ formed g-C3N4 and graphene composite showed a further

enhancement of the performance, benefiting from the strong electronic interaction between

them. The photo–generated electrons on the conduction band of g-C3N4 can transfer to

graphene, resulting in an enhanced hole–electron separation and diminished carrier

recombination.[32]

3. Conclusion

Our work describes a simultaneous electrochemical exfoliation method at both anode and

cathode in a single–compartment electrochemical cell for the production of high–

performance 2D composite materials. The process is easy, cheap, ecofriendly, and efficient

and is intrinsically scalable for large–scale processing. The method enables the

compositing of semiconductive and even insulating 2D materials with graphene, reducing

the possibility of contaminations, and allowing access to their intrinsic catalytic properties,

which are usually hindered by poor conductivity. Such an in situ formation approach

Page 16: DOI: 10.1002/ ((please add manuscript number)) Full Paper

16

therefore offers the advantage of enhanced performance, compared to the physical mixing

of single components of the composite, for applications in energy fields such as electro–

/photo–catalysis and energy storage. Our method is applicable to generating 2D nanosheets

and their composites from a great range of materials with layered structures, which will

contribute to numerous applications, including catalysis, energy conversion and storage,

and flexible electronics.

Supporting Information

Supporting Information is available from the Wiley Online Library or from the author.

Acknowledgements

This research was supported by the Australian Research Council (ARC) through the ARC

Centre of Excellence for Electromaterials Science. F.L. acknowledges Monash University

for his Postgraduate Publication Award. D.R.M. is grateful to the ARC for his Australian

Laureate Fellowship. M.X. acknowledges Natural Science Foundation of China

(21622407) for financial support. F.L. thanks Prof. David Officer (University of

Wollongong) for provision of natural MoS2 crystal used in this study. The authors

acknowledge the use of facilities within the Monash Centre for Electron Microscopy and

the Monash X–Ray Platform.

Received: ((will be filled in by the editorial staff))

Revised: ((will be filled in by the editorial staff))

Published online: ((will be filled in by the editorial staff))

Page 17: DOI: 10.1002/ ((please add manuscript number)) Full Paper

17

References

[1] D. D. Chung, Composite Materials: Functional Materials for Modern Technologies,

Springer-Verlag, London, 2003.

[2] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos,

I. V. Grigorieva, A. A. Firsov, Science 2004, 306, 666.

[3] a) F. Bonaccorso, L. Colombo, G. Yu, M. Stoller, V. Tozzini, A. C. Ferrari, R. S.

Ruoff, V. Pellegrini, Science 2015, 347, 1246501; b) R. Raccichini, A. Varzi, S.

Passerini, B. Scrosati, Nat. Mater. 2015, 14, 271; c) H. Bai, C. Li, G. Shi, Adv.

Mater. 2011, 23, 1089; d) N. Gao, X. Fang, Chem. Rev. 2015, 115, 8294.

[4] M. Chhowalla, H. S. Shin, G. Eda, L. J. Li, K. P. Loh, H. Zhang, Nat. Chem. 2013,

5, 263.

[5] R. Ma, T. Sasaki, Acc. Chem. Res. 2015, 48, 136.

[6] X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen,

M. Antonietti, Nat. Mater. 2009, 8, 76.

[7] L. Song, L. Ci, H. Lu, P. B. Sorokin, C. Jin, J. Ni, A. G. Kvashnin, D. G. Kvashnin,

J. Lou, B. I. Yakobson, P. M. Ajayan, Nano Lett. 2010, 10, 3209.

[8] B. Anasori, M. R. Lukatskaya, Y. Gogotsi, Nat. Rev. Mater. 2017, 2, 16098.

[9] A. J. Mannix, B. Kiraly, M. C. Hersam, N. P. Guisinger, Nat. Rev. Chem. 2017, 1,

0014.

[10] T. Rodenas, I. Luz, G. Prieto, B. Seoane, H. Miro, A. Corma, F. Kapteijn, F. X.

Llabres i Xamena, J. Gascon, Nat. Mater. 2015, 14, 48.

[11] a) Y. P. Zhu, C. Guo, Y. Zheng, S. Z. Qiao, Acc. Chem. Res. 2017, 50, 915; b) Y.

Sun, S. Gao, F. Lei, Y. Xie, Chem. Soc. Rev. 2015, 44, 623.

[12] a) Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong, H. Dai, J. Am. Chem. Soc. 2011, 133,

7296; b) D. Deng, K. S. Novoselov, Q. Fu, N. Zheng, Z. Tian, X. Bao, Nat.

Nanotechnol. 2016, 11, 218.

[13] C. G. Williams, M. A. Edwards, A. L. Colley, J. V. Macpherson, P. R. Unwin, Anal.

Chem. 2009, 81, 2486.

[14] a) A. M. Abdelkader, A. J. Cooper, R. A. Dryfe, I. A. Kinloch, Nanoscale 2015, 7,

6944; b) Z. Zeng, T. Sun, J. Zhu, X. Huang, Z. Yin, G. Lu, Z. Fan, Q. Yan, H. H.

Hng, H. Zhang, Angew. Chem. Int. Ed. 2012, 51, 9052; c) A. J. Cooper, N. R.

Wilson, I. A. Kinloch, R. A. W. Dryfe, Carbon 2014, 66, 340; d) J. Wang, K. K.

Manga, Q. Bao, K. P. Loh, J. Am. Chem. Soc. 2011, 133, 8888; e) Y. L. Zhong, T.

M. Swager, J. Am. Chem. Soc. 2012, 134, 17896; f) A. T. Najafabadi, E. Gyenge,

Carbon 2014, 71, 58; g) C. Y. Su, A. Y. Lu, Y. Xu, F. R. Chen, A. N. Khlobystov,

L. J. Li, ACS Nano 2011, 5, 2332; h) K. Parvez, R. Li, S. R. Puniredd, Y. Hernandez,

F. Hinkel, S. Wang, X. Feng, K. Mullen, ACS Nano 2013, 7, 3598; i) K. Parvez, Z.

S. Wu, R. Li, X. Liu, R. Graf, X. Feng, K. Mullen, J. Am. Chem. Soc. 2014, 136,

6083; j) S. Yang, A. G. Ricciardulli, S. Liu, R. Dong, M. R. Lohe, A. Becker, M.

A. Squillaci, P. Samori, K. Mullen, X. Feng, Angew. Chem. Int. Ed. 2017, 56, 6669.

[15] N. Liu, P. Kim, J. H. Kim, J. H. Ye, S. Kim, C. J. Lee, ACS Nano 2014, 8, 6902.

[16] A. Ambrosi, Z. Sofer, J. Luxa, M. Pumera, ACS Nano 2016, 10, 11442.

[17] M. B. Erande, S. R. Suryawanshi, M. A. More, D. J. Late, Eur. J. Inorg. Chem.

2015, 2015, 3102.

[18] A. T. Najafabadi, E. Gyenge, Carbon 2015, 84, 449.

Page 18: DOI: 10.1002/ ((please add manuscript number)) Full Paper

18

[19] S. H. Song, B. H. Kim, D. H. Choe, J. Kim, D. C. Kim, D. J. Lee, J. M. Kim, K. J.

Chang, S. Jeon, Adv. Mater. 2015, 27, 3152.

[20] X. Chia, A. Y. Eng, A. Ambrosi, S. M. Tan, M. Pumera, Chem. Rev. 2015, 115,

11941.

[21] J. Lu, J. X. Yang, J. Wang, A. Lim, S. Wang, K. P. Loh, ACS Nano 2009, 3, 2367.

[22] T. F. Jaramillo, K. P. Jorgensen, J. Bonde, J. H. Nielsen, S. Horch, I. Chorkendorff,

Science 2007, 317, 100.

[23] D. Voiry, R. Fullon, J. Yang, E. S. C. de Carvalho Castro, R. Kappera, I. Bozkurt,

D. Kaplan, M. J. Lagos, P. E. Batson, G. Gupta, A. D. Mohite, L. Dong, D. Er, V.

B. Shenoy, T. Asefa, M. Chhowalla, Nat. Mater. 2016, 15, 1003.

[24] J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou, X. W. Lou, Y. Xie,

Adv. Mater. 2013, 25, 5807.

[25] Z. Li, R. Ye, R. Feng, Y. Kang, X. Zhu, J. M. Tour, Z. Fang, Adv. Mater. 2015, 27,

5235.

[26] a) S. Zhang, P. Kang, T. J. Meyer, J. Am. Chem. Soc. 2014, 136, 1734; b) E. Yoo,

T. Okata, T. Akita, M. Kohyama, J. Nakamura, I. Honma, Nano Lett. 2009, 9, 2255.

[27] a) A. B. Laursen, P. C. Vesborg, I. Chorkendorff, Chem. Commun. 2013, 49, 4965;

b) J. D. Benck, Z. Chen, L. Y. Kuritzky, A. J. Forman, T. F. Jaramillo, ACS Catal.

2012, 2, 1916.

[28] L. Peng, X. Peng, B. Liu, C. Wu, Y. Xie, G. Yu, Nano Lett. 2013, 13, 2151.

[29] J. Yan, Z. Fan, T. Wei, W. Qian, M. Zhang, F. Wei, Carbon 2010, 48, 3825.

[30] Y. Omomo, T. Sasaki, L. Wang, M. Watanabe, J. Am. Chem. Soc. 2003, 125, 3568.

[31] X. Yang, C. Cheng, Y. Wang, L. Qiu, D. Li, Science 2013, 341, 534.

[32] Q. Xiang, J. Yu, M. Jaroniec, J. Phys. Chem. C 2011, 115, 7355.

Page 19: DOI: 10.1002/ ((please add manuscript number)) Full Paper

19

Scheme 1. The electrochemical exfoliation configuration, where graphene is generated from

anode while another 2D material is generated at the cathode simultaneously, and the as–

generated two kinds of nanosheets in situ form the composites in the solution due to electrostatic

interactions.

Figure 1. Cathodic exfoliation of graphite. (a) The exfoliation yields when electrolyte with

different ion diameter was used. (b) The yield and quality (revealed by the intensity ratio

between the D and G bands in the Raman spectrum, where a lower ratio corresponds to fewer

defects) of graphene when different potentials were applied.

Page 20: DOI: 10.1002/ ((please add manuscript number)) Full Paper

20

Figure 2. Cathodic exfoliation of MoS2 crystal. Photographs of a piece of MoS2 crystal before

(a) and after (b) expansion. (c) SEM image of the expanded crystal after applying –10 V for 10

minutes in acetonitrile with 0.1 M [TBA][PF6]. (d) XRD patterns of the crystal before and after

expansion. (e) Photograph of MoS2 dispersion in ethanol/water (1:1 v/v) after 3 months,

displaying the Tyndall effect. TEM (f), HREM (g) and Raman spectrum (h) of the exfoliated

MoS2. Insert in (g) is the corresponding FFT pattern, scale bar: 2 1/nm.

Figure 3. In situ formation of MoS2–graphene composite via electrochemical exfoliation and

its performance for HER electrocatalysis. (a) Cathodic exfoliation yield of MoS2 powder in

acetonitrile at –10 V with different electrolytes. (b) The yields of MoS2 in 0.1 M [TBA][PF6]

in acetonitrile at different applied potentials. TEM (c) and AFM (d) images of as–exfoliated

MoS2 nanosheets. (e) Yield of the MoS2–G composite and the ratio between graphene and

MoS2. (f) TEM image of MoS2–G. (g) Polarization curves of Pt, MoS2–G, bare MoS2

nanosheets (MoS2 NS) and bulk MoS2 powder in 0.5 M H2SO4 at a scan rate of 2 mV s–1 with

iR compensation. (h) Tafel plots derived from (f).

Page 21: DOI: 10.1002/ ((please add manuscript number)) Full Paper

21

Figure 4. In situ formation of MnO2–graphene composite via electrochemical exfoliation and

its performance in supercapacitors. (a) Cathodic exfoliation yield of MnO2 in acetonitrile at –

10 V with different electrolytes (0.1 M). (b) SEM image of the exfoliated MnO2 nanosheets. (c)

Yield of the MnO2–G composite and the ratio between graphene and MnO2. (d) Photograph of

the MnO2–G film produced by vacuum filtration. (e, f) Cross–sectional SEM images of the

MnO2–G film. (g) Galvanostatic charge–discharge curves of the planar, symmetric all–solid–

state supercapacitor device at different current densities. (h) Galvanostatic charge–discharge

curves of the device at a current density of 2.5 A g–1 at different mechanical conditions.

Figure 5. In situ formation of g-C3N4–graphene composite via electrochemical exfoliation and

its performance in photocatalytic HER. (a) cathodic exfoliation yield of g-C3N4 in acetonitrile

at –10 V with different electrolytes. (b) The yields of in 0.1 M [Py1,4][BF4] in acetonitrile at

different applied potentials. (c,d) TEM image of the exfoliated g-C3N4 nanosheets. (e) Statistics

of the lateral size of 100 pieces of g-C3N4 nanosheets in the TEM images. (f) The yield of the

g-C3N4–G composite and the ratio between graphene and g-C3N4. (g) TEM image of the g-

C3N4–G composite. (h) H2 evolution curves for bulk g-C3N4, cathodic exfoliated g-C3N4

nanosheets and g-C3N4–G composite and the H2 evolution rate is 276, 1137, 2749 μmol g–1 h–

1, respectively. The weight ratio of graphene in the g-C3N4–G composite was 23% obtained by

exfoliation at 7 V.

Page 22: DOI: 10.1002/ ((please add manuscript number)) Full Paper

22

Functional 2D composite materials can be generated in situ by a simultaneous cathodic and

anodic exfoliation method in a single-compartment electrochemical cell regardless of the low

electrical conductivity of parent 2D materials. The formed 2D composites show excellent

performance for electro- and photo-chemical energy conversion and storage applications.

Keywords: electrochemical exfoliation, 2D materials, composite materials, energy storage,

energy conversion

F. Li, M. Xue, X. Zhang, L. Chen, G. P. Knowles, D. R. MacFarlane*, J. Zhang*

Advanced composite two–dimensional energy materials by simultaneous anodic and

cathodic exfoliation

ToC figure

Page 23: DOI: 10.1002/ ((please add manuscript number)) Full Paper

23

Copyright WILEY-VCH Verlag GmbH & Co. KGaA, 69469 Weinheim, Germany, 2016.

Supporting Information

Advanced composite two–dimensional energy materials by simultaneous anodic and

cathodic exfoliation

Fengwang Li, Mianqi Xue, Xiaolong Zhang, Lu Chen, Gregory P. Knowles, Douglas R.

MacFarlane* and Jie Zhang*

1. Experimental Section

1.1 Chemicals and apparatus

LiBF4, 1–ethyl–3–methylimidazolium tetrafluoroborate (EMIMI–BF4, ≥98.0%),

tetramethylammonium tetrafluoroborate ([TMA][BF4], 97%), tetrahexylammonium

hexafluorophosphate ([THA][PF6], ≥97%), N–methyl–2–pyrrolidone (NMP), graphite powder

(<150 μm, 99.99% trace metals basis), Molybdenum (IV) sulfide (MoS2, powder), Lead powder

(~100 mesh, 99.95%), dicyandiamide, chloroplatinic acid hydrate (H2PtCl6·xH2O, ~38% Pt

basis), poly(vinyl alcohol) (PVA, 87–89% hydrolyzed, average Mw 31,000–50,000) and Nafion

(5 wt%) were purchased from Sigma–Aldrich. Natural MoS2 crystal was purchased from

Manchester Nanomaterials Ltd (U.K.). 1–butyl–3–methylimidazolium tetrafluoroborate

([BMIM][BF4], 99%), 1–butyl–1–methylpyrrolidinium tetrafluoroborate ([Py1,4][BF4], 99%),

1–butyl–3–methylimidazolium hexafluorophosphate ([Py1,4][PF6], 99%), 1–butyl–1–

methylpyrrolidinium bis(trifluoromethylsulfonyl)imide ([Py1,4][TFSI], 99%) were purchased

from IOLITEC GmbH. KMnO4 was purchased from Ajax Finechem. Ethanol (96%),

acetonitrile (97%), H2SO4 (95–98%), H3PO4 (≥99%) were purchased from Merck.

Tetrabutylammonium hexafluorophosphate ([TBA][PF6], 98%) was purchased from Wako

(Japan) and recrystallized twice from ethanol before use. All other chemicals were used without

further purification. Deionized water was used for the preparation of all aqueous solutions.

DC regulated power supply (voltage range: 0–30 V; current range: 0–5 A, Jaycar electronics)

was used for all the electrochemical exfoliation process. All other electrochemical experiments

were conducted on a CHI 760E electrochemical workstation (CH Instruments, USA) at room

temperature (22 ± 2 °C). A 300 W Xe lamp (MicroSolar300, PerfectLight, Beijing) through a

UV–cutoff filter (>400 nm) was used as the light source (15 cm away from the reactor). Raman

spectra was obtained using a Renishaw inVia Microscope with a 514–nm laser source and X–

Page 24: DOI: 10.1002/ ((please add manuscript number)) Full Paper

24

ray Diffraction (XRD) data were collected with a Bruker D2 PHASER powder diffractometer

(Cu Kα radiation, λ = 0.15406 nm). Transmission electron microscopic (TEM) images were

collected on a FEI Tecnai G2 T20 TWIN TEM. Scanning electron microscopic (SEM) images

and energy dispersive spectrum (EDS) were recorded on a FEI Nova NanoSEM 450 FEG SEM

equipped with Bruker Quantax 400 X–ray analysis system. Gas chromatography (GC) was

performed with an Agilent 7820 A gas chromatography system equipped with a HP–plot

molesieve (5A) column and a thermal conductivity detector (TCD). The carrier gas was

nitrogen (99.99%).

1.2 Materials synthesis

δ-MnO2 was synthesized by calcinating KMnO4 at 600 °C in air atmosphere for 5 hours, as

previously reported.[1] After cooling down naturally, the powder was washed thoroughly with

water until the purple color disappeared, and then dried at 80 °C overnight. g-C3N4 was

synthesized by calcinating dicyandiamide at 550 °C for 4 hours under N2 atmosphere, as

previously reported.[2] After cooling down naturally, the large powder was ground into fine

powder for further use.

All the exfoliation processes were carried out in a two–electrode system: two–compartment

electrochemical cell (Figure S1) was used to study the parameters for single material exfoliation

and preparation; a single compartment electrochemical cell (Scheme 1) was used for the

preparation of the composite materials from exfoliation at both electrodes. The electrodes were

placed in parallel at a fixed distance of ~1.5 cm. The solution was acetonitrile with 0.1 M salt

as supporting electrolyte. A direct current voltage was used applied for 0.5 h for the exfoliation.

The current magnitude ranges from 0.04 A to 0.2 A. Graphite rod (3 mm in diameter) was used

as anode for all the exfoliation experiments. To exfoliate graphite at cathode, graphite rod was

also used as cathode. To exfoliate MoS2 and MnO2, a tablet–type electrode of 13 mm in

diameter was used as cathode. The tablet was prepared via a Specac 15 Ton Manually Operated

Hydraulic Press in conjunction with 13 mm Specac Evacuable Pellet Die with an applied load

of 8 ton for a period of 20 minutes on the powder consisting of 0.2 g corresponding material

and 0.05 g graphite powder. To prepare the g-C3N4 tablet, graphite was replaced by lead powder

and other conditions were kept the same. Once the exfoliation was finished, all the materials in

the electrochemical cell were centrifuged at 5,000 rpm for 3 minutes, washed with acetonitrile

and recollected by centrifugation. The washing step was repeated for at least 5 times to ensure

that the salts were entirely removed. Afterwards, the obtained powder was dispersed in a solvent

(NMP for graphene via cathodic exfoliation, ethanol for MoS2 and water for all other materials)

via mild sonication for 30 min. The large particles were removed by centrifugation at 3,000

Page 25: DOI: 10.1002/ ((please add manuscript number)) Full Paper

25

rpm for 5 minutes. The final products were obtained by centrifugation of the supernatant at

12,000 rpm for 20 min and drying at 80 °C overnight. The ratio between graphene and other

materials in the composites was determined by elemental analysis in EDS.

1.3 Electrocatalytic hydrogen evolution reaction measurement

4 mg catalyst and 40 µl of 5 wt% Nafion solution were dispersed in 1 ml water/ethanol (4:1

v/v) mixture by sonication for 30 min to form a homogeneous ink. Then 5 μl of the catalyst ink

was pipetted onto a pre–polished glassy carbon electrode of 3 mm in diameter and dried under

an infrared lamp. Linear sweep voltammetry was carried out in 0.5 M H2SO4 at a scan rate of 2

mV s–1 with iR compensation. A graphite rod counter electrode, an Ag/AgCl (3 M KCl)

reference electrode and the glassy carbon electrode working electrode were used for the

measurements. Sample with different ratio between graphene and MoS2 was tested and the

optimum one was determined by the smallest overpotential to reach 10 mA cm–2. To convert

the potentials (E) to the reversible hydrogen electrode (RHE) reference scale, the following

formula was used.

E (vs. RHE) = E (vs. Ag/AgCl) + 0.210 V + 0.059 V × pH

1.4 Supercapacitor assembly and measurement

The optimum ratio between graphene and MnO2 was determined by cyclic voltammetry (CV).

Briefly, 4 mg sample and 40 µl of 5 wt% Nafion solution were dispersed in 1 ml water/ethanol

(4:1 v/v) mixture by sonication for 30 min to form a homogeneous ink. Then 5 μl of the ink was

pipetted onto a pre–polished glassy carbon electrode of 3 mm in diameter and dried under an

infrared lamp. CV was carried out in 1.0 M Na2SO4 at a scan rate of 50 mVs–1 until a stable

trace was obtained. A Pt wire counter electrode, an Ag/AgCl (3 M KCl) reference electrode and

the glassy carbon electrode working electrode modified with the sample were used in the test.

The optimum sample was determined by the largest area of the CV curve, and was used for

further device assembly and measurement.

To obtain the MnO2–G film, 30 ml MnO2–G dispersion (0.4 mg mL–1) was vacuum filtered

through a mixed cellulose ester filter membrane (0.05 μm pore size, 50 mm in diameter). The

vacuum was disconnected immediately once no free dispersion was left on the filtrate cake. The

filter membrane with MnO2–G film on it was cut into the desired size and transferred to a PET

substrate by gently applying a small force with a glass rod. Then the filter membrane was slowly

peeled off. The density of the MnO2–G film after fully drying is ~1.4 g cm–3 and the thickness

of the film is ~ 4 μm. The interdigitated electrode was obtained by scraping the film along the

transparent path in Figure S14b using a knife. Two copper tapes were stuck on each side of the

electrodes to work as current collectors. The planar supercapacitor was finally coated with

Page 26: DOI: 10.1002/ ((please add manuscript number)) Full Paper

26

PVA/H3PO4 gel electrolyte on the parallel region to fill the channel between the two working

electrodes. The assembled device was left overnight to allow the permeation of the electrolyte

into the electrodes and the dry of the gel. The PVA/H3PO4 electrolyte was prepared as follow:

2 g PVA was added into 20 mL water and the mixture was heated to 85 °C under stirring for

0.5 hour. Afterwards, 2 g H3PO4 was added and the solution was continuously stirred at 85 °C

until becoming clear.

The galvanostatic charge–discharge curves were recorded on CHI 760D using two–electrode

system at an operating potential range of 0.8 V at different current densities ranging from 2.5–

50 A g–1.

The specific capacitance (C) of the device was estimated from the discharge curve based on

the following equation:

(1)

where I is the constant current applied to the planar supercapacitor; ∆t and ∆V represent the

discharging time and potential range (0.8 V) respectively and m is the total mass of two

electrodes. Energy density (E) and power density (P) were derived based on the following

equations:

(2)

(3)

1.5 Photocatalytic hydrogen evolution reaction measurement

The experiments were performed in a 100 mL Pyrex flask at atmospheric pressure and room

temperature (22 ± 2 °C). A 300 W Xe lamp through a UV–cutoff filter (>420 nm) was used as

the light source (15 cm away from the reactor). The measurement protocol follows the literature

report.[3] In a typical experiment, 80 mg of photocatalyst powder was suspended in 60 mL H2O

and 20 mL methanol was added as the scavenger. 1.5 wt % Pt cocatalyst was loaded by directly

dissolving H2PtCl6·xH2O into the above 80 mL mixed solution. The suspension was then

illuminated by the lamp for 20 minutes under stirring. After that, the suspension was firstly

sonicated for 30 minutes and then bubbled with nitrogen for 30 minutes to remove the dissolved

oxygen. During the photocatalytic experiments, circulating water was used to maintain the

temperature of the flask. 0.2 mL of gas was sampled every one hour through the rubber septum

of the flask, and analyzed by GC.

Page 27: DOI: 10.1002/ ((please add manuscript number)) Full Paper

27

2. Supplementary Figures

Figure S1. Schematic illustration of the two–compartment electrochemical cell to investigate

of the parameter for electrochemical exfoliation at either anode or cathode.

Figure S2. Photograph (a), SEM (b) and XRD pattern (c) of the graphite rod. The interplanar

spacing between each graphene layer in the graphite can be derived from the (002) diffraction

peak according to Bragg’s law.

Page 28: DOI: 10.1002/ ((please add manuscript number)) Full Paper

28

Figure S3. Crystallographic diameters of ions used for the intercalation in this research. The

data were obtained from reference [4].

Page 29: DOI: 10.1002/ ((please add manuscript number)) Full Paper

29

Figure S4. XRD patterns (a), Raman spectrum (b), perpendicular SEM image (c) and cross–

sectional SEM image (d) of the natural MoS2 single crystal. (e) Low magnification SEM image

of the expanded MoS2 crystal. The upper part was not contacted with the electrolyte, which

showed no expansion, compared to the lower part with obvious expansion. (f) Some fully

exfoliated flakes in the expanded MoS2 crystal, which indicates the exfoliated flakes can be as

large as several ten micrometers.

Figure S5. SEM (a) and XRD patterns (b) of the MoS2 powder. The interplanar spacing

between each layer (0.618 nm) can be derived from the (002) diffraction peak according to

Bragg’s law.

Page 30: DOI: 10.1002/ ((please add manuscript number)) Full Paper

30

Figure S6. (a) Schematic illustration of the electrode used for cathodic exfoliation of MoS2,

MnO2 and g-C3N4. Resistance of the electrode of MoS2 (b), MnO2 (c) and g-C3N4 (d). The

resistance was small using graphite or Pb powder as conducting additive.

Figure S7. (a) HR–TEM image of as–exfoliated MoS2. (b) Statistics of the lateral size from

100 randomly selected nanosheets in TEM images. (c) Height profile of the line in the inserted

AFM image (the same image as Figure 3d). (d) Statistics of the height from100 nanosheets in

AFM images.

Page 31: DOI: 10.1002/ ((please add manuscript number)) Full Paper

31

Figure S8. (a) iR compensated polarization curves of MoS2–G composites with different

graphene weight contents (obtained under different applied potentials). (b) iR compensated

polarization curves of MoS2–G and MoS2 nanosheets–graphene mixture. Electrolyte: 0.5 M

H2SO4; scan rate is 2 mV s–1.

Figure S9. Raman spectra of exfoliated MoS2 NS, in situ formed MoS2–G and the physical

mixture of exfoliated MoS2 and graphene (MoS2–G mix). A redshift of the E12g and A1g bands

was observed only in the MoS2–G sample, whereas there is no redshift in the MoS2–G mix

sample. The redshift was assigned to be the n–dope of carbon materials to the MoS2 nanosheets,

as previously reported,[5] which indicates the electronic interaction between them.

Figure S10. Chronoamperometric response of the MoS2–G electrode at an applied potential of

–0.125 V (vs. RHE). After 24–hour electrolysis, only ~2% activity loss was observed.

Page 32: DOI: 10.1002/ ((please add manuscript number)) Full Paper

32

Figure S11. SEM (a) and XRD patterns (b) of MnO2 powder. 2θ = 12.74° at (001) diffraction

peak, corresponding to an interplanar spacing of 0.694 nm according to Bragg’s law.

Figure S12. Statistics of lateral size (a) of 100 pieces of MnO2 nanosheets in the TEM images

and height (b) of 50 pieces of MnO2 nanosheets in the AFM images.

Figure S13. SEM and EDS mapping results of the MnO2–G film obtained by exfoliation at 8

V. MnO2 distributes uniformly in the film. Graphene was 31.7 wt%.

Page 33: DOI: 10.1002/ ((please add manuscript number)) Full Paper

33

Figure S14. Schematic illustration (a) and photograph (b) of the planar supercapacitor.

Figure S15. (a) Specific capacitance of the device at different current densities. (b) Ragone plot

of the supercapacitor. (c) Cycle performance of the device with a potential of 0.8 V at a current

density of 25 A g–1.

Figure S16. TEM (a) and XRD patterns (b) of the bulk g-C3N4 powder. The interplanar spacing

between each layer (0.322 nm) can be derived from the (002) diffraction peak according to

Bragg’s law.

Page 34: DOI: 10.1002/ ((please add manuscript number)) Full Paper

34

Figure S17. Statistics of the height of 50 pieces of g-C3N4 nanosheets in the AFM images.

About 60% of nanosheets is in the range of 4–6 nm in height.

Page 35: DOI: 10.1002/ ((please add manuscript number)) Full Paper

35

3. Supplementary Tables

Table S1. Representative parameters in the electrochemical exfoliation method.

Target

material

Starting

material

Exfoliation

electrode Electrolyte Voltage[a] Ref.

graphene natural graphite

flakes anode 0.1 M H2SO4 in H2O +10 V [6]

graphene

HOPG[b] or

natural graphite

flakes

anode 0.5 M H2SO4 in H2O +1 V (30 min),

+10 V (1 min) [7]

graphene graphite rod anode 0.1 M Py1,4–TFSI in

acetonitrile +7 V [8]

graphene graphite rod anode 1 mM poly(sodium–4–

styrenesulfonate) +5 V [9]

carbon

nanoribbons

and

nanoparticles,

graphene

HOPG anode BMIM–BF4/H2O mixture range from

+1.5V – +8 V [10]

graphene natural graphite

flakes anode

0.1 M (NH4)2SO4 or

K2SO4 or Na2SO4 in H2O +10 V [11]

graphene graphite foil anode 0.05 M (NH4)2SO4 + 1

mg mL–1 TEMPO[c] +10 V [12]

graphene graphite foil anode/cathode[d] 0.1 M TBA–HSO4 in

H2O

AC (±10 V, 0.1

Hz) [13]

graphene graphite rod cathode 0.1 M TMA–ClO4 in

NMP

–2.4 V vs.

Ag/AgClO4 [14]

graphene HOPG cathode 30 mg mL–1 LiClO4 in

proprylene carbonate –15 V±5 V [15]

graphene graphite foil cathode

0.1 M LiClO4 + 0.03 M

TBA–ClO4 in proprylene

carbonate

–5 V [16]

BN, NbSe2,

WSe2, Sb2Se3,

Bi2Te3

corresponding

crystals cathode

1m LiPF6 in ethyl

carbonate/dimethyl

carbonate (1:1 v/v)

constant

current of

0.025 mA

[17]

Bi2Se3/Bi2Te3 synthesized

crystals cathode 0.5 M Na2SO4 in H2O

altering +10 V

and –10V [18]

MoS2 natural single

crystal anode 0.5 M Na2SO4 in H2O

+2 V, 10 min,

then +10 V [19]

BP[e] BP crystal anode 0.5 M Na2SO4 in H2O +7 V [20]

graphene–

other 2D

nanosheet

composite

graphite rod,

MoS2, MnO2 or

g-C3N4 powder

both anode

and cathode

0.1 M Py1,4–BF4 or

TBA–PF6 in acetonitrile

range from +7

V to 20 V

this

work

[a]We use + to represent the voltage used in anodic exfoliation and – for the voltage used in cathodic exfoliation. [b]highly oriented pyrolytic graphite. [c](2,2,6,6–tetramethylpiperidin–1–yl)oxyl. [d]AC voltage was applied on the

electrodes. Essentially only one material – graphene – was produced. [e]black phosphorus.

Page 36: DOI: 10.1002/ ((please add manuscript number)) Full Paper

36

Table S2. Comparison of the HER performance of the representative state–of–the–art

electrocatalyst based on MoS2–carbon materials in acidic media.

Catalyst Synthesis method η10

[a]

(mV)

Tafel slope

(mV dec–1) Ref.

O–incorporated MoS2

vertically on rGO

Hummers’ method for GO; hydrothermal method

for composite ~180 40 [21]

MoSx on graphene

coated Ni foam CVD for graphene, thermolysis for MoSx n.a.[b] 43 [22]

MoS2/N–doped rGO

composite

Hummers’ method for GO; hydrothermal method

for composite 107 41.3 [23]

S–rich MoS2 in N–

doped carbon nanofibers

electrospinning polymer fiber containing

precursors, treated with S vapor at 400 °C,

graphited at 1000 °C

120 38 [24]

MoS2 nanoparticles on

rGO

Hummers’ method for GO; hydrothermal method

for composite ~150 41 [25]

MoS2 quantum dots

decorated rGO

Hummers’ method for GO; programmed

sonication for MoS2 quantum dots 64 63 [26]

MoSx/N–doped CNT[c]

forest hybrid

plasma–enhanced CVD for CNT, decomposition

of precursor at 90 °C for decoration of MoSx on

CNT

~110 40 [27]

vertically aligned MoS2

on rGO

Hummers’ method for GO; microwave method

for composite 104 63 [28]

MoS2 on mesoporous

rGO

Hummers’ method for GO; hydrothermal method

for composite ~100 42 [29]

MoS2 on

rGO/CNT/polyimide

film

Hummers’ method for GO; electrodeposition

method for composite ~200 61 [30]

Mo2C on rGO/CNT Hummers’ method for GO; thermolysis method

for composite ~130 58 [31]

MoS2 nanosheets/CNT

hybrid

hydrothermal method for MoS2; CVD for CNT;

simple dispersion in solvent for the hybrid ~120 40.8 [32]

MoS2 nanoflower

decorated rGO paper

Hummers’ method for GO; filtration for GO

paper; hydrothermal method for the composite ~300 95 [33]

3D N–doped rGO

supported MoS2

nanoparticles

Hummers’ method for GO; hydrothermal method

for composite with polypyrrole as nitrogen source ~290 44 [34]

Ni–doped MoS2

nanosheets on carbon

fiber cloth

hydrothermal method 200 85.3 [35]

In situ formed MoS2–

graphene composite

electrochemical exfoliation in a single

electrochemical cell using graphite and MoS2

powder as precursors

125 41 this

work

[a]overpotential at the current density of 10 mA cm–2; [b]not available; [c]carbon nanotube.

Page 37: DOI: 10.1002/ ((please add manuscript number)) Full Paper

37

Table S3. Comparison of the performance of representative state–of–the–art MnO2–carbon

material based supercapacitors.

Electrode Synthesis method Specific device

capacitance (F g–1)

Energy density

(Wh kg–1)

Power density

(kW kg–1) Ref.

3D

graphene/M

nO2

networks

CVD method to grow

graphene on Ni foam;

electrodeposition method

for MnO2

116[a] (2 mV s–1) 6.8[b] 0.06 36

carbon

nanoparticle

s/MnO2

nanorod

hybrid

flame synthesis method for

carbon nanoparticles,

electrodeposition for

MnO2

200[a] (5 mV s–1) 4.8[b] 14 37

graphene/M

nO2

not available for graphene

synthesis; microwave

method for the composite

78[a] (2 mV s–1) 51.1[c] 0.1

38 8.2[c] 16.5

GO-MnO2

nanospheres

Hummers’ method for GO;

chemical co-precipitate

method for MnO2

307[d] (0.2 A g–1) 42.6[d] 0.276 39

graphene/M

nO2

Hummers’ method for GO;

microwave method for

MnO2

78[a] (2 mV s–1) n.a. n.a. 40

MnO2

nanowire/rG

O

Hummers’ method for GO;

ion intercalation for MnO2 31[e] (0.5 A g–1) 7.0[c] 5 41

MnO2/GO/C

NT

Hummers’ method for GO;

wet chemical method for

MnO2; CVD for CNT

93[a] (10 mV s–1) 8.9 0.106

42 3.2 1.28

planar 2D

graphene/M

nO2

Hummers’ method for GO;

liquid phase exfoliation for

MnO2

267[e] (0.2 A g–1) 17[b] 2.52 43

planar 2D

MnO2–

graphene

In situ formed composite

via electrochemical

exfoliation

265[e] (2.5 A g–1) 20.7[b] 2.57 this

work

[a]data estimated from reported single electrode on the basis of ; [b]data from symmetric

devices; [c]data from asymmetric devices, with carbon material as counter electrode; [d]data from asymmetric

devices, with MoO3–carbon as counter electrode; [e]data based on the active materials of the whole device.

Page 38: DOI: 10.1002/ ((please add manuscript number)) Full Paper

38

4. Supplementary Note

As demonstrated in Figure S18a, we were able to intercalate several anions, including BF4–,

PF6– and TFSI– (bis(trifluoromethylsulfonyl)imide), into graphite anode to exfoliate graphene

with a yield of nearly 50% under an applied potential of 10 V. The as–exfoliated graphene

demonstrates a typical thin–layer morphology with an average lateral size of ~3.5 μm and can

be dispersed well in water without the assistance of any surfactants (Figure S19). Cations have

negligible effect on the exfoliation yield, as is apparent when Py1,4+ was replaced by the

somewhat larger TBA+; this is reasonable since the anion is the species intercalating into the

graphite anode. The irrelevance of cation species to anodic exfoliation offers the advantage that

a great number of salts can be introduced as supporting electrolyte. The crystallographic

diameters of anions are larger than the interplanar spacing of graphite (0.334 nm). The huge

internal stress induced by the intercalation of the anions leads to the expansion and further

exfoliation of graphite anode. Applied potential can influence the exfoliation efficiency and the

quality of produced graphene (Figure S18b). Increasing the applied potential leads to higher

exfoliation rate, but introduces more defects, as revealed by an increased intensity ratio between

the D and G bands in the Raman spectrum.44 Nevertheless, in the potential range of 7–14 V, the

ID/IG is still less than 1.0, which is smaller than rGO obtained via chemical reduction of GO.44

The lower ratio is related to fewer defects and a higher degree of preservation of the conjugated

structure of graphene, which is believed to be beneficial to the migration of charge carriers.

Figure S18. Anodic exfoliation of graphite. (a) The exfoliation yields when electrolyte with

different ion diameter was used. (b) The yield and quality (revealed by the intensity ratio

between the D and G bands in the Raman spectrum, where a lower ratio corresponds to fewer

defects) of graphene when different potentials were applied.

Page 39: DOI: 10.1002/ ((please add manuscript number)) Full Paper

39

Figure S19. (a) Photograph of the graphene dispersion (0.1 mg mL–1) exfoliated at the anode,

displaying the Tyndall effect. (b) TEM image of the graphene nanosheet. (c) Statistics of the

lateral size of the graphene by 100 nanosheets from randomly navigated view field in the TEM

measurement. (d) AFM image of the graphene.

Page 40: DOI: 10.1002/ ((please add manuscript number)) Full Paper

40

5. Supplementary References

[1] S. Komaba, N. Kumagai and S. Chiba, Electrochim. Acta 2000, 46, 31-37.

[2] X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and

M. Antonietti, Nat. Mater. 2009, 8, 76-80.

[3] Q. Xiang, J. Yu and M. Jaroniec, J. Phys. Chem. C 2011, 115, 7355-7363.

[4] H. Tokuda, K. Hayamizu, K. Ishii, M. A. B. H. Susan and M. Watanabe, J. Phys. Chem.

B 2004, 108, 16593-16600.

[5] Z. Li, R. Ye, R. Feng, Y. Kang, X. Zhu, J. M. Tour and Z. Fang, Adv. Mater. 2015, 27,

5235-5240.

[6] K. Parvez, R. Li, S. R. Puniredd, Y. Hernandez, F. Hinkel, S. Wang, X. Feng and K.

Mullen, ACS Nano 2013, 7, 3598-3606.

[7] C. Y. Su, A. Y. Lu, Y. Xu, F. R. Chen, A. N. Khlobystov and L. J. Li, ACS Nano 2011,

5, 2332-2339.

[8] A. T. Najafabadi and E. Gyenge, Carbon, 2014, 71, 58-69.

[9] G. Wang, B. Wang, J. Park, Y. Wang, B. Sun and J. Yao, Carbon 2009, 47, 3242-3246.

[10] J. Lu, J. X. Yang, J. Wang, A. Lim, S. Wang and K. P. Loh, ACS Nano 2009, 3, 2367-

2375.

[11] K. Parvez, Z. S. Wu, R. Li, X. Liu, R. Graf, X. Feng and K. Mullen, J. Am. Chem. Soc.

2014, 136, 6083-6091.

[12] S. Yang, S. Bruller, Z. S. Wu, Z. Liu, K. Parvez, R. Dong, F. Richard, P. Samori, X.

Feng and K. Mullen, J. Am. Chem. Soc. 2015, 137, 13927-13932.

[13] S. Yang, A. G. Ricciardulli, S. Liu, R. Dong, M. R. Lohe, A. Becker, M. A. Squillaci,

P. Samori, K. Mullen and X. Feng, Angew. Chem. Int. Ed. 2017, 56, 6669-6675.

[14] A. J. Cooper, N. R. Wilson, I. A. Kinloch and R. A. W. Dryfe, Carbon 2014, 66, 340-

350.

[15] J. Wang, K. K. Manga, Q. Bao and K. P. Loh, J. Am. Chem. Soc. 2011, 133, 8888-8891.

[16] Y. L. Zhong and T. M. Swager, J. Am. Chem. Soc. 2012, 134, 17896-17899.

[17] Z. Zeng, T. Sun, J. Zhu, X. Huang, Z. Yin, G. Lu, Z. Fan, Q. Yan, H. H. Hng and H.

Zhang, Angew. Chem. Int. Ed. 2012, 51, 9052-9056.

[18] A. Ambrosi, Z. Sofer, J. Luxa and M. Pumera, ACS Nano 2016, 10, 11442-11448.

[19] N. Liu, P. Kim, J. H. Kim, J. H. Ye, S. Kim and C. J. Lee, ACS Nano 2014, 8, 6902-

6910.

[20] M. B. Erande, S. R. Suryawanshi, M. A. More and D. J. Late, Eur. J. Inorg. Chem. 2015,

2015, 3102-3107.

[21] A. Liu, L. Zhao, J. Zhang, L. Lin and H. Wu, ACS Appl. Mater. Interf. 2016, 8, 25210-

25218.

[22] Y. H. Chang, C. T. Lin, T. Y. Chen, C. L. Hsu, Y. H. Lee, W. Zhang, K. H. Wei and L.

J. Li, Adv. Mater. 2013, 25, 756-760.

[23] Y.-J. Tang, Y. Wang, X.-L. Wang, S.-L. Li, W. Huang, L.-Z. Dong, C.-H. Liu, Y.-F. Li

and Y.-Q. Lan, Adv. Energy Mater. 2016, 6, 1600116.

[24] H. Zhu, M. Du, M. Zhang, M. Zou, T. Yang, S. Wang, J. Yao and B. Guo, Chem.

Commun. 2014, 50, 15435-15438.

[25] Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong and H. Dai, J. Am. Chem. Soc. 2011, 133,

7296-7299.

[26] F. Li, J. Li, Z. Cao, X. Lin, X. Li, Y. Fang, X. An, Y. Fu, J. Jin and R. Li, J. Mater.

Chem. A 2015, 3, 21772-21778.

[27] D. J. Li, U. N. Maiti, J. Lim, D. S. Choi, W. J. Lee, Y. Oh, G. Y. Lee and S. O. Kim,

Nano Lett. 2014, 14, 1228-1233.

[28] M. Chatti, T. Gengenbach, R. King, L. Spiccia and A. N. Simonov, Chem. Mater. 2017,

29, 3092-3099.

Page 41: DOI: 10.1002/ ((please add manuscript number)) Full Paper

41

[29] L. Liao, J. Zhu, X. Bian, L. Zhu, M. D. Scanlon, H. H. Girault and B. Liu, Adv. Funct.

Mater. 2013, 23, 5326-5333.

[30] Y. Jiang, X. Li, S. Yu, L. Jia, X. Zhao and C. Wang, Adv. Funct. Mater. 2015, 25, 2693-

2700.

[31] D. H. Youn, S. Han, J. Y. Kim, J. Y. Kim, H. Park, S. H. Choi and J. S. Lee, ACS Nano

2014, 8, 5164-5173.

[32] Y. Cai, X. Yang, T. Liang, L. Dai, L. Ma, G. Huang, W. Chen, H. Chen, H. Su and M.

Xu, Nanotechnology 2014, 25, 465401.

[33] C. B. Ma, X. Qi, B. Chen, S. Bao, Z. Yin, X. J. Wu, Z. Luo, J. Wei, H. L. Zhang and H.

Zhang, Nanoscale 2014, 6, 5624-5629.

[34] H. Dong, C. Liu, H. Ye, L. Hu, B. Fugetsu, W. Dai, Y. Cao, X. Qi, H. Lu and X. Zhang,

Sci. Rep. 2015, 5, 17542.

[35] J. Miao, F. X. Xiao, H. B. Yang, S. Y. Khoo, J. Chen, Z. Fan, Y. Y. Hsu, H. M. Chen,

H. Zhang and B. Liu, Sci. Adv. 2015, 1, e1500259.

[36] Y. He, W. Chen, X. Li, Z. Zhang, J. Fu, C. Zhao and E. Xie, ACS Nano 2013, 7, 174-

182.

[37] L. Yuan, X. H. Lu, X. Xiao, T. Zhai, J. Dai, F. Zhang, B. Hu, X. Wang, L. Gong, J.

Chen, C. Hu, Y. Tong, J. Zhou and Z. L. Wang, ACS Nano 2012, 6, 656-661.

[38] Z. Fan, J. Yan, T. Wei, L. Zhi, G. Ning, T. Li and F. Wei, Adv. Funct. Mater. 2011, 21,

2366-2375.

[39] J. Chang, M. Jin, F. Yao, T. H. Kim, V. T. Le, H. Yue, F. Gunes, B. Li, A. Ghosh, S.

Xie and Y. H. Lee, Adv. Funct. Mater. 2013, 23, 5074-5083.

[40] J. Yan, Z. Fan, T. Wei, W. Qian, M. Zhang and F. Wei, Carbon 2010, 48, 3825-3833.

[41] Z. S. Wu, W. Ren, D. W. Wang, F. Li, B. Liu and H. M. Cheng, ACS Nano 2010, 4,

5835-5842.

[42] Y. Cheng, S. Lu, H. Zhang, C. V. Varanasi and J. Liu, Nano Lett. 2012, 12, 4206-4211.

[43] L. Peng, X. Peng, B. Liu, C. Wu, Y. Xie and G. Yu, Nano Lett. 2013, 13, 2151-2157.

[44] S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y.

Wu, S. T. Nguyen and R. S. Ruoff, Carbon 2007, 45, 1558-1565.