Journal Name

1 downloads 0 Views 876KB Size Report
proposed for a plethora of next generation uses with potential applications in ..... This is a small but clear reduction of the onset potential which trends with films ...
Dynamic Article Links ►

Journal Name Cite this: DOI: 10.1039/c0xx00000x

ARTICLE TYPE

www.rsc.org/xxxxxx

Molybdenum disulfide/pyrolytic carbon hybrid electrodes for scalable hydrogen evolution Hugo Nolan,a,b Niall McEvoy,*a Maria O’Brien,a,b Nina C. Berner,a Chanyoung Yim,a,b Toby Hallam,a Aidan R. McDonaldb and Georg S. Duesberg*a,b 5

10

15

Received (in XXX, XXX) Xth XXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX DOI: 10.1039/b000000x

The electrochemical generation of hydrogen fuel via the proton reduction in the Hydrogen Evolution Reaction (HER) in aqueous media is currently dependent on the use expensive noble metal catalysts for which alternatives must be sought. Molybdenum disulfide (MoS2) has shown great promise as a suitable electrocatalyst in this regard. While many lab-scale experiments on the HER activity of this material have demonstrated its viability and explored some fundamental mechanistic features of HER at MoS2, these experimental techniques are often ill-suited to large scale production of such electrodes. In this study we present work on the fabrication of MoS2/pyrolytic carbon (PyC) electrodes via vapour phase sulfurization of Mo thin films. These hybrid electrodes combine the catalytic activity of MoS2 with the conductivity and stability of PyC, whilst using industrially compatible processing techniques. Structural defects in the sulfur lattice were found to be key catalytically active sites for HER and thinner MoS 2 films displayed a higher quantity of these defects and, hence, an improved HER activity. The observed Tafel slope of 95 mV/dec is comparable to previous literature works on MoS2 HER performance.

Introduction 20

25

30

35

40

45

Efficient and sustainable production of hydrogen is essential for a hydrogen based economy to reduce our dependence on fossil fuel energy sources.1 The electrochemical splitting of water via proton reduction in the Hydrogen Evolution Reaction (HER) provides a feasible route towards this without dependence on consumable resources. However, to date, the best HER performance has been achieved using prohibitively expensive platinum electrodes, severely limiting the practical applications of this technology. If the HER is to be exploited as a viable energy source, alternative electrocatalysts to Pt must be sought. Recent years have seen ever increasing focus placed on the synthesis and characterisation of layered 2D materials, based principally on their novel properties.2 Much work in the literature at present is concerned with layered Transition Metal Dichalcogenides (TMDs).2,3 This group of materials has been proposed for a plethora of next generation uses with potential 7–9 applications in 2D electronics,4–6 gas sensing and, most 10–12 significantly for this study, catalysis of the HER, having been demonstrated in recent years. Thus far, molybdenum disulfide (MoS2) has been the most heavily studied material in the context of catalysis, although studies on the HER behaviour of other materials such as MoSe2, 13–15 WS2 and WSe2 have recently emerged. Depending on the synthesis route undertaken, HER activity has been attributed to different features of TMD layers. Where MoS 2 flakes have been exfoliated from bulk MoS2 via intercalation of Li-ions, it has been demonstrated that the HER activity is related to the presence 16,17 of the so-called 1T polymorph. This is not found in bulk This journal is © The Royal Society of Chemistry [year]

50

55

60

65

70

75

MoS2, which consists solely of the 2H polymorph. Many other studies have attributed HER activity of MoS2 to the presence of unsaturated sulfur atoms along molybdenum edges of the MoS2 18,19 structure. Furthermore, theoretical studies have shown that sulfur vacancies catalyse various reactions at the MoS2 surface.20 These defects, in the otherwise catalytically inactive basal plane of the sulfur sub-lattice, act as catalytically active sites. The majority of studies on HER by TMDs deal with dispersions of TMD flakes or nanoparticles which are then immobilised on an electrode surface for HER measurements. Often these same studies incorporate further processing conditions or stages to enhance the HER activity of these films by establishing active sites on the material. Li et al. used a solvothermal process to synthesise MoS2 flakes and found that incorporating Reduced Graphene Oxide (rGO) to create an MoS2/rGO hybrid improved 10 performance for HER. Xie et al. reported a synthetic chemical route where the reaction precursors were controlled to create defective MoS2 nanoflakes which exhibit favourable HER 21 activity. Wang et al. isolated MoS2 flakes exfoliated from 22 bulk and subsequently attributed increased HER performance of smaller nanoparticles to an increase in available unsaturated sulfur sites.23 Yu et al. found that the HER activity of MoS2 grown via Chemical Vapour Deposition (CVD) from an MoCl5 precursor improved with a reduction in the number of MoS 2 layers.24 They correlated the activity to the hopping of electrons between these layers. Typically characterisation of the HER activity of such materials is carried out by depositing very small amounts of the material on conductive substrates such as glassy carbon. Merki et al. developed a technique whereby MoS2 flakes can be electrochemically deposited on conductive substrates via [journal], [year], [vol], 00–00 | 1

25

5

10

15

20

25

repeated cyclic voltammetry in an electrolyte of (NH4)2[MoS4]. Such direct deposition/synthesis techniques could conceivably allow the large-scale fabrication of MoS2 electrodes for energy applications. Nonetheless, many techniques employed in literature studies on HER at MoS2, while useful for lab-scale experiments, offer poor scalability. Ultimately, if the implementation of TMDs as HER catalysts is to be successful, large scale production of such electrocatalysts for industrial applications must be viable. Here, we propose an electrode fabrication process that involves the deposition of thin films of metallic Mo directly on conductive substrates followed by vapour phase sulfurisation. The poor conductivity of MoS2 precludes the use of electrodes fabricated entirely from the material. Mo films were evaporated on to Pyrolytic Carbon (PyC) films grown via CVD. Our previous experience with PyC has shown that this material is highly suitable as an electrode material given its chemical, thermal and mechanical durability with many reported electrochemical 26–29 applications. This investigation aims to show that such thin layers of MoS 2, grown via vapour phase sulfurisation directly on a substrate of conductive PyC, present a suitable electrode for HER with few limits to up-scaling production in an industrially compatible process. In addition, details of the reaction mechanism are discussed and correlated to spectroscopic and microscopic characterisation of the MoS2/PyC films.

Results and Discussion

30

35

40

45

50

Electrodes were fabricated on SiO2/Si substrates with an initial step of CVD growth of PyC. Subsequent deposition of Mo was performed in a defined region of the sample to control the area of the resultant electrode. Finally, vapour phase sulfurisation was carried out to form MoS2 on the PyC coated substrate. All further material characterisation and electrochemical measurements were performed on such samples.

55

60

Material Characterisation Raman Spectroscopy MoS2 samples 4.5, 10 and 53 nm thick were investigated and spectral features of both PyC and MoS2 are evident. The broad D and G bands (~1353 and ~1583 cm-1, respectively) coupled with the highly suppressed 2D band at 2700 cm-1 are indicative of a disordered graphitic system such as PyC. The peaks observed at ~383 and ~408 cm-1 are characteristic of MoS2 and are denoted the E12g and A1g peaks and arise due to in plane and out of plane vibrations, respectively.30,31 Fig.1 (a) shows a comparison of the Raman spectra of each sample, normalised to the intensity of the A1g peak, with the D and G band region of PyC highlighted in grey and the MoS2 peaks highlighted in cyan. It is immediately apparent that the relative PyC characteristic signal is much reduced as the MoS2 layer thickness increases. The greater quantity of MoS2 present atop of the PyC surface causes a decreased penetration of the Raman laser into the PyC substrate.

2 | Journal Name, [year], [vol], 00–00

65

70

75

Fig 1: (a) Raman spectra for MoS2 films on PyC substrates. Also included is the spectrum of the bare substrate. The cyan region marks MoS2 characteristic peaks while the grey region marks PyC characteristic peaks. (b) High spectral resolution scan of the MoS2 peaks with the reduced peak separation of the thinner films indicated.

Analysis of the MoS2 characteristic peaks divulges information regarding the nature of the MoS2 films. Fig.1 (b) shows that the spectral region of interest for MoS2 is rather similar for all samples. However, closer inspection shows a shift in peak position which is dependent on MoS2 thickness. The coupling between electronic transitions and phonons is modified as the MoS2 thickness is decreased, causing a narrowing of peak separation between the E12g and A1g modes. This behaviour is reflected here with the dotted lines highlighting a reduction in peak separation of ~2 cm-1 on going from 53 nm to 4.5 nm thick MoS2. From this it is apparent that the 4.5 nm film can be considered distinct from bulk MoS2. X-Ray Photoelectron Spectroscopy XPS analysis of the MoS2/PyC films revealed a wealth of chemical information. Survey spectra of each MoS2 film and the bare PyC substrate are shown in Fig.2 (a) with characteristic photoemission peaks corresponding to C, S and Mo highlighted in grey, yellow and cyan, respectively. It is apparent, as with the Raman spectroscopy, that the spectral contributions from the PyC substrate diminish as the MoS2 layer thickness increases.

This journal is © The Royal Society of Chemistry [year]

35

40

45

5

10

15

20

25

30

the higher magnification image in Fig.3 (b) where randomly oriented crystallites can be discerned by virtue of the crystalline planes. Further analysis of MoS2/PyC cross sections is included in the ESI where an MoS2 interlayer spacing of approximately 6.6 Å was found, consistent with previously reported values.33 Further TEM analysis was carried out on MoS2 grown on SiO2 using an identical technique as used for that grown on PyC. This substrate allowed the MoS2 film to be transferred to TEM grids via a polymer supported transfer method. A top-down image of such a MoS2 film is shown in Fig.3 (c). Hexagonal symmetry is clearly visible in the lattice. Inset in this panel is a Fast Fourier Transform (FFT) corresponding to this image. The clear hexagonal pattern further confirms the structure of the film with the presence of multiple MoS 2 layers evidenced by the outer ring in the FFT spot pattern.

Fig 2: (a) XPS survey scans of the MoS2/PyC electrodes investigated in this study. Carbon characteristic core level peaks marked by grey, Mo peaks by cyan and the S peaks by yellow. (b) S2p core level peak for 4.5 nm MoS2 film with spectral contributions fitted.

To gain a greater insight into the chemical nature of the film, higher resolution scans of the S2p and Mo3d spectral regions were recorded and spectral contributions were fitted to the measured data. Fig.2 (b) shows the S2p region for the 4.5 nm film including peak contributions, which are consistent with the binding energy values found by Eda et al.32 The two key features are the doublet peak which corresponds to sulfur in bulk MoS 2 (marked in blue) and a smaller doublet peak at a slightly higher binding energy which is associated with unsaturated sulfur in edge crystallographic states (marked in green). It was found that this “edge-sulfur” contribution was appreciably smaller for thicker MoS2 films. The concentration of edge sulfur as determined by XPS was found to increase from approximately 7.6% in the thickest film to approximately 13.6% in the thinnest film. A detailed discussion of the S2p fitting and a comparison of all three MoS2 film thicknesses is included in the ESI in Fig.S1. From analysis of the Mo3d spectral region it was found that low levels of oxides in the form of MoO2 and MoO3 are present in the MoS2. However, these quantities remain invariant with MoS2 film thickness and, as such, do not influence any trends in HER activity of the MoS2/PyC electrodes (see ESI for detailed discussion, Fig.S2). Electron Microscopy Fig.3 (a) displays a cross-sectional TEM image of the 53 nm MoS2/PyC film. The indicated Au/Pd and Pt are part of a protective capping layer involved in the fabrication of the TEM lamella. A uniform thickness is observed throughout the film as shown. The polycrystalline nature of the MoS2 layer is evident in This journal is © The Royal Society of Chemistry [year]

50

Fig 3: (a) Cross sectional TEM image showing uniformity of the 53 nm MoS2/PyC hybrid electrode. (b) HR-TEM of the PyC/MoS2 interface. Crystalline planes are visible in the MoS2 layer. (c) High resolution top-

Journal Name, [year], [vol], 00–00 | 3

down image of MoS2 film on TEM grid with evident hexagonal symmetry. Inset: Corresponding FFT image.

HER Catalysis

5

10

15

20

Linear Sweep Voltammetry (LSV) was performed at a scan rate of 5 mV s-1 for several MoS2/PyC samples. The voltammetry data is presented in Fig.4. An understanding of the mechanism by which HER can proceed is required to fully characterise the HER activity of different electrode materials. A Tafel plot displays voltammetric data on a log –linear plot. The linear portion of the curve on a Tafel plot is described by the Tafel equation.

Where η is the overpotential, b is the Tafel slope, j is the current density and a captures several thermodynamic constants. The kinetics of different catalytic reactions can be related to the slope. As a result, analysis of the Tafel slope then provides an insight into the HER mechanism at the electrode surface. The following 34 is covered in detail by Compton and Banks and also Conway 35 and Tilak . In acidic media, an initial step involves the reduction of a proton in solution followed by adsorption of intermediate Hads to the electrode surface as follows: (Volmer reaction) A Tafel slope of approximately 120 mV/dec is associated with this reaction step. This is followed by desorption of H2 by either of the following reactions: (Heyrovsky reaction)

25

Fig 4: (a) Linear sweep voltammograms for MoS2/PyC hybrid electrodes compared to the bare PyC substrate and a platinum electrode. (b) Corresponding Tafel plots with fitted slopes. 55

(Tafel reaction) 35

30

35

40

45

50

According to Conway and Tilak, the latter reaction has a Tafel slope of approximately 30 mV/dec. The former has a slope of approximately 120 mV/dec when surface coverage of Hads is relatively low. In cases of higher surface coverage, this reduces to approximately 40 mV/dec. This is caused by the different steps of the mechanism dominating at the different Hads surface coverages. It is immediately apparent from Fig.4 (a) that the MoS2 films have a dramatically increased HER activity compared to the bare PyC substrate. Given the inertness of the PyC, any contributions from it towards the performance of the MoS2 films can be considered negligible. The onset potential for the different materials was determined by measuring the overpotential at which a current density of 2 mA cm-2 was achieved. It was found that the bare PyC substrate only achieved this current density at an overpotential of approximately 870 mV. Pt is known to have excellent HER catalytic performance and, as expected, a nearzero overpotential was observed. Interestingly, the onset potential of the MoS2/PyC films was found to decrease with decreasing film thickness. For 53 nm MoS2 an onset potential of 540 mV was recorded, the 10 nm film displayed a potential of 520 mV while an onset of 503 mV was observed for the 4.5 nm sample. This is a small but clear reduction of the onset potential which trends with films thickness.

4 | Journal Name, [year], [vol], 00–00

60

65

70

75

80

Further information about the HER activity of the various samples was garnered by analysing the linear region of the Tafel plots shown in Fig.4 (b). As previously mentioned, some details about the reaction mechanism at the electrode surface can be extracted from this data. It is known that the HER at a Pt electrode proceeds via the socalled Volmer-Tafel mechanism with a Tafel slope of 30 mV/dec. This is reflected in our experimental data where the Pt electrode exhibited an almost ideal Tafel slope of 31 mV/dec. Of the MoS2 films investigated in this study, the thinnest (4.5 nm) displayed the lowest Tafel slope of 95 mV/dec. The thicker the MoS2 film, the greater the Tafel slope observed. Li et al. proposed that the HER at MoS2 electrodes proceeds via a Volmer-Heyrovsky 10 mechanism. A modification of the structural morphology of the MoS2 flakes used in their MoS2/rGO electrodes caused a reduction in Tafel slope which suggested an increase in site density for Hads, as is consistent with this mechanism. In addition, it was proposed that more effective electronic coupling between the rGO and MoS2 flakes contributed to the enhanced HER performance. Our investigations indicate a reduction in Tafel slope as the film thickness decreases. It is likely that decreasing the MoS2 film thickness exposes a greater density of active edge sites on the surface of the film which, in turn, provides a greater number of accessible sites for Hads. This is consistent with behaviour observed by Wang et al., where smaller MoS2 flakes displayed enhanced HER activity compared to larger ones which was attributed to an increase in unsaturated sulfur sites at the MoS2 flakes.23 The Tafel slope of 95 mV/dec for our thinnest film is well below that of 110 mV/dec reported for untreated This journal is © The Royal Society of Chemistry [year]

5

10

15

20

exfoliated MoS2 flakes16 and comparable to that observed by Xie 36 et al. for calcined synthetic MoS2 nanosheets and by Li et al. for 10 synthetic MoS2 particles. The MoS2/rGO hybrid material of Ma et al. exhibits an identical Tafel slope.37 It should be noted that a secondary phenomenon may contribute to the improved HER performance of the thinner MoS2 films in this study whereby the thinner films facilitate electron migration to the MoS2 surface. While the results presented here do not match the best performance of MoS2 materials for HER available in the literature, a greater understanding of the mechanism by which the reaction proceeds has been achieved. Sufficiently thin MoS 2 layers exhibit behaviour distinct from bulk and a greater number of defects in the sulfur lattice is seen. Importantly, our electrodes display both good reproducibility and stability as shown in Fig.S6 of the supplementary information. It is envisaged that further modification of the MoS2/PyC hybrid stack, by chemical or physical means, could result in additional enhancement of HER catalysis. In addition, the facile fabrication of MoS2 electrodes on conductive substrates renders this process highly applicable for large scale production.

Conclusions

25

30

35

40

45

In conclusion, we have developed a robust method of fabricating electrodes of MoS2 on conductive PyC substrates and demonstrated encouraging results when this is employed as a HER catalyst. This combination of materials leads to a mechanically durable electrode assembly which can easily be manufactured on a large scale, unlike many techniques which are used in the literature to investigate HER at TMDs. Spectroscopic techniques confirm the stoichiometry of the MoS2 film and also indicate that our thinnest films behave in a manner distinct from bulk MoS2. The improved performance of the thinner MoS2 films over thicker ones is correlated to an apparent increase in sulfur defects in the thinner sample. We propose that this increase in crystalline defects allows for an increased surface coverage of adsorbed H (Hads) on the electrode surface. Controlling the thickness of the starting Mo layer is sufficient to enhance the HER performance of the resultant electrode. The performance of the MoS2 catalyst is comparable to other published results for untreated MoS2 materials. Further engineering of the MoS2/PyC electrode presented here to increase surface roughness and defect levels could further enhance the HER performance. Additionally, our methods could readily be extended to synthesise electrodes from other members of the TMD family (WS2, MoSe2 etc.), potentially improving their catalytic performance.

60

65

All chemicals used throughout the synthesis and characterisation were purchased from Sigma-Aldrich and used as-received. PyC films were grown by CVD of acetylene at 950° C for 30 minutes to a thickness of 300-400 nm on 300 nm thermal SiO2 on Si substrates in a hot wall quartz tube furnace as previously 26 reported. Following CVD of PyC, the samples were cooled under Ar atmosphere to room temperature. Subsequently, electron beam evaporation was carried out in a Temescal FCThis journal is © The Royal Society of Chemistry [year]

2000 to deposit Mo to thicknesses of approximately 2 nm, 4 nm and 20 nm. Areas of the PyC substrates were physically masked so that only a 1 cm2 area of the PyC sample was exposed to the Mo flux. This ensured a reproducible electrode area from sample to sample. Sulfurisation was then performed as previously 9,33 reported. Briefly, the Mo/PyC samples were placed in a two zone hot wall quartz tube furnace and heated to 750° C under Ar flow (150 sccm). Sulfur powder was thermally sublimed upstream of the samples at a temperature of 120° C and sulfurisation of the Mo took place in the higher temperature region. Samples were then cooled to room temperature under Ar flow. Electrochemical Measurements

70

75

80

85

90

95

Electrochemical measurements were carried out to evaluate the performance of MoS2/PyC films as electrodes for HER. Sulfuric acid (0.5 M) was used as the electrolyte. A three electrode configuration was employed with MoS2 as the working electrode, a graphite counter electrode and a Reference Hydrogen Electrode (RHE) as the reference electrode. A Gamry Reference 600 potentiostat was used for all electrochemical measurements. Spectroscopy Raman spectroscopy was performed using a Witec Alpha 300 R confocal Raman microscope with a laser wavelength of 532 nm at a power of ~ 0.5 mW. For wide spectral scans a grating with 600 lines/mm was used. Higher spectral resolution measurements required the use of a 1800 lines/mm grating. XPS was carried out on a VG Scientific ESCAlab MKII system, using Al Kα X-rays (1486.7 eV) as incident light. An analyser pass energy of 50 eV was used for survey scans while a pass energy of 20 eV was used to obtain high resolution spectra of characteristic core levels. Thickness measurements of as-deposited Mo and sulfurised MoS2 films were performed by spectroscopic ellipsometry as described in a previous work.38 Bare SiO2 substrates were included in the same deposition runs as the PyC samples of each thickness and these samples were used to calibrate the actual film thicknesses for each sample set. Mo films of 2, 4 and 20 nm were found to increase in thickness by a factor of approximately 2.5 after sulfurisation. The resultant MoS2 films were found to be approximately 4.5, 10 and 53 nm thick, respectively. Detailed discussion of these measurements is included in the ESI. Microscopy

Experimental Electrode Fabrication

50

55

100

105

Cross sections of the MoS2 on PyC were fabricated by Focussed Ion Beam (FIB) milling of lamellar cross sections followed by TEM imaging. A PMMA polymer support layer was spun on before the SiO2 substrate was etched using 2M NaOH. The resultant films were transferred to a TEM grid and the PMMA layer dissolved in acetone. A Carl Zeiss Auriga FIB was used for lamella fabrication and TEM images were recorded using an FEI Titan TEM at an accelerating voltage of 300 kV. Scanning Electron Microscopy (SEM) was performed with a Carl Zeiss Ultra FE SEM at low accelerating voltages with a secondary electron detector. Atomic Force Microscopy (AFM) images were obtained using a Journal Name, [year], [vol], 00–00 | 5

Veeco Dimension 3100 atomic force microscope in tapping mode with 300 kHz resonant Si Tips.

65

Acknowledgements 5

10

This work is supported by the SFI under Contract No. 12/RC/2278 and PI_10/IN.1/I3030. NM acknowledges the EU under FP7-2010-PPP Green Cars (Electrograph No. 266391) and MO is supported by the Irish Research Council under the Enterprise Partnership Scheme, Project #201517, Award #12508. The authors thank Mr. Davide Betto for his assistance with electron beam evaporation. Mr. Dermot Daly and Mr. Clive Downing, of the Advanced Microscopy Laboratory, are thanked for their assistance with FIB and TEM techniques, respectively.

70

75

80

References 15

1. 2.

20

3. 4. 5.

25

6. 7. 30

8. 9. 10.

35

11. 12. 13. 40

14. 15. 45

16. 17. 50

18.

19. 55

20. 21. 22. 60

23. 24.

J. A. Turner, Science, 2004, 305, 972–974. S. Z. Butler, S. M. Hollen, L. Cao, Y. Cui, J. A. Gupta, H. R. Gutierrez, T. F. Heinz, S. S. Hong, J. Huang, A. F. Ismach, E. Johnston-Halperin, M. Kuno, V. V. Plashnitsa, R. D. Robinson, R. S. Ruoff, S. Salahuddin, J. Shan, L. Shi, M. G. Spencer, M. Terrones, W. Windl, and J. E. Goldberger, ACS Nano, 2013. M. Chhowalla, H. S. Shin, G. Eda, L.-J. Li, K. P. Loh, and H. Zhang, Nat. Chem., 2013, 5, 263–275. B. Radisavljevic, M. B. Whitwick, and A. Kis, ACS Nano, 2011, 5, 9934–9938. K. Lee, H. Kim, M. Lotya, J. N. Coleman, G. Kim, and G. S. Duesberg, Adv. Mater., 2011, 23, 4178–4182. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti, and A. Kis, Nat. Nanotechnol., 2011, 6, 147–150. Q. He, Z. Zeng, Z. Yin, H. Li, S. Wu, X. Huang, and H. Zhang, Small, 2012, 8, 2994–2999. H. Li, Z. Yin, Q. He, H. Li, X. Huang, G. Lu, D. W. H. Fam, A. I. Y. Tok, Q. Zhang, and H. Zhang, Small, 2012, 8, 63–67. K. Lee, R. Gatensby, N. McEvoy, T. Hallam, and G. S. Duesberg, Adv. Mater., 2013, 25, 6699–6702. Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong, and H. Dai, J. Am. Chem. Soc., 2011, 133, 7296–7299. D. Merki and X. Hu, Energy Environ. Sci., 2011, 4, 3878–3888. Y. Zhao, Y. Zhang, Z. Yang, Y. Yan, and K. Sun, Sci. Technol. Adv. Mater., 2013, 14, 043501. D. Kong, H. Wang, J. J. Cha, M. Pasta, K. J. Koski, J. Yao, and Y. Cui, Nano Lett., 2013, 13, 1341–1347. H. Wang, D. Kong, P. Johanes, J. J. Cha, G. Zheng, K. Yan, N. Liu, and Y. Cui, Nano Lett., 2013, 13, 3426–3433. D. Voiry, H. Yamaguchi, J. Li, R. Silva, D. C. B. Alves, T. Fujita, M. Chen, T. Asefa, V. B. Shenoy, G. Eda, and M. Chhowalla, Nat. Mater., 2013, 12, 850–855. M. A. Lukowski, A. S. Daniel, F. Meng, A. Forticaux, L. Li, and S. Jin, J. Am. Chem. Soc., 2013, 135, 10274–10277. D. Voiry, M. Salehi, R. Silva, T. Fujita, M. Chen, T. Asefa, V. B. Shenoy, G. Eda, and M. Chhowalla, Nano Lett., 2013, 13, 6222– 6227. B. Hinnemann, P. G. Moses, J. Bonde, K. P. Jørgensen, J. H. Nielsen, S. Horch, I. Chorkendorff, and J. K. Nørskov, J. Am. Chem. Soc., 2005, 127, 5308–5309. T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S. Horch, and I. Chorkendorff, Science, 2007, 317, 100–102. D. Le, T. B. Rawal, and T. S. Rahman, J. Phys. Chem. C, 2014. J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou, X. W. (David) Lou, and Y. Xie, Adv. Mater., 2013, 25, 5807–5813. T. Wang, L. Liu, Z. Zhu, P. Papakonstantinou, J. Hu, H. Liu, and M. Li, Energy Environ. Sci., 2013, 6, 625–633. T. Wang, D. Gao, J. Zhuo, Z. Zhu, P. Papakonstantinou, Y. Li, and M. Li, Chem. – Eur. J., 2013, 19, 11939–11948. Y. Yu, S.-Y. Huang, Y. Li, S. N. Steinmann, W. Yang, and L. Cao, Nano Lett., 2014, 14, 553–558.

6 | Journal Name, [year], [vol], 00–00

85

90

95

100

25. D. Merki, S. Fierro, H. Vrubel, and X. Hu, Chem. Sci., 2011, 2, 1262–1267. 26. N. McEvoy, N. Peltekis, S. Kumar, E. Rezvani, H. Nolan, G. P. Keeley, W. J. Blau, and G. S. Duesberg, Carbon, 2012, 50, 1216– 1226. 27. G. Keeley, N. McEvoy, S. Kumar, N. Peltekis, M. Mausser, and G. Duesberg, Electrochem. Commun., 2010, 12, 1034–1036. 28. G. P. Keeley, N. McEvoy, H. Nolan, S. Kumar, E. Rezvani, M. Holzinger, S. Cosnier, and G. S. Duesberg, Anal. Methods, 2012, 4, 2048–2053. 29. H. Nolan, N. McEvoy, G. P. Keeley, S. D. Callaghan, C. McGuinness, and G. S. Duesberg, Phys. Chem. Chem. Phys., 2013. 30. C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone, and S. Ryu, ACS Nano, 2010, 4, 2695–2700. 31. H. Li, Q. Zhang, C. C. R. Yap, B. K. Tay, T. H. T. Edwin, A. Olivier, and D. Baillargeat, Adv. Funct. Mater., 2012, 22, 1385–1390. 32. G. Eda, H. Yamaguchi, D. Voiry, T. Fujita, M. Chen, and M. Chhowalla, Nano Lett., 2011, 11, 5111–5116. 33. Y. Zhan, Z. Liu, S. Najmaei, P. M. Ajayan, and J. Lou, Small, 2012, 8, 966–971. 34. R. G. Compton and C. E. Banks, Understanding Voltammetry, World Scientific, 2011. 35. B. E. Conway and B. V. Tilak, Electrochimica Acta, 2002, 47, 3571– 3594. 36. J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou, X. W. (David) Lou, and Y. Xie, Adv. Mater., 2013, n/a–n/a. 37. C.-B. Ma, X. Qi, B. Chen, S. Bao, Z. Yin, X.-J. Wu, Z. Luo, J. Wei, H. Zhang, and H. Zhang, Nanoscale, 2014. 38. C. Yim, M. O’Brien, N. McEvoy, S. Winters, I. Mirza, J. G. Lunney, and G. S. Duesberg, Appl. Phys. Lett., 2014, 104, 103114. Notes a Centre for Research on Adaptive Nanostructures and Nanodevices (CRANN) & Advanced Materials Bio-Engineering Research Centre (AMBER),Trinity College Dublin, Dublin 2, Ireland. E-mail: [email protected], [email protected] b School of Chemistry, Trinity College Dublin, Dublin 2, Ireland. † Electronic Supplementary Information (ESI) available: [Detailed XPS analysis, further TEM, ellipsometry measurements, AFM]. See DOI: 10.1039/b000000x/

This journal is © The Royal Society of Chemistry [year]