Intercalated 2D MoS2 Utilizing a Simulated Sun Assisted Process

Jan 7, 2016 - Tel: +61 3 9925 3254 (K.K.)., *E-mail: [email protected]. Tel: +61 3 9925 2946 (J.Z.O.)., *E-mail: [email protected]. Tel...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF CALIFORNIA SAN DIEGO LIBRARIES

Article 2

Intercalated 2D MoS Utilizing a Simulated Sun Assisted Process: Reducing the HER Overpotential Yichao Wang, Benjamin Carey, Wei Zhang, Adam F Chrimes, Liangxian Chen, Kourosh Kalantar-zadeh, Jian Zhen Ou, and Torben Daeneke J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b10939 • Publication Date (Web): 07 Jan 2016 Downloaded from http://pubs.acs.org on January 13, 2016

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

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32

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

The Journal of Physical Chemistry

Intercalated 2D MoS2 Utilizing a Simulated Sun Assisted Process: Reducing the HER Overpotential Yichao Wang,† Benjamin J. Carey,† Wei Zhang,† Adam F. Chrimes, † Liangxian Chen,‡ Kourosh Kalantar-zadeh,*,† Jian Zhen Ou,*,† and Torben Daeneke*,† †

School of Electrical and Computer Engineering, RMIT University, Melbourne, Vic 3000,

Australia ‡

Institute for Advanced Materials and Technology, University of Science and Technology

Beijing, Beijing 100083, P. R. China

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry

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

Page 2 of 32

ABSTRACT: Molybdenum disulphide (MoS2) in its two dimensional (2D) morphology has favorable catalytic properties for the hydrogen evolution reaction (HER). To fully take advantage of this capability and in order to enhance this material’s HER performance, viable methods for synthesizing and tuning electronic properties of 2D MoS2 should be developed. Here, we demonstrate a facile and non-hazardous approach to partially intercalate Li+ ions into the 2D MoS2 host structure. We show that such an intercalation is possible in a non-hazardous saline Li+ containing solution using a grinding/sonication assisted exfoliation method in both dark and simulated sun irradiation conditions. A partial phase transformation from semiconducting (2H) to metallic (1T) phase is observed for the majority of flakes, after the process. We observe a notable enhancement of the HER activity for samples prepared under light illumination in the presence of a Li+ containing solution, in comparison to both pristine 2D MoS2 samples and samples processed in dark. This method provides an effective approach for phase engineering of 2D MoS2 for enhancing its HER performance.

ACS Paragon Plus Environment

2

Page 3 of 32

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

The Journal of Physical Chemistry

1. INTRODUCTION Two-dimensional (2D) transition metal dichalcogenides (TMDs), in particular 2D molybdenum disulphide (MoS2), have attracted significant attention in recent years due to their unique electronic and optical properties.1-6 Depending on the S atoms arrangement, MoS2 usually exhibits two distinct crystal phases at room temperature: 2H with trigonal prismatic and 1T with octahedral coordination.7 The two phases can reversibly convert from one to the other through intra-layer atomic plane gliding, which involves a transversal displacement of one of the S planes.8 While the 2H phase features semiconducting electronic behaviour with a direct bandgap energy of ~1.9 eV for single atomic layer MoS2, the 1T phase demonstrates metallic properties. Two dimensional MoS2 has found numerous applications including those in the development of field-effect transistors (FETs),9-12 photodetectors,13 energy storages,14-15 gas separation membranes,16 bio and gas sensors,17-20 and photovoltaics.21 This material has also been found to be an economical alternative to noble metals catalysts for the electrochemical hydrogen evolution reaction (HER).22-25 Both theoretical and experimental studies show that the active catalytic sites for HER on semiconducting 2H 2D MoS2 are located along the edges of the S−Mo−S plane, while the basal planes remain chemically inert.26 Mo− edges were found to contribute to a much larger extend to the catalytic activity of the material than S−terminated edges due to the lower hydrogen binding energy of exposed Mo sites.27 The phase transformation of MoS2 to metallic 1T phase can further enhance its HER performance, possibly due to the reduction of the charge transfer resistance28 and the activation of the inactive basal plane.29 Therefore, optimisation of the HER performance of 2D MoS2 can be achieved through activating catalytic sites and phase engineering.

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry

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

Page 4 of 32

So far, there have been several approaches to realise the 2H→1T phase transition, such as electrochemical alkali ion (Li+, Na+ and K+) intercalation,2 chemical intercalation of MoS2 utilising highly reducing organo-alkali compounds and the direct chemical synthesis of 1T nanosheets.30-32 Electrochemical and chemical alkali ion intercalation appear to be effective methods to achieve the phase transition. Specifically, Li+ ion intercalation can decrease the hydrogen binding energy at Mo− edges and initialize the MoS2 phase transition.2-3 At the same time, the Li+ ion intercalation facilitates the exfoliation of bulk MoS2 into individual or few layers flakes, which can also improve the electrical conductivity and increase the surface to volume ratio. The chemical insertion of Li+ ions into the MoS2 lattice has however its own drawbacks. It is a slow reaction process, leading to long reaction times and/or requiring elevated temperatures.30, 33 Currently, one of the most commonly used methods for achieving Li+ intercalation into MoS2 is the reaction of bulk MoS2 with butyllithium suspended in a moderate boiling point solvent. This reaction has to occur in an inert atmosphere since butyllithium is a very reactive and explosive reagent.30 The long reaction time implies energetically unfavourable reaction kinetics. The process also results in the deterioration of the crystallinity and 2D configuration of the flakes.33 Another popular process for Li+ intercalation is the electrochemical method.2 In this process, the driving force for the intercalation reaction is provided by electrochemical means avoiding hazardous reagents and reducing reaction time. In this case, pre-exfoliated 2D MoS2 nanoflakes are typically drop-casted onto a conductive substrate and a Li+-containing solution is utilised as the electrolyte. The drawback is that the MoS2 nanoflakes have to be transferred from suspension onto a substrate followed by an annealing treatment, aiming to fully evaporate the process

ACS Paragon Plus Environment

4

Page 5 of 32

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

The Journal of Physical Chemistry

solvent. Annealing can easily lead to the oxidation of the MoS2 surface.34 Moreover, due to the poor conductivity of MoS2, this method only affects MoS2 in the vicinity of the surface of the working electrode substrate and several intercalation steps have to be carried out to produce an appreciable yield. These extra steps can cause excessive exfoliation of the film and cause undesirable pin-holes and cracks in the films made of 2D sheets. Therefore, a facile and non-hazardous Li+ intercalation process that can accelerate the exfoliation and 1T transformation reactions and simultaneously soften the reaction condition is sought after in order to form efficient HER electrocatalysts. In order to avoid the aforementioned problems, alternative approaches should be developed. It has been observed that sonication treatment can assist the intercalation process of guest molecules into the host stratified MoS2 structure.35 During this process, acoustic cavitation can be generated by ultrasonic forces, leading to the formation, growth and collapse of bubbles in the solution. The powerful collapses of these bubbles results in shock waves and the local increase in temperature and pressure, which can assist the guest molecule intercalation. Similarly, photoirradiation treatment was found to be a suitable method for the activation of electrochemical intercalation.36-37 Electron-hole pairs are generated in the host material upon photo-irradiation. It was found that the photo generated electrons remain in the host material, while the holes are transferred to the electrolyte, presumably reacting with water or solvent molecules. Band bending effects facilitate this process when working with solid MoS2 electrodes. The material is polarised in this situation, producing an electric field that provides an assisting driving force for Li+ ion intercalation.38

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry

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

Page 6 of 32

In this study, we demonstrate sonication-assisted Li+ intercalation with and without the irradiation of broad spectrum light of a sun simulator, producing 2D Li+ intercalated MoS2 nanoflakes with dominant 1T crystal phase. The effects of intercalating Li+ ions and the phase transformation on the electronic and optical properties of 2D MoS2 nanoflakes are investigated. Finally the catalytic HER performance of these materials is evaluated in an electrolysis cell. 2. EXPERIMENTAL SECTION 2.1. Materials. MoS2 powder was purchased from Sigma Aldrich (Fluka, Product Number: 69860, 99% purity). The powder had an average particle size of 6 µm. N-methyl-2-pyrrolidone (NMP, 99% anhydrous) and lithium perchlorate (LiClO4, 98% purity) were purchased from Sigma Aldrich and used without further purification. 2.2. Synthesis of 2D MoS2 and Li+ intercalated MoS2 Nanoflakes. Suspensions of 2D MoS2 and LixMoS2 were produced using a facile two-step grinding/sonication process,39-40 For the synthesis of 2D MoS2 nanoflakes, 150 mg of MoS2 powder was added to 0.4 mL of NMP in a mortar and manually ground for 30 minutes. The mixture was placed in a vacuum oven to dry overnight at room temperature, then re-dispersed in 10 mL NMP. The obtained mixture was probe-sonicated (Ultrasonic Processor GEX500) for 90 minutes at a power of 125 W (20 kHz). The supernatant containing 2D MoS2 nanoflakes was collected after being centrifuged for 90 minutes at a speed of 4000 rpm. The sonication time was chosen based on previously published studies.2 The same sonication duration was used for sonication-assisted intercalation process, both in the dark and under illumination. For the synthesis of 2D Li+ intercalated MoS2 nanoflakes, 0.1 g LiClO4 was added to the solution of re-dispersed MoS2 in NMP prior to sonication. The obtained suspension was divided into two equal parts. The first aliquot was

ACS Paragon Plus Environment

6

Page 7 of 32

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

The Journal of Physical Chemistry

probe-sonicated for 90 min at 125 W power in the dark, the second aliquot was irradiated while sonicating, using a sun simulator modified from a broad spectrum light system (AM1.5 simulated light, Abet Technologies LS-150) at a power density of 100 mW cm-2. The intensities of the solar simulator light at its different wavelengths are presented in Figure S1 in the Supporting Information. The same sonication duration of 90 minutes was used for sonicationassisted intercalation process, both in dark and under the sun simulator illumination. The schematics of the three processes are presented in Figure 1. The two batches of peocessed nanoflakes were collected in the supernatant after centrifugation for 90 min at 4000 rpm. 2.3. Characterisation of 2D MoS2 and Li+ intelated MoS2 nanoflakes. The nanoflake suspensions were diluted and their average lateral dimensions were measured using a dynamic light scattering (DLS) equipment (ALV fast DLS particle-sizing spectrometer). Atomic force microscopy (AFM - Phillips D3100, Digital Instruments) in tapping mode was used for assessing the thickness of nanoflakes. The nanoflake suspensions were dropped onto a 300 mesh copper holey carbon grid (ProSciTech, Australia) for high resolution transmission electron microscopy (HRTEM) imaging (JEOL 2100F) characterisation. The crystal structure was characterised by Xray diffraction (XRD), with a Bruker D4 Endeavour wide-angle diffractometry equipped with a (Cu K-alpha) 0.15418 nm X-ray source. Raman spectra were obtained using a micro-Raman spectrometer (Renishaw InVia microscope) with a 2400 l/mm blazed grating and a 457 nm, 20 mW laser excitation source. Acquisitions were carried out for 20 seconds with three averages using the ×100 lens. X-ray photoelectron spectroscopy (XPS) measurements were performed on a Thermo scientific K-Alpha VG-310F instrument using aluminium monochromated X-rays (20 kV, 15 mA) with the hemispherical energy analyser set at a pass energy of 20 eV for the peak scans. A Cary 60 spectrophotometer (Agilent Technologies) was used for obtaning UV-vis

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry

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

Page 8 of 32

spectra. The nanoflake suspensions were diluted 6-fold in NMP before UV-vis measurement. The photoluminescence (PL) spectra of 2D nanoflake suspensions were obtained using a PerkinElmer LS 55 luminescence spectrometer at multiple excitation wavelengths from 250 to 500 nm. 2.4. Hydrogen evolution reaction. HER performances of pristine 2D MoS2 and MoS2 nanoflakes (processed in dark and under simulated sun illumination) were evaluated using a potentiostat (CHI760D, CH Instrument Co.) in linear sweep mode (scan rate 0.01 V/s) between −0.5 and 0.1 V vs RHE and impedance spectroscopy (EIS) technique. A conventional threeelectrode setup was used for these measurements. The working electrodes were made by dropcasting of 2D MoS2 and LixMoS2 suspensions onto a mirror polished glassy carbon (MPGC) substrate. The surface area of 0.13 cm2 was defined through a kapton tape mask with defined dimensions. An Ag/AgCl (in 3 M KCl) reference electrode was used and a platinum wire was the counter electrode. The electrolyte contained 0.5 M H2SO4 in deionised water. 3. RESULTS AND DISCUSSION The lateral dimensions of the nanoflakes were measured by DLS (see Figure 2a), and the distribution of the dimension was determined by an equation (Supporting Information) developed by Lotya et al.41 All samples were found to show high polydispersity. The median lateral dimension for pristine 2D MoS2 nanoflakes was ~110 nm. The flake lateral dimension was reduced by approximately 8% upon sonication with the addition of LiClO4 in the dark and there was a further ~14% reduction of the size under simulated sun light irradiation.

ACS Paragon Plus Environment

8

Page 9 of 32

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

The Journal of Physical Chemistry

Figure 1. The two-step grinding/sonication process: (a) sonication in NMP; (b) sonication in NMP with LiClO4 and (c) sonication in NMP with LiClO4 under simulated sun irradiation.

The thicknesses of the nanoflakes were assessed using AFM. Figures 2b−d present representative AFM images for pristine 2D MoS2 and processes nanoflakes. The AFM profiles clearly indicate that the change in the 2D nature of the flakes is retained after Li+ intercalation with and without the presence of sun simulation.

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry

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

Page 10 of 32

Figure 2. (a) The lateral dimension distribution of the 2D MoS2 and processed nanoflakes using DLS; AFM profiles of (b) 2D MoS2; (c) 2D nanflakes synthesised in the prsence of LiClO4 without illumination, and (d) 2D under simulated sun irradiation (scale bar is 100 nm) and the statistical analysis of the layer thickness distribution of the samples (e, f, and g).

The thicknesses of pristine 2D MoS2 ranges from 1.5 to 3.5 nm, corresponding to between 2 and 5 fundamental layers (thickness of a single MoS2 monolayer is around 0.7 nm) which are shown

ACS Paragon Plus Environment

10

Page 11 of 32

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

The Journal of Physical Chemistry

in Figures 2b and 2e.42 The thickness of Li+ intercalated MoS2 nanoflakes synthesised in the dark and under light irradiation lies within a similar range (Figures 2c, 2d, 2f, and 2g). HRTEM images of the edges of the nanoflakes in Figure S2 (Supporting Information) show that 2D MoS2 exhibits near perfect planar structure (Figure S2a), while Li+ intercalated MoS2 nanoflakes have disordered polycrystalline features (Figure S2b and c). This can be rationalised by the fact that the intercalated Li+ ions reside within the van der Waal’s gap of MoS2, resulting in topotactical transformation, expansion of crystal structure and eventually cracking of the crystal.3, 43 As can be seen from Figure 2g, the exposure to simulated light seems to result in the larger number of layers in flakes. The maximum number of flakes in sun simulated exposed samples is 4, while it is 2 for non-processed MoS2 flakes. This can be due to faster cracking of the flakes during the light illumination that cracks the flakes before the full exfoliation occurs. The crystal structures of the pristine and processed 2D nanoflakes were investigated using XRD observations. The XRD patterns shown in Figure 3a lead to the conclusion that the 2D MoS2 nanoflakes are presented in the 2H MoS2 phase with a dominant peak at 14.5o, reflecting the (002) plane (ICDD card no. 77-1716).44 The spacing between adjacent layers in the lattice is c/2 = 0.61 nm. The processed samples synthesised both in the dark and when illuminated present shifted (002) plane reflections at 13.6o and 13.4o, corresponding to interlayer distances of c1/2 = 0.65 nm and c2/2 = 0.66 nm respectively. This provides strong evidence for the intercalation of Li+ which we link to the expansion of the lattice along the c-axis and formation of substoichiometric compounds.45 The increased expansion of the interlayer distance in the case of the sun light-assisted intercalation sample indicates a higher degree of Li+ intercalation, likely due to slightly increased reaction driving force and the induced optical field. Broadening of the (002) peak of the sonicated sample in the presence of LiClO4 indicate loss of crystallinity of the

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry

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

Page 12 of 32

flakes and increased polydispersity. This is likely due to cracking of the crystal due to the higher degree of Li+ intercalation,46 which is well consistent with the HRTEM observations presented in Supporting Information. The effects of the sonication assisted Li+ intercalation on the crystal structure of 2D MoS2 nanoflakes was also investigated using Raman spectroscopy. The Raman spectrum of reference 2D MoS2 in Figure 3b has two prominent peaks which are identified as the in-plane mode   located at 383 cm-1 and the out-of-plane mode  at 407 cm-1.44, 47-49

 The  Raman mode is shifted towards lower wavenumbers of 380 cm-1 (in comparison to 383

cm-1 for pristine 2D MoS2) for both of the sonicated samples in the presence of LiClO4 with and without exposure to sun simulators. There was no observable change in the position of the  mode after the sonication. This indicates that the intercalated Li+ ions mainly suppresses the vibration of Mo−S bonds due to additional strain in the crystal lattice.44 This provides further evidence that when the bulk MoS2 is probe-sonicated in the presence of LiClO4, the Li+ ions are likely to be inserted into the exfoliated 2D nanoflakes under acoustic cavitation forces.50 Under sun light irradiation an additional field is providing driving force possibly enhancing the Li+ intercalation process in the 2D MoS2 nanoflakes.50 The intercalated Li+ ions in the 2D MoS2 planes occupy the interlayer S−S tetrahedron sites in 1T LixMoS2, resulting in a 2H→1T phase transition. The synthesis results in polytype superlattices and the formation of LixMoS2 compounds.51

ACS Paragon Plus Environment

12

Page 13 of 32

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

The Journal of Physical Chemistry

Figure 3. (a) XRD patterns of 2D MoS2 and LixMoS2 nanoflakes. (b) Raman spectra of 2D MoS2 and the spectra change after Li+ ion intercalation. (c) XPS spectra of elemental Mo before and after intercalation. (d) XPS spectra of elemental Li after intercalation.

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry

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

Page 14 of 32

The wavenumber difference between the two Raman vibration modes can be used for identifying the number of MoS2 layers47 and this difference can be associated to the number of layers.52 As  Raman mode peak shifts to a lower wavenumber after Li+ ion can be seen from Figure 3b,   intercalation and the difference between  and  increases. This means there is an increase

of the thickness according to the Raman measurements which is consistent with the AFM results in Figure 2. The phase composition of the intercalated nanoflakes was studied by XPS. For the pristine 2D MoS2, as shown in Figure 3c, the Mo 3d spectrum consists of peaks at approximately 229.5 and 232.5 eV that correspond to Mo4+3d5/2 and Mo4+3d3/2 components, respectively, which indicates that the Mo keeps the +4 oxidation state and the pristine 2D crystal structures is dominated by the 2H semiconducting phase.2, 25 When the bulk MoS2 is probe sonicated in the presence of LiClO4 under the exclusion of light, shoulders appear at 231.3 and 228.2 eV. Deconvolution of the signal reveals additional peaks that are shifted to lower binding energies by ~1.2 eV with respect to the Mo 3d peak positions of the pristine MoS2, suggesting that parts of the crystal exhibit a 2H→1T phase transition.2, 30 The composition ratio between 1T and 2H phases is nearly 0.5 and the amount of intercalated Li+ ions (x) equals to approximately 0.3 for LixMoS2, estimated by the ratio of the 2H to 1T Mo 3d5/2 peak areas using the method demonstrated by Wang et al..22, 30 Deconvolution of the XPS data corresponding to the light-assisted intercalation sample reveals that the binding energies of these emerging peaks are downshifted and their intensities are further enhanced, indicating that a greater portion of the 2D MoS2 nanoflakes undergo the 2H→1T phase transition. The eventual phase composition ratio between 1T and 2H is increased to 1.33 and x = 0.8 for LixMoS2.22, 30 The photoenhanced Li intercalation yield in 2D LixMoS2 nanoflakes is directly confirmed by the increase of the Li 1s region count (Figure 3d).

ACS Paragon Plus Environment

14

Page 15 of 32

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

The Journal of Physical Chemistry

All in all, the sonication assisted intercalation method, both in the dark and under simulated sun illumination, was proven to be an efficient approach to alter the phase composition of 2D MoS2. The electronic band structures and optical properties of pristine 2D MoS2 and intercalated LixMoS2 were studied further by measuring their absorption spectra using UV−vis absorption spectroscopy (Figure 4). The spectrum of pristine MoS2 shows two distinct absorption maxima at approximately 615 nm and 670 nm, corresponding to A and B excitonic peaks arising from the K point of the Brillouin zone.53-54 After Li+ intercalation, these excitonic A and B peaks are quenched. This provides further evidence that the crystals are dominated by 1T phase.55 Interestingly, for the samples sonicated in the presence of LiClO4, a new absorption peak at ~350 nm appears which is more prominent for the samples synthesised under light irradiation. This additional peak represents the plasmon resonance of LixMoS2 originating from the increased free charge carrier concentration (electrons) generated by the Li+ ions intercalation, which is in accordance with our previous study.3

Figure 4. UV−vis absorbance spectra of the 2D MoS2 and LixMoS2 nanoflakes. ExA and ExB indicate the excitonic resonances A and B of 2H MoS2.

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

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

Page 16 of 32

The PL properties of pristine 2D MoS2 nanoflakes and intercalated nanoflakes, prepared both in the dark and under light irradiation, were also investigated. The PL spectra of each sample excited at various wavelengths (250, 300, 350, 400, 450, and 500 nm) are shown in Figure 5 a-f. It was found that the pristine 2D MoS2 nanoflakes have excitation-dependent PL emission characteristics.30,

56-57

Red shifting of the PL emission peaks with increasing excitation

wavelengths was observed (Figure 5 a-f) which has been linked to quantum confinement effects in laterally polydisperse samples.18, 58-59 Upon the absorption of photons with sufficient energy, excitons are generated.60 These electron-hole pairs remain attracted to each other through coulomb forces which are relatively strong in smaller nanoflakes, increasing confinement of excitons in low dimensional materials. Therefore low energy PL emission peaks seen here correspond to the direct band-gap recombination of flakes with relatively large lateral dimension, while both large and small flakes contribute to the high energy PL emission. The samples of 2D MoS2 intercalated with Li+ ions show more quenched PL properties. This is likely to be linked to the phase transition of the nanoflakes from the original semiconducting (2H) to the metallic phase (1T).2 However, it is observed that the PL intensity is quenched to a much larger degree at smaller emission wavelengths, while there is almost no change for PL measurements excited using 450 and 500 nm wavelength sources. This indicates that sonication-assisted Li+ intercalation is more efficient in smaller flakes compared to larger flakes regardless of whether sun light is applied during the grinding/sonication process. This is likely due to the existence of a larger density of surface and edge defects in small flakes, that are favourable for ion intercalation.27

ACS Paragon Plus Environment

16

Page 17 of 32

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

The Journal of Physical Chemistry

Figure 5. PL spectra of MoS2 and LixMoS2 under excitation wavelengths of (a) 250, (b) 300, (c) 350, (d) 400, (e) 450, and (f) 500 nm.

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry

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

Page 18 of 32

To gain insight on whether the phase transition from 2H to 1T can lead to increased catalytic activity and therefore reduce the overpotential of HER, the materials were tested in an electrolysis cell. Figure 6a shows the current density (J) plotted against the applied potentials vs reversible hydrogen electrode (RHE) for 2D MoS2 and LixMoS2 nanoflakes. The obtained polarisation curves after iR-correction for LixMoS2 show high catalytic activity, with an onset of HER activity at approximately −150 mV vs RHE. This onset potential is similar to most of the MoS2 based catalysts.61 The excellent H2 evolution currents of 3.1 and 7.2 mA/cm2 were observed at about −300 mV vs RHE. The HER electrocatalytic activity of 2D LixMoS2 nanoflakes was further investigated through Tafel plot analysis, wherein the slopes were found to be 110, 92 and 58 mV per decade (mV/dec) for 2D MoS2, LixMoS2 synthesised without and with simulated sun irradiation, respectively (Figure 6b). These numbers show that while for nonirradiated sample the Tafel reaction dominates for the sun simulated irradiated sample the Volmer reaction dominates.62 It is important to mention that the observed Tafel slope for Li+ intercalated samples is still higher than that of previous reports on pure 1T MoS2 in which Tafel slopes in the order of 40 mV/dec are reported

22, 28-29

The outcomes of our study, however, are comparable with defect rich MoS2

systems.63 Xie et al. engineered defect structures which led to the increased exposure of active edge sites through the formation of cracks on the nanosheet surface. The Tafel slope was found to be in the order of 50 mV/dec when defects were introduced into the MoS2 structure.63 The higher conductivity of the processed LixMoS2 samples is due to the reduction of resistivity across the basal planes of the flakes, leading to more efficient electron transfer across these planes. This is also confirmed by electrochemical impedance spectroscopy which is presented in

ACS Paragon Plus Environment

18

Page 19 of 32

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

The Journal of Physical Chemistry

Figure 6c. As can be seen, the Nyquist plots reveal that in comparison to non-processes MoS2 sample, decreased charge-transfer resistance is seen for both LixMoS2 samples. The effect is much more dramatic for the sample that was developed by the exposure to simulated light, resulting in a relatively more facile electron transfer between LixMoS2 flakes and the HER electrode. The reaction kinetics are significantly enhanced in the samples of sonication-assisted Li+ intercalated MoS2, synthesised in the dark and under broad spectrum illumination. The Tafel slope was found to be remarkably reduced by 16.4% and 47.3% for samples synthesised in the dark and under sun illumination respectively, when compared to 2D MoS2. The low Tafel slope suggests that the hydrogen desorption process is the rate determining step.29,

62, 64

There are

several factors that may give rise to the enhanced HER activity of 2D LixMoS2. Mo-edges contribute to the catalytic activity in a much larger extent than S-edges due to the lower hydrogen binding site energy. Introducing Li+ ions into the 2D structure can lower the S-edges site energy while maintaining the site energy of Mo-edges, hence increasing the activity of sites with previously low activity.27 Secondly, samples synthesised under simulated sun illumination are smaller in lateral dimension, leading to a larger ratio of highly active edge sites to the overall material volume.27 It seems that in our work the number of layers has less effect on the reduction of HER overpotential in comparison to the lateral dimension. The number of layers is 2 for exfoliated MoS2 flakes and 4 for the light processed flakes, and the MoS2 flakes have also larger lateral dimensions, hence relatively larger surface-to-volume ratio, than the light processed flakes. Obviously, the effect of surface defects and edges are more important in making the HER system more efficient than enhanced surface-to-volume ratio in this specific case.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry

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

Page 20 of 32

Furthermore the increased d-spacing may also reduce steric hindrance at the flake edges, increasing the accessible volume around the catalytic sites leading to increased activity. The increased conversion yield to the 1T metallic phase during broad spectrum illumination leads also to increased conductivity of nanoflakes, which will reduce ohmic losses during electrolysis.27

Figure 6. (a) Polarisation curve. (b) Tafel slopes for the three samples. (c) EIS Nyquist plots of the three samples.

ACS Paragon Plus Environment

20

Page 21 of 32

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

The Journal of Physical Chemistry

Altogether, the increased HER activity in our study is likely to be a combination of the aforementioned effects. The decrease of the Tafel slope suggests that the density of both active catalytic edges and defect sites are increased through the intercalation of Li+ ions. The catalyst property was also increased due to the partial 2H→1T phase transition, resulting in enhanced HER reaction kinetics and improved conductivity.27 Further optimisation of the intercalation process and improved starting materials consisting of smaller MoS2 flakes could result in better outcomes. Improvements in the design of the electrolysis cell may also lead to higher current densities. In addition to the HER activity, another crucial parameter for the practical applicability of the catalyst is the stability test. In order to evaluate the stability of the catalyst in the acidic solution (0.5 M H2SO4), a stability test was conducted using LixMoS2 synthesised by light assisted intercalation method in the HER. The cathodic current was measured in time at the applied potential −300 mV for 7200 s. As present in Figure S3 in the Supporting Information, there is only a moderate loss of ~15% after 7200 s of continuous running, suggesting reasonable durability. 4. CONCLUSIONS In summary, we successfully demonstrated a facile synthesis method for the preparation of LixMoS2 based on grinding / probe sonication in presence of a Li+ containing salt (LiClO4) in the presence and absence of simulated sun irradiation. After Li+ ion intercalation, nanoflakes with smaller lateral dimensions and larger interlayer d-spacings were obtained. The success of the intercalation, achieving smaller flakes and the presence of Li+ ions in the interlayer spacing were demonstrated utilising a suit of characterisation techniques. The simultaneous phase transition of

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry

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

Page 22 of 32

a significant portion of the materials from the semiconducting 2H phase to a metallic 1T phase alongside the Li+ intercalation process was observed. Due the energy applied during the process and when Li+ ion presented in the solution, partial intercalation occurred and an increased defect density was created in the nanoflakes. The intercalation effect was more enhanced under the exposure to simulated sun. The LixMoS2 synthesised by our light assisted intercalation method exhibits high HER activity with small overpotential of 150 mV and Tafel slope of 58 mV/dec. The main advantage of the proposed process is its use of a green and non-hazardous process that only uses sun energy and lithium salt. It does not include hazardous liquids such as butyllithium and is highly scalable in comparison to electrochemical intercalation processes. The demonstrated method is straightforward and simple. As such, it is translatable to other intercalating guests and other TMDs hosts. ASSOCIATED CONTENT Supporting Information Solar light simulator spectrum, equation for the calculation of lateral dimensions of nanosheets, HRTEM image of a pristine 2D MoS2 nanoflakes, 2D nanoflakes synthesised in the presence of LiClO4 without light illumination, and 2D nanoflakes under simulated sun irradiation, and HER durability test. This information is available free of charge via the Internet at http://pubs.acs.org.

ACS Paragon Plus Environment

22

Page 23 of 32

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

The Journal of Physical Chemistry

AUTHOR INFORMATION Corresponding Author *Email: [email protected]; [email protected]; [email protected]. Tel: +61 3 9925 3254 Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS The authors gratefully acknowledge the facilities, and the scientific and technical assistance of the Australian Microscopy & Microanalysis Research Facility at the RMIT Microscopy & Microanalysis Facility, at RMIT University. REFERENCES 1.

Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S. Electronics and

Optoelectronics of Two-Dimensional Transition Metal Dichalcogenides. Nat. Nanotechnol. 2012, 7, 699-712. 2.

Wang, Y.; Ou, J. Z.; Balendhran, S.; Chrimes, A. F.; Mortazavi, M.; Yao, D. D.; Field,

M. R.; Latham, K.; Bansal, V.; Friend, J. R.; et al. Electrochemical Control of Photoluminescence in Two-Dimensional MoS2 Nanoflakes. ACS Nano 2013, 7, 10083-10093. 3.

Wang, Y.; Ou, J. Z.; Chrimes, A. F.; Carey, B. J.; Daeneke, T.; Alsaif, M. M. Y. A.;

Mortazavi, M.; Zhuiykov, S.; Medhekar, N.; Bhaskaran, M.; et al. Plasmon Resonances of Highly Doped Two-Dimensional MoS2. Nano Lett. 2015, 15, 883–890.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry

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

4.

Page 24 of 32

Coleman, J. N.; Lotya, M.; O’Neill, A.; Bergin, S. D.; King, P. J.; Khan, U.; Young, K.;

Gaucher, A.; De, S.; Smith, R. J.; et al. Two-Dimensional Nanosheets Produced by Liquid Exfoliation of Layered Materials. Science 2011, 331, 568-571. 5.

Nasir, M. Z. M.; Sofer, Z.; Ambrosi, A.; Pumera, M. A Limited Anodic and Cathodic

Potential Window of MoS2: Limitations in Electrochemical Applications. Nanoscale 2015, 7, 3126-3129. 6.

Cunningham, G.; Khan, U.; Backes, C.; Hanlon, D.; McCloskey, D.; Donegan, J. F.;

Coleman, J. N. Photoconductivity of Solution-Processed MoS2 Films. J. Mater. Chem. C 2013, 1, 6899-6904. 7.

Eda, G.; Fujita, T.; Yamaguchi, H.; Voiry, D.; Chen, M.; Chhowalla, M. Coherent

Atomic and Electronic Heterostructures of Single-Layer MoS2. ACS Nano 2012, 6, 7311-7317. 8.

Lin, Y. C.; Dumcenco, D. O.; Huang, Y. S.; Suenaga, K. Atomic Mechanism of the

Semiconducting-to-Metallic Phase Transition in Single-Layered MoS2. Nat. Nanotechnol. 2014, 9, 391-396. 9.

Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-Layer MoS2

Transistors. Nat. Nanotechnol. 2011, 6, 147-150. 10.

Liu, B.; Chen, L.; Liu, G.; Abbas, A. N.; Fathi, M.; Zhou, C. High-Performance

Chemical Sensing Using Schottky-Contacted Chemical Vapor Deposition Grown Monolayer MoS2 Transistors. ACS Nano 2014, 8, 5304-5314. 11.

Che, Y.; Lin, Y.-C.; Kim, P.; Zhou, C. T-Gate Aligned Nanotube Radio Frequency

Transistors and Circuits with Superior Performance. ACS Nano 2013, 7, 4343-4350.

ACS Paragon Plus Environment

24

Page 25 of 32

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

The Journal of Physical Chemistry

12.

Li, H.; Yin, Z.; He, Q.; Li, H.; Huang, X.; Lu, G.; Fam, D. W. H.; Tok, A. I. Y.; Zhang,

Q.; Zhang, H. Fabrication of Single- and Multilayer MoS2 Film-Based Field-Effect Transistors for Sensing NO at Room Temperature. Small 2012, 8, 63-67. 13.

Lopez-Sanchez, O.; Lembke, D.; Kayci, M.; Radenovic, A.; Kis, A. Ultrasensitive

Photodetectors Based on Monolayer MoS2. Nat. Nanotechnol. 2013, 8, 497-501. 14.

Weng, Q.; Wang, X.; Wang, X.; Zhang, C.; Jiang, X.; Bando, Y.; Golberg, D.

Supercapacitive Energy Storage Performance of Molybdenum Disulfide Nanosheets Wrapped with Microporous Carbons. J. Mater. Chem. A 2015, 3, 3097-3102. 15.

Wang, J. Z.; Lu, L.; Lotya, M.; Coleman, J. N.; Chou, S. L.; Liu, H. K.; Minett, A. I.;

Chen, J. Development of MoS2–CNT Composite Thin Film from Layered MoS2 for Lithium Batteries. Adv. Energy Mater. 2013, 3, 798-805. 16.

Berean, K. J.; Ou, J. Z.; Daeneke, T.; Carey, B. J.; Nguyen, E. P.; Wang, Y.; Russo, S. P.;

Kaner, R. B.; Kalantar-zadeh, K. 2D MoS2 PDMS Nanocomposites for NO2 Separation. Small 2015, 11, 5035-5040. 17.

Kalantar-zadeh, K.; Ou, J. Z.; Daeneke, T.; Strano, M. S.; Pumera, M.; Gras, S. L. Two-

Dimensional Transition Metal Dichalcogenides in Biosystems. Adv. Funct. Mater. 2015, 25, 5086-5099. 18.

Ou, J. Z.; Chrimes, A. F.; Wang, Y.; Tang, S. Y.; Strano, M. S.; Kalantar-zadeh, K. Ion-

Driven Photoluminescence Modulation of Quasi-Two-Dimensional MoS2 Nanoflakes for Applications in Biological Systems. Nano Lett. 2014, 14, 857-863. 19.

He, Q.; Zeng, Z.; Yin, Z.; Li, H.; Wu, S.; Huang, X.; Zhang, H. Fabrication of Flexible

MoS2 Thin-Film Transistor Arrays for Practical Gas-Sensing Applications. Small 2012, 8, 29942999.

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry

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

20.

Page 26 of 32

Zhu, C.; Zeng, Z.; Li, H.; Li, F.; Fan, C.; Zhang, H. Single-Layer MoS2-Based

Nanoprobes for Homogeneous Detection of Biomolecules. J. Am. Chem. Soc. 2013, 135, 59986001. 21.

Wi, S.; Kim, H.; Chen, M.; Nam, H.; Guo, L. J.; Meyhofer, E.; Liang, X. Enhancement of

Photovoltaic Response in Multilayer MoS2 Induced by Plasma Doping. ACS Nano 2014, 8, 5270-5281. 22.

Wang, H.; Lu, Z.; Xu, S.; Kong, D.; Cha, J. J.; Zheng, G.; Hsu, P. C.; Yan, K.; Bradshaw,

D.; Prinz, F. B.; et al. Electrochemical Tuning of Vertically Aligned MoS2 Nanofilms and its Application in Improving Hydrogen Evolution Reaction. Proc. Natl Acad. Sci. USA 2013, 110, 19701-19706. 23.

Hatay, I.; Ge, P. Y.; Vrubel, H.; Hu, X.; Girault, H. H. Hydrogen Evolution at Polarised

Liquid/Liquid Interfaces Catalyzed by Molybdenum Disulfide. Energy Environ. Sci. 2011, 4, 4246-4251. 24.

Ambrosi, A.; Sofer, Z.; Pumera, M. 2H→1T Phase Transition and Hydrogen Evolution

Activity of MoS2, MoSe2, WS2 and WSe2 Strongly Depends on the MX2 Composition. Chem. Commun. 2015, 51, 8450-8453. 25.

Ambrosi, A.; Sofer, Z.; Pumera, M. Lithium Intercalation Compound Dramatically

Influences the Electrochemical Properties of Exfoliated MoS2. Small 2015, 11, 605-612. 26.

Lukowski, M. A.; Daniel, A. S.; English, C. R.; Meng, F.; Forticaux, A.; Hamers, R. J.;

Jin, S. Highly Active Hydrogen Evolution Catalysis from Metallic WS2 Nanosheets. Energy Environ. Sci. 2014, 7, 2608-2613.

ACS Paragon Plus Environment

26

Page 27 of 32

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

The Journal of Physical Chemistry

27.

Benck, J. D.; Hellstern, T. R.; Kibsgaard, J.; Chakthranont, P.; Jaramillo, T. F. Catalyzing

the Hydrogen Evolution Reaction (HER) with Molybdenum Sulfide Nanomaterials. ACS Catal. 2014, 4, 3957-3971. 28.

Lukowski, M. A.; Daniel, A. S.; Meng, F.; Forticaux, A.; Li, L.; Jin, S. Enhanced

Hydrogen Evolution Catalysis from Chemically Exfoliated Metallic MoS2 Nanosheets. J. Am. Chem. Soc. 2013, 135, 10274-10277. 29.

Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.; Asefa, T.; Shenoy, V. B.; Eda, G.;

Chhowalla, M. Conducting MoS2 Nanosheets as Catalysts for Hydrogen Evolution Reaction. Nano Lett. 2013, 13, 6222-6227. 30.

Eda, G.; Yamaguchi, H.; Voiry, D.; Fujita, T.; Chen, M.; Chhowalla, M.

Photoluminescence from Chemically Exfoliated MoS2. Nano Lett. 2011, 11, 5111-5116. 31.

Zak, A.; Feldman, Y.; Lyakhovitskaya, V.; Leitus, G.; Popovitz-Biro, R.; Wachtel, E.;

Cohen, H.; Reich, S.; Tenne, R. Alkali Metal Intercalated Fullerene-Like MS2 (M = W, Mo) Nanoparticles and Their Properties. J. Am. Chem. Soc. 2002, 124, 4747-4758. 32.

Mahler, B.; Hoepfner, V.; Liao, K.; Ozin, G. A. Colloidal Synthesis of 1T-WS2 and 2H-

WS2 Nanosheets: Applications for Photocatalytic Hydrogen Evolution. J. Am. Chem. Soc. 2014, 136, 14121-14127. 33.

Benavente, E.; González, G. Microwave Activated Lithium Intercalation in Transition

Metal Sulfides. Mater. Res. Bull. 1997, 32, 709-717. 34.

Wu, J.; Li, H.; Yin, Z.; Li, H.; Liu, J.; Cao, X.; Zhang, Q.; Zhang, H. Layer Thinning and

Etching of Mechanically Exfoliated MoS2 Nanosheets by Thermal Annealing in Air. Small 2013, 9, 3314-3319.

ACS Paragon Plus Environment

27

The Journal of Physical Chemistry

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

35.

Page 28 of 32

Belova, V.; Möhwald, H.; Shchukin, D. G. Sonochemical Intercalation of Preformed

Gold Nanoparticles into Multilayered Clays. Langmuir 2008, 24, 9747-9753. 36.

Mangos, D. N.; Nakanishi, T.; Lewis, D. A. A Simple Method for the Quantification of

Molecular Decorations on Silica Particles. Sci. Tech. Adv. Mater. 2014, 15, 015002. 37.

Fan, X.; Xu, P.; Zhou, D.; Sun, Y.; Li, Y. C.; Nguyen, M. A. T.; Terrones, M.; Mallouk,

T. E. Fast and Efficient Preparation of Exfoliated 2H MoS2 Nanosheets by Sonication-Assisted Lithium Intercalation and Infrared Laser-Induced 1T to 2H Phase Reversion. Nano Lett. 2015, 15, 5956-5960. 38.

Frank, S. N.; Bard, A. J. Semiconductor Electrodes. 12. Photoassisted Oxidations and

Photoelectrosynthesis at Polycrystalline Titanium Dioxide Electrodes. J. Am. Chem. Soc. 1977, 99, 4667-4675. 39.

Carey, B. J.; Daeneke, T.; Nguyen, E. P.; Wang, Y.; Zhen Ou, J.; Zhuiykov, S.; Kalantar-

zadeh, K. Two Solvent Grinding Sonication Method for the Synthesis of Two-Dimensional Tungsten Disulphide Flakes. Chem. Commun. 2015, 51, 3770-3773. 40.

Nguyen, E. P.; Carey, B. J.; Daeneke, T.; Ou, J. Z.; Latham, K.; Zhuiykov, S.; Kalantar-

zadeh, K. Investigation of Two-Solvent Grinding-Assisted Liquid Phase Exfoliation of Layered MoS2. Chem. Mat. 2015, 27, 53-59. 41.

Lotya, M.; Rakovich, A.; Donegan, J. F.; Coleman, J. N. Measuring the Lateral Size of

Liquid-Exfoliated Nanosheets with Dynamic Light Scattering. Nanotechnology 2013, 24, 265703. 42.

Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C. Y.; Galli, G.; Wang, F.

Emerging Photoluminescence in Monolayer MoS2. Nano Lett. 2010, 10, 1271-1275.

ACS Paragon Plus Environment

28

Page 29 of 32

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

The Journal of Physical Chemistry

43.

Spahr, M. E.; Novák, P.; Haas, O.; Nesper, R. Electrochemical Insertion of Lithium,

Sodium, and Magnesium in Molybdenum(VI) Oxide. J. Power Sources 1995, 54, 346-351. 44.

Balendhran, S.; Ou, J. Z.; Bhaskaran, M.; Sriram, S.; Ippolito, S.; Vasic, Z.; Kats, E.;

Bhargava, S.; Zhuiykov, S.; Kalantar-zadeh, K. Atomically Thin Layers of MoS2 via a Two Step Thermal Evaporation-Exfoliation Method. Nanoscale 2012, 4, 461-466. 45.

Zheng, J.; Zhang, H.; Dong, S.; Liu, Y.; Tai Nai, C.; Suk Shin, H.; Young Jeong, H.; Liu,

B.; Ping Loh, K. High Yield Exfoliation of Two-Dimensional Chalcogenides Using Sodium Naphthalenide. Nat. Commun. 2014, 5, DOI: 10.1038/ncomms3995. 46.

Yan, S.; Xie, J.; Wu, Q.; Zhou, S.; Qu, A.; Wu, W. Highly Efficient Solid Polymer

Electrolytes Using Ion Containing Polymer Microgels. Polym. Chem. 2015, 6, 1052-1055. 47.

Li, H.; Zhang, Q.; Yap, C. C. R.; Tay, B. K.; Edwin, T. H. T.; Olivier, A.; Baillargeat, D.

From Bulk to Monolayer MoS2: Evolution of Raman Scattering. Adv. Funct. Mater. 2012, 22, 1385-1390. 48.

Ling, X.; Fang, W.; Lee, Y.-H.; Araujo, P. T.; Zhang, X.; Rodriguez-Nieva, J. F.; Lin, Y.;

Zhang, J.; Kong, J.; Dresselhaus, M. S. Raman Enhancement Effect on Two-Dimensional Layered Materials: Graphene, H-BN and MoS2. Nano Lett. 2014, 14, 3033-3040. 49.

Senthilkumar, V.; Tam, L.; Kim, Y.; Sim, Y.; Seong, M. J.; Jang, J. I. Direct Vapor Phase

Growth Process and Robust Photoluminescence Properties of Large Area MoS2 Layers. Nano Res. 2014, 7, 1759-1768. 50.

Benavente, E.; Santa Ana, M. A.; Mendizábal, F.; González, G. Intercalation Chemistry

of Molybdenum Disulfide. Coord. Chem. Rev. 2002, 224, 87-109. 51.

Wang, L.; Xu, Z.; Wang, W.; Bai, X. Atomic Mechanism of Dynamic Electrochemical

Lithiation Processes of MoS2 Nanosheets. J. Am. Chem. Soc. 2014, 136, 6693-6697.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry

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

52.

Page 30 of 32

Liu, K. K.; Zhang, W.; Lee, Y. H.; Lin, Y. C.; Chang, M. T.; Su, C. Y.; Chang, C. S.; Li,

H.; Shi, Y.; Zhang, H.; et al. Growth of Large-Area and Highly Crystalline MoS2 Thin Layers on Insulating Substrates. Nano. Lett. 2012, 12, 1538-1544. 53.

Chikan, V.; Kelley, D. F. Size-Dependent Spectroscopy of MoS2 Nanoclusters. J. Phys.

Chem. B 2002, 106, 3794-3804. 54.

Gopalakrishnan, D.; Damien, D.; Shaijumon, M. M. MoS2 Quantum Dot-Interspersed

Exfoliated MoS2 Nanosheets. ACS Nano 2014, 8, 5297-5303. 55.

Song, X.; Hu, J.; Zeng, H. Two-Dimensional Semiconductors: Recent Progress and

Future Perspectives. J. Mater. Chem. C 2013, 1, 2952-2969. 56.

Enyashin, A. N.; Seifert, G. Density-Functional Study of LixMoS2 Intercalates (0