• Tidak ada hasil yang ditemukan

1 s2.0 S2452262717300648 main

N/A
N/A
Protected

Academic year: 2017

Membagikan "1 s2.0 S2452262717300648 main"

Copied!
40
0
0

Teks penuh

(1)

Accepted Manuscript

Tailoring catalytic activities of transition metal disulfides for water splitting Seokhoon Choi, Ki Chang Kwon, Soo Young Kim, Ho Won Jang

PII: S2452-2627(17)30064-8

DOI: http://dx.doi.org/10.1016/j.flatc.2017.06.010

Reference: FLATC 31

To appear in: FlatChem

Received Date: 30 April 2017 Revised Date: 25 June 2017 Accepted Date: 25 June 2017

Please cite this article as: S. Choi, K.C. Kwon, S.Y. Kim, H.W. Jang, Tailoring catalytic activities of transition metal disulfides for water splitting, FlatChem (2017), doi: http://dx.doi.org/10.1016/j.flatc.2017.06.010

(2)

Tailoring catalytic activities of transition metal

disulfides for water splitting

Seokhoon Choi1,‡, Ki Chang Kwon1,2,‡, Soo Young Kim2,*, Ho Won Jang1,*

1Department of Materials Science and Engineering, Research Institute of Advanced Materials,

Seoul National University, Seoul 08826, Republic of Korea

2School of Chemical Engineering and Materials Science, Chung-Ang University, Seoul 06974,

Republic of Korea

These authors contributed equally.

KEYWORDS

Transition metal disulfide, hydrogen evolution reaction, water splitting, catalytic, molybdenum

(3)

Abstract

Development of clean and renewable energy to attain a sustainable society is one of the most

urgent tasks at present. In this respect, the production of hydrogen (H2) through electrochemical

or photoelectrochemical water splitting is a promising way to utilize sustainable energy source

such as sunlight. Recently, transition metal disulfides (TMDs) have received tremendous

attention as hydrogen evolution reaction (HER) catalysts due to their favorable chemical and

catalytic properties. Theoretical and experimental demonstrations have revealed that TMD-based

catalysts exhibit excellent catalytic activity at edge sites. However, there is limitation that the

basal plane which constitutes most of the TMD material remains chemically inert. Herein, we

overview recent progress in tailoring catalytic activities of TMDs for electrochemical and

photoelectrochemical hydrogen production and discuss the future strategies to overcome the

(4)

1. Introduction

The potential greenhouse gas emission of fossil fuels as well as their uncertain reserves has

triggered the need for renewable energy development to create a sustainable society. It has been

an urgent issue to find out renewable and carbon-neutral energy sources in research field.

Hydrogen has the highest energy density than any other chemical fuels and produces water for

combustion products. Therefore, hydrogen can be utilized as future energy carrier by storing

energy in a chemical bond between hydrogen atoms [1-3]. Since the atmosphere contains little

hydrogen, it must be separated from other sources. The most economically competitive method

for generating hydrogen is steam reforming of the fossil fuels [4], but there is a severe problem

of the greenhouse gas emission such as carbon dioxides. The other method for hydrogen

production is electrolysis of water by electrochemical (EC) [5, 6] or photoelectrochemical (PEC)

[1, 7, 8] water splitting. For practical purpose, the storing solar energy in the form of hydrogen

via water splitting is the ultimate goal to meet the growing global energy demand with reducing

dependence on fossil fuels. However, under standard conditions, the free energy change of ΔG =

237.2 kJ/mol is required to decompose one molecule of H2O to H2 and 1/2O2. Furthermore, the

charge transfer processes at solid/liquid interface are sluggish due to the high activation energy

of water splitting reaction. These process could be accelerated by loading electrocatalysts which

can reduce activation energy [9-11]. Noble metal catalysts, such as Pt, Pd, Ru, and Ir, exhibit

outstanding hydrogen evolution reaction (HER) performance [12-14]. However, the high cost of

noble-metal catalysts due to their scarcity limits their widespread utilization. Hence, it is

significant to develop the noble-metal-free electrocatalysts which are composed of earth

abundant materials for the sustainable water splitting devices.

During the past several years, there have been a surge in the development of noble-metal-free

(5)

[15, 16] and hydroxides [17, 18], carbon based materials and their hybrids [19-21]. Even though

these electrocatalysts display excellent improvement for HER, they still show lower HER

performance compared to Pt-based catalysts.

2-dimensional transition metal disulfides (TMDs) which have fascinating optical and

electrical properties became one of the most extensively investigated materials as HER catalysts

in the last decade [6, 22]. There are various types of bulk TMDs including insulators such as

HfS2, semiconductors such as MoS2 and WS2, to true metals such as NbS2 [23, 24]. Each layer of

TMDs which have the general formula of MS2 (M = transition metal) is composed of three

atomically stacked layer (S-M-S) and interconnected by weak van der Waals forces. When the

bulk TMDs become atomically thin 2D films, the 2D TMDs not only maintain their original

characteristics but also obtain additional properties due to quantum confinement effects [23].

Since 2005, MoS2 was spotlighted as a promising HER electrocatalyst by the study of the

Norskov’s group. They found that the binding free energy of atomic hydrogen to Mo-edge in the

MoS2 is close to zero which is comparable to Pt [25, 26]. To the best of our knowledge, it is the

first time that the edge sites of TMDs are identified as electrochemically active site.

Subsequently, other 2D TMDs (WS2, TiS2, and TaS2) were investigated by experimental and

theoretical studies [27-30]. Such studies demonstrated that other 2D TMDs could be also

efficient electrocatalysts for hydrogen production. However, one crucial shortcoming of a

TMD-based catalyst is the limited number of catalytically active edge sites since the most of the

surface consists of the electrochemically inactive basal plane in its presented form.

In this review, electrocatalysts based on 2D TMDs are covered in comprehensive manner

with major emphasis on the active site engineering. We summarize the major strategies for

tailoring catalytic activity of 2D TMDs with synthetic method. While the electrocatalysts based

(6)

obstacles that impede practical utilization. We will discuss the challenges to realize TMD based

catalysts comparable to noble metal catalysts such as Pt.

2. Tailoring catalytic activity

2.1. By nanostructuring

The layered chalcogenide materials including MoS2 show poor catalytic activity for the

HER in its bulk form [31]. Recent work showed MoS2 to be a promising electrocatalyst for the

HER. Both computational and experimental results confirmed that the HER activity stemmed

from the Mo-edges of MoS2 plates, while their basal planes were catalytically inert [25, 32]. For

these reasons, the nanosized MoS2 synthesized by various methods such as lithiation [33, 34] and

liquid exfoliation [35] should be more active for the HER catalysis than materials in bulk form.

The nanoparticles of layered materials usually exhibit platelet-like morphology, in which the

basal planes exposed, due to the anisotropic bonding and the general tendency to minimize its

surface energy. The increasing of the edge dimension is challenging for improving the catalytic

activity of layered chalcogenide materials [36]. Furthermore, the anisotropic conductivity of

MoS2 poses a challenge for high surface area structures as the conductivity of MoS2 is poor

along certain crystallographic directions [37].

Recently, the synthesis of nanostructured MoS2 have been extensively reported to improve

the HER catalytic activity of MoS2 [32, 38-40].As shown in Fig. 1A–D, Chen et al. [41] have

developed a core-shell MoO3-MoS2 nanowire catalyst with the following design criteria: (1) a

partially reduced MoO3 core with high conductivity, (2) a MoS2 shell (surface) with high

catalytic activity and stability in acids, and (3) a morphology of vertically aligned core-shell

nanowires to provide a high aspect ratio and a high surface area. The core-shell MoO3-MoS2

(7)

with a various temperature from 200 to 700 °C as shown in the scanning electron microscopy

(SEM) and transmission electron spectroscopy (TEM) images. The shapes and morphologies of

nanowires are gradually changed to the more blade-like (grass-like) surface. Nanowire with

synthesis temperature of 200 °C exhibits an onset for the HER at the overpotential at 1 mA/cm2

of approximately 150–200 mV, consistent with earlier results seen for nanostructured MoS2

particles [32, 38, 39]. The nanowire synthesized with the higher temperatures shows the lower

HER catalytic activity due to the excess sulfurization in the MoO3 core structure. The MoO3–

MoS2 core-shell nanowire could maintain the almost 100 % of its initial current density and

overpotential value at 10 mA/cm2 even after 10000 cycles. The partially reduced MoO3 core

serves as a nanostructured and conductive core, while the ultrathin and conformal MoS2 shell

serves as both a HER catalyst and a passivation layer.

In order to enhance the conductivity and the electrochemically active surface area (ECSA) it

can be an effective way to combine TMDs with conductive scaffolds. Li et al. [42] reported the

MoS2/reduced graphene oxide (RGO) composite nanostructures which exhibited excellent

electrocatalytic activity for HER. The MoS2/RGO catalyst was synthesized through facile

solvothermal method as illustrated Fig. 1E. The RGO plays a role as a nucleation sites for MoS2

nanosheets, enabling well dispersed growth of the MoS2 nanosheets. In contrast, the MoS2

nanosheets grown without RGO aggregate to spherical particles (see Fig. 1F). HRTEM image of

Fig. 1G shows highly dispersed MoS2 nanosheets on RGO template. As shown in Fig. 1H,

electrical coupling of MoS2 nanosheets with relatively more conducting RGO template

significantly reduced the charge-transfer impedance, leading to a small Tafel slope of 41 mV/dec.

Meng et al. [43] also reported p-MoS2/n-rGO composite nanostructures as a photocatalyst. The

nanoscale pn junction greatly enhanced PEC performance, shifting the onset potential, increasing

(8)

most semiconductor-based electrodes insufficiently absorb the entire solar spectrum and have a

limitation on recombination of electron-hole pairs. TMDs can enlarge the light absorption

spectrum of photocatalyst materials. Moreover, TMDs can be applied to reduce the activation

energy of hydrogen evolution and improve the separation of photogenerated electron-hole pairs

[44]. Liang et al. [45] demonstrated vertical VS2 nanoplate arrays on carbon paper grown by

facile hydrothermal method. The VS2 nanoplate arrays as a HER electrocatalyst exhibited

excellent catalytic activity with a small Tafel slope of 36 mV/dec, a high exchange current

density of 0.955 mA/cm2

, and a low overpotential of 42 mV at a current density of -10 mA/cm2 .

Because the VS2 which is the intrinsic metallic TMDs (e.g., VS2, NbS2, and TaS2) is catalytically

active in both basal planes and the edges, compared with MoS2 and WS2 that have catalytically

inert basal planes. Furthermore, the directly grown VS2 nanoplate arrays on conductive scaffold

(i.e., carbon paper) enabled the fast charge transport and also facilitated the detachment of

evolved hydrogen bubbles from the surface.

To increase the active edge sites and dangling bond on its surface, Kong et al. [46] have

demonstrated that the vertically aligned MoS2 and MoSe2 layers to maximally expose the

catalytic active sites by using a rapid sulfurization/selenization process as displayed in Fig. 2A.

Although the edge-terminated film has been considered unstable, it could be synthesized on

diverse substrates through a kinetically controlled rapid growth method. The chemical

conversion of the Mo thin films to the MoS2/MoSe2 occurs much faster than the diffusion across

the Mo thin film at the high temperature due to the anisotropic structure of MoS2/MoSe2 layer.

The edge-terminated thin films of MoS2 and MoSe2 layer was synthesized by controlling the

kinetics and temperature gradient as shown in Fig. 2B and C. The layers tend to be perpendicular

to the substrate, with exposed van der Waals gaps for fast reaction. They have evaluated the

(9)

The overpotential at 1 mA/cm2 of both materials are similar to each other with value of

approximately 350 mV. Interestingly, the turn over frequency (TOF) of MoS2 is very close to the

reported value of 0.016 s-1 obtained from combined scanning tunneling microscopy (STM)

studies and in situ electrochemical measurements on MoS2 nano-platelet catalysts [32]. It

confirmed that nearly all of the surface sites on synthesized edge-terminated films are

catalytically active. Due to the increased number of edge sites, the exchange current densities of

edge-terminated films were about ten times higher than previous MoS2 nanoparticle-based

electrodes and compare favorably to most common metal catalysts [32, 47].

Kwon et al. [48] demonstrated the wafer-scale fabrication of MoS2/p-Si heterojunction

using a thin-film transfer method for high-performance PEC hydrogen production.

Thickness-controlled MoS2 layers were synthesized by the thermolysis of a solution precursor layer on

SiO2/Si substrate as shown in Fig. 2E. The synthesized 2H-MoS2 layers contain a-domains with

layer-by-layer stacked (001) planes and c-domains with vertically stacked (100) planes. The

serrated surface of the synthesized 2H-MoS2 thin film provides a larger surface area than the

atomically flat surface of a single-crystalline MoS2 monolayer. Furthermore, the partially rotated

MoS2 basal planes, Moiré fringes, roughen the surface of the transferred MoS2 thin film as

observed in Fig. 2F. The films mainly contain a-domains with the preferred out-of-plane (001)

orientation of the 2H-MoS2 phase and c-domains with the preferred in-plane (001) orientation of

the 2H-MoS2 phase. The 13-nm-thick MoS2 layer heterojunction with p-Si photocathode have

displayed the high photocurrent density (24.6 mA/cm2 at 0 V vs. RHE), the large shifts in

overpotential (0.79 V at 10 mA/cm2), and the long-term stability (over 10000 s) as shown in Fig.

2G. The overpotential at 10 mA/cm2 of this heterojunction photocathode is lower than the those

(10)

at 0 V vs. RHE of this sample shows the highest value among the similar materials system.

These improved PEC performances compared to the atomically flat MoS2 layer originate from

the vertically stacked c-domains in synthesized thin films which can act as efficient catalytic

active region. They also have demonstrated the versatility of the synthesize thin films, the

various p-type III-V semiconductors were employed as photoelectrodes. When the 13-nm-thick

MoS2 films are transferred onto the p-InP, p-GaAs, and p-GaP photocathodes, the onset

potentials are shifted toward the anodic direction without notable losses in the saturation

photocurrents.

According to these previous reports, the nanostructured MoS2 could act as not only the

efficient catalyst but also the passivation layer for the stable electrolysis. The increase of

catalytic active edge sites is still main challenging issue for the efficient hydrogen production by

utilizing the TMD catalyst.

2.2 By anion doping

The electronic nature of TMDs can be manipulated by adjusting chemical composition. For

TMD nanosheets, the dopants are easily exposed on the surface of TMDs and take part in the

HER. Jin’s group [49] reported the amorphous MoQxCly (Q = S, Se) electrocatalyst for

electrochemical and photoelectrochemical hydrogen production. The catalytic activity of the

amorphous MoQxCly electrocatalysts is comparable to the state-of-art MoS2 based electrocatalysts.

According to ultraviolet photoelectron spectroscopy (UPS), the valence band edge position of the

amorphous MoQxCly is shifted close to the Fermi level, compared to the crystalline MoS2.

Obviously, the doping of Cl changes the electronic structures of MoS2 due to generation of defect

states into the band gap. Moreover, the additional Cl atoms increase disorder of the amorphous

(11)

MoQxCly electrocatalyst intergrating with n+pp+ Si micropyramid (MP) absorber. The n+pp+ Si

MPs were utilized in order to improve the light absorption of planar p-Si by scattering effect.

Furthermore, a built-in pn junction of n+p Si can enlarge the photovoltageof Si MPs by

modifying the limited p-Si/electrolyte junction [51]. The extra p+ layer on the backside of n+p Si

could improve the majority carrier (hole) collection, resulting in enhanced fill factor [52]. It is

also important to achieve high photocurrent density (see Fig. 3A). A uniform MoQxCly thin films

were deposited on the surface of the Si MPs using CVD method. As shown in Fig. 3B, MoSxCly

thin film exhibited a smooth surface, while MoSexCly thin film formed a nanogranular surface.

MoQxCly/n+pp+ Si MPs photocathodes showed a significant enhancement in the onset potential

and the photocurrent density as illustrated in Fig. 3C. The onset potential of MoSxCly/n+pp+ Si

MPs phothocathode showed cathodic shift of 0.14 V with respect to the MoSxCly/planar p-Si

photocathode due to the band bending of the n+pp+ structure. The onset potential of

MoSexCly/n+pp+ Si MPs also exhibited the cathodic shift of 0.18 V compared to MoSexCly/planar

p-Si. Furthermore, at 0 V vs. RHE, the photocurrent density of Si MPs photocathodes

significantly showed two- or three-fold increase compared to planar p-Si due to the improved

light-trapping property of Si MPs.

In recent reports, the phosphorus (P), which places on the near sulfur (S) in the periodic

table, was considered as a substitutional dopant for improving the catalytic activity to efficiently

generate the hydrogen [53-55]. Ye et al. [56] have reported the one-step approach to improve the

HER activity of MoS2 and molybdenum phosphide (MoP) via formation of a MoS2(1-X)PX (X = 0

to 1) solid solution. The bulk particulate MoS2 (diameter 6 m) was annealed at 750 °C with

different amounts of red phosphorus in Ar/H2 flow. They have calculated the hydrogen

adsorption Gibbs free energy by using the Projector Augmented Wave pseudopotentials, an

(12)

initio Simulation Package [57-61]. A supercell composed of 6 × 6 MoS2 primitive cells was

used to model the P-alloying and H adsorption as shown in Fig. 3D. They found that ΔGH = 2.2

eV for the pristine MoS2 surface and ΔGH < 0.8 eV for the neighnoring S atoms in the P-alloying

MoS2 surface. The drastic enhanced H adsorption on the P-alloying MoS2 surface could be stated

by using the “state-filling” model. The work required to accommodate the H electron in much

lower in the surface of the P-alloyed MoS2 than that of the pristine MoS2, leading to a stronger H

binding. To observe the morphology of the pristine and P-alloyed MoS2 particles, they measured

the SEM images and Brunauer-Emmett-Teller (BET) analysis. As shown in the Fig. 3E, the size

of the P-alloyed MoS2 particles became the larger than that of the pristine MoS2 after the

phosphorization reaction. Furthermore, the BET surface area MoS2(1-X)PX increased gradually

from 1.3 to 9.4 m2/g as a function of increase of X. To clarify the substitution of S by P in the

MoS2 lattice, the energy-dispersive-X-ray-spectroscopy (EDS)-scanning transmission electron

microscopy (STEM) of MoS0.94P0.53 was conducted as shown in the Fig. 3E. The Mo, P, and S

are mostly homogeneously distributed among the sample, except the region with low P intensity

which is marked by circle in the Fig. 3E. The P atoms could be easily substituted into the MoS2

lattice due to the similarity of the Mo–S and Mo–P bond lengths, indicating that the MoS2(1-X)PX

particles could be easily synthesized without breakage of Mo–S bonding. The HER activities of

MoS2(1-X)PX are gradually enhanced until the X value reaches at 0.53 (see Fig. 3F). It showed the

overpotential of 150 mV at the current density of 10 mA/cm2. However, the lower HER activities

were observed in the samples with the larger P ratio than S. Also, the Tafel slopes of these

samples have shown the similar tendency. The reaction with phosphorus significantly reduced

the Tafel slope from 131 mV/dec (MoS2) to 57 mV/dec (MoS0.60P0.70), indicating that the HER

(13)

stability with 20000 cylces clearly show that the P-alloyed MoS2 particles could endure the

acidic nature of measurement system, while the pristine MoS2 were degraded. Based on these

results, the P-alloyed MoS2 particles could act as an efficient catalyst with high acidic stability.

Kwon et al. [62] also synthesized sulfur-doped molybdenum phosphide (S:MoP) thin film

through simple thermolysis method by using [(NH4)2MoS4] solution and powder precursors with

different sulfur/phosphorus weight ratios. The S:MoP thin film catalyst, which is transferred on

p-Si substrates, showed a high photocurrent density (33.13 mA/cm2 at 0 V vs. RHE), large

overpotential shift (0.9 V at 10 mA/cm2) and long-term stability. Similar to above mentioned

report, DFT calculations indicated that the surface of S:MoP thin film absorb hydrogen better

than that of MoS2 thin film. Cabán-Acevedo et al. [63] studied pyrite-type cobalt

phosphosulphide (CoPS) as a HER catalyst. According to their DFT calculations, ΔGH can be

controlled by tuning the anion constituents of sulfur and phosphorus. After spontaneous H

adsorption at opened P sites, the ΔGH of the adjacent Co sites becomes very close to that of Pt.

Nanostructured CoPS only needed an overpotential of 45mV to achieve 10 mA/cm2 with long

term stability. The CoPS/n+pp+ Si micropyramids photocathode achieved photocurrent density up

to 35 mA/cm2 at 0 V vs. RHE.

The anions doping in MoS2 or other TMDs could be the one of the possible approach to

improve the catalytic inactivity of MoS2 in basal planes. The substituted anions facilitate the

decrease of the hydrogen adsorption Gibbs free energy in the neighboring S atoms in the MoS2

lattice by filling previously unoccupied electronic states within the substrate. The strong H

binding in the substitutional anion doped structure might be useful to improve the catalytic

(14)

2.3. By cation doping

Two-dimensional (2D) MoS2 nanomaterials have exhibited potential, but catalytic activity

only originates from their limited edge sites with massive, useless in-plane domains. It is great

interest to explore strategies to stimulate the catalytic activity of inert region which is in-plane

atoms of 2D hexagonal structure of MoS2, but it still remains as a great challenge due to the lack

of efficient experimental techniques and theoretical approaches [42, 64]. From the previous

technological approaches in graphene, the electronic properties of the carbon atoms neighboring

the doped heteroatoms can be significantly modulated thereby altering the catalytic activity of

in-plane carbon atoms [65-67]. To facilitate the useless in-in-plane domains for the catalysts, the

single-atom replacing ‘Mo’ atoms in MoS2 is one of the efficient way to enhance its catalytic

activity. The doped metal atoms could tune the adsorption behavior of H atoms on the

neighboring S sites. From this system, a significance enhanced HER activity on the MoS2 surface

could be anticipated by density-functional theory (DFT) calculations in recent reports.

Deng et al. [68] have demonstrated that the catalytic activity of in-plane S atoms of MoS2

can be triggered via single-atom metal (Pt, Ni, Co) substitutional doping into the Mo site. They

used a one-pot chemical reaction to synthesize the single Pt atom-doped, few-layer MoS2

nanosheets by using (NH4)6Mo7O24, H2PtCl6 and CS2 as precursors. From sub-angstrom

resolution high-angle annular dark field-scanning transmission electron microscopy

(HAADF-STEM) images, the single Pt atoms uniformly dispersed in the 2D MoS2 nanosheets were

observed as shown in Fig. 4A. The single Pt atoms could exactly occupy the positions of the Mo

atoms. The DFT calculations were carried out to characterize the catalytic process in Pt-MoS2.

Based on the optimized DFT calculations results, the in-plane S atoms, in the presence of Pt

(15)

and thereby the introduction of Pt atoms into the MoS2 can significantly enhance the HER

activity, as shown in Fig. 4B. Furthermore, as predicted in the volcano plot for varied

single-atom metal-doped MoS2, the non-noble-metal dopants such as Co and Ni could also be the good

alternative of Pt to tune the catalytic activity of MoS2. The HER polarization curves for the

Ni-MoS2, Co-MoS2, and Pt-MoS2 is displayed in Fig. 4C. In the case of Co-MoS2 nanosheets, the

obvious enhanced HER activities are observed by reducing the overpotential of about 20 mV

compared to flower-like MoS2 at the current density of 10 mA/cm2. The trend of HER activity in

these substitutional dopants is Pt > Co > Ni, which is good agreement with the calculated in

volcano plot as shown in Fig. 4B.

In MoS2, the Mo-edge has been demonstrated to be active sites for HER whereas the S-edge

is inert. Facilitating the catalytic inert S-edge sites is the one of challenging issues to improve the

catalytic activity of MoS2. In the recent report, Wang et al. [69] have predicted that the S-edge

site could also become HER active sites by incorporating transition metal atoms (Fe, Co, Ni, or

Cu) on the vertically standing, edge-terminated MoS2 nanofilms. From the DFT calculations, all

doped S-edge sites display the closer hydrogen adsorption Gibbs free energy value to

thermos-neutral than the that of S-edge of pristine MoS2 layer. Doping with Co, Fe, Ni, and Cu could

elevate the catalytic activity of S-edge sites to make it comparable to or better than pristine

Mo-edge sites. As shown in Fig. 4D, the transition metal-doped MoS2 thin films could be synthesized

by sulfurization process of the thin transition metal film/Mo film layer. In the kinetics controlled

synthesis condition, the Mo film changed to vertically aligned structure with many catalytic

active sites on its surface. The thin transition metal film could also be substituted into the top

layer of the MoS2 layer. The schematic illustration of synthesized nanofilms is shown in Fig. 4D.

(16)

pristine and Co-doped MoS2 nanofilms. No particles or clusters are also observed, indicating that

the Co atoms are uniformly distributed on the surface edge sites without formation of cobalt

sulfide. The electrochemical characterization of the synthesized nanofilms are displayed in Fig.

4H. The pristine MoS2 nanofilm reaches 0.6 mA/cm2 at 300 mV overpotential, while the Fe, Co,

Ni and Cu-doped MoS2 nanofilms achieve 2.3, 3.5, 2.4, and 2.6 mA/cm2 at 300 mV,

respectively. The catalytic activity was enhanced about 4 times compared to the pristine one.

Although the current density value at the 300 mV is drastically improved compared to the

pristine one, the value of Tafel slope is comparable with each other, regardless of kinds of

dopants. The unaffected Tafel slopes suggested that the rate-determining step is not changed

after the doping process, and thus the comparison of the exchange current densities as well as the

turn over frequency per active site is meaningful. They calculated the exchange current densities

and the turn over frequencies of the synthesized materials. Based on the theoretical and

experimental trends, the order of catalytic activity should be Fe-doped MoS2 > Co-doped MoS2 >

pristine MoS2, Cu-doped MoS2 > Ni-doped MoS2. This results paved the way to improve the

electrochemical catalysts by incorporating promoters into particular atomic sites, and for using

well-defined systems in order to understand the origin of the promotion effects.

The catalytic activity of materials is mainly decided by its structure, the number of edge

sites, and elements, but there are some promising way to improve the catalytic activity by using

various approaches. The substitutional doping method is one of the promising technique with

changing the catalytic inactive element to the other catalytic active element.

2.4 By phase transformation

2D TMDs has diverse polymorphs. Above all, 2H-TMDs mostly exhibit semiconducting

(17)

maximizing the exposure of edge sites to improve HER performance. However, 2H-TMDs have

one drawback to improve their performance due to the limited number of active edge sites. It has

been recently demonstrated that phase transition of semiconducting 2H-TMDs into metallic

1T-TMDs could enhance the HER performance [34]. Joensen et al. [33] showed that single layer

MoS2 nanosheets could be exfoliated to stable colloidal suspension by Li intercalation. Moreover,

chemical exfoliation via Li intercalation leads to transformations of the crystal structure owing to

the electron transfer from Li compounds to MoS2 nanosheets. According to the calculation work

[70], octahedral coordination is desirable in order to accept these surplus electrons in the d

orbitals of the transition metal. The metallic property of 1T-TMDs has come of these results.

Ambrosi et al. [71] explored the electrocatalytic activities of MoS2, MoSe2, WS2, and WSe2

for HER using similar exfoliation method of Li intercalation. They showed that the degree of

phase transformation and exfoliation are apparently different for the four TMD materials. The

differences in catalytic activities originated in the portion of the 1T phase which exist in the

exfoliated TMDs. Among them, WS2 was the most efficient catalyst for HER as the WS2

nanosheets exhibited the lowest overpotential. Lukowski et al. [72] demonstrated the facile

procedure to obtain 1T-TMD (1T-WS2) through a microwave-assisted reaction. The exfoliated

1T-WS2 nanosheets showed rapid kinetics, metallic property and increased number of active sites,

which enable a dramatic improvement in HER performance compared to 2H-WS2.

To elucidate the HER mechanism of two different phase MoS2 (2H and 1T), Voiry et al. [73]

partially oxidized the 2H-MoS2 and 1T-MoS2 nanosheets over few days in aqueous solution

which was saturated by oxygen. The oxidation was identified by the confirmation of several new

XPS peaks. The partial oxidation of the edges was also confirmed by TEM (see Fig. 5A) that the

edges of MoS2 nanosheets are very disordered after oxidation process, preserving the basal plane

(18)

showed large overpotential over 250 mV and low current densities due to the limited number of

edge site. On the other hand, chemically exfoliated 1T-MoS2 nanosheets showed significantly

enhanced electrochemical activity with reduced overpotentials of ~ 100 mV and low Tafel slopes

of ~ 40mV/dec (iR-corrected). After oxidation, HER activity of 2H-MoS2 nanosheets was

drastically deteriorated, while 1T-MoS2 exhibited thoroughly unaffected HER activity as shown

in Fig. 5B and C. 1T-MoS2 nanosheets showed no change in HER performance regardless of

edge oxidation, suggesting that the primary active site of 1T-MoS2 is located in the basal plane.

Furthermore, the metallic edge sites have quite smaller contribution on the HER in contrast with

2H-MoS2 nanosheets. Wang et al. [74] investigated the d-spacing, oxidation state and the ratio of

phase transformation of MoS2 catalysts using the battery testing system (see Fig. 5D). They

found that when the phase transformation from 2H to 1T takes place, the electrocatalytic

properties are already enhanced as the oxidation state of Mo is lowered. The decreased oxidation

states of Mo might affect the electronic structure which dramatically changes the hydrogen

adsorption free energy, ∆GH, and reduce the activation energy, resulting in the improvement of

HER performance. As the applied voltage vs. Li+

/Li is continuously reduced from 2.1 V to 1.1 V,

the first-order phase transition of 2H-MoS2 to 1T-MoS2 occurred between 1.2 V and 1.1 V. This

explains the superior HER activity of 1.1 V lithiated MoS2 as shown in Fig. 5F which well

accords with other recent reports [34, 75]. Gao et al. [40, 76] investigated the phase transition of

monolayer MoS2 from 2H to 1T and to 1T′ (distorted 1T) by density functional theory (see Fig.

5G). The atomic and electronic structures of 2H-, 1T- and 1T′-MoS2 are illustrated in Fig. 5H.

They found that the phase transition of MoS2 monolayer is seriously affected by the negative

charge. In general, the 1T phase MoS2 is not stable and easily transform to 2H-MoS2 or 1T′-MoS2.

The negative charge could thermodynamically stabilize 1T-MoS2 in comparison with the

(19)

eV when 4 electrons are supplied per MoS2 unit. At the same time, 1T-MoS2 is transformed to

1T′-MoS2. According to calculation (see Fig. 5I), the hydrogen adsorption free energy, ∆GH, on

(1010) Mo-edge of 1T′-MoS2 is as low as the (1010) Mo-edge of 2H-MoS2. This result suggest

that the electrocatalytic activity of 1T′-MoS2 exhibits comparable to the 2H-MoS2. Considering

the charge transfer, the 1T′-MoS2 would exhibit higher catalytic activity than the 2H-MoS2. Tang

et al. [77] also conducted DFT calculations to demonstrate the HER mechanism of the basal

planes of metallic 1T-MoS2 nanosheets. Compared to the catalytically inert basal plane of

2H-MoS2, the superior catalytic activity of the basal plane of 1T-MoS2 originated in its affinity for H

at the S sites on the surface. They found that the optimum HER on the basal plane of 1T-MoS2

will occur at the surface H coverage of 12.5 % ~ 25 %. Within this coverage, their results

suggest that HER could mainly occur on the basal plane of 1T-MoS2 through a combination of

the Volmer-Heyrovsky mechanism rather than the Volmer-Tafel mechanism. Moreover, they

discovered that the electrocatalytic activity of 1T-MoS2 can be further improved by substituting

other metal atoms (e.g., Mn, Cr, Cu, Ni, Fe) for Mo sites in the MoS2 lattice.

2.5 By defects generation

The catalytically active edge sites of 2H-TMDs are attributed to the unsaturated sulfur sites

of the edge. This resulted in the fabrication of amorphous TMD in which a myriad of sulfur was

unsaturated [40, 78]. Active edge sites could be also increased through the structural control of

the TMD nanosheets [79, 80]. Xie et al. [79] demonstrated defect-rich MoS2 nanosheets

synthesized by using high concentrated precursors and controlling the amounts of thiourea. The

defect-rich structure is formed when thiourea is used in excess, because excess thiourea adsorbs

(20)

defects resulting from the partial cleavage of the catalytically inactive basal planes leads to the

creation of extra active edge sites. A corresponding HRTEM images of top and side view are

displayed in Fig. 6B. The top-view HRTEM image shows the d-spacing of 2.7 Å which well

coincides with the d-spacing of the 2H-MoS2 (001) planes. Through the side view of HRTEM

image, numerous dislocations and distortions of defect-rich structures could be recognized. As

shown in Fig. 6C, the defect-rich MoS2 nanosheets show high HER performance with low onset

potential of 120 mV and the low Tafel slope of 50 mV/dec by virtue of the creation of extra

active edge sites. Ye et al. [81] also reported the structural defect-induced monolayer MoS2 via

oxygen plasma exposure and hydrogen annealing. The 2H-MoS2 monolayers are grown by CVD

method, similar to their previous report [82]. They exposed the 2H-MoS2 monolayers to oxygen

plasma for 10, 20, 30 s, reported as a successful method to generate defects into the monolayer

MoS2 [83-85]. After 10 s of oxygen plasma treatment, short and individual cracks could be

recognized as illustrated in Fig. 6D that begin to grow in the basal plane of 2H-MoS2

monolayers. The lowest Tafel slope of 171 mV/dec and the smallest onset potential (Fig. 6F) are

obtained from the sample which exposed to oxygen plasma for 20 s. Moreover, they also

intentionally generated structural defects through another approach using hydrogen annealing.

The hydrogen annealing has been utilized to create defects on graphene such as hexagonal holes

or fractal patterns [86, 87]. These analogous approach was applied to create additional active

sites by etching the catalytically inert basal plane of MoS2 monolayers. Unlike the case of

producing defects through oxygen plasma, the triangular holes created by hydrogen annealing

are non-uniform and omnipresent on basal plane as displayed in Fig. 6H. Samples treated with

hydrogen annealing apparently exhibit reduced onset potential and increased current density as

shown in Fig. 6G. Since the increase in current density is proportional to the exposed edge length

(21)

plasma method in improving the catalytic activity of MoS2 monolayers. Yin et al. [88] proposed

a novel hydrothermal method which simultaneously control crystal phase and disorder of

partially crystallized 1T-MoSe2 nanosheets. The phase transition of 2H-phase to 1T-phase, using

an excess content of reductant (NaBH4), enhanced the intrinsic conductivity and HER activity.

Moreover, a lower reaction temperature made disordered structure of abundant unsaturated

defects which play a role as active sites. These two synergistic factors led to an outstanding HER

performance with a low onset potential of 120 mV to exhibit – 10 mA/cm2 and the low Tafel

slope of 52 mV/dec.

As mentioned above, most approaches focus on optimizing the edge sites by generating

structural defects. Meanwhile, there are a few approaches that activate the basal planes which

occupy the majority of 2H-MoS2. Li et al. [89] reported that the inert basal plane of 2H-MoS2

monolayer could be activated and optimized by generating S-vacancies and exerting strains as

illustrated in Fig. 6H. They first used DFT calculations to determine the hydrogen adsorption

free energy, ∆GH. The maximum catalytic activity can be expected when the ∆GH approaches 0 eV

[90]. This is because the lower ∆GH let the atomic hydrogen more strongly bound, making it

difficult to desorb molecular hydrogen, while higher ∆GH hinder the adsorption of atomic

hydrogen on the catalyst. For basal planes of pristine 2H-MoS2, ∆GH is around 2 eV, which

exhibits an inert property. As S-vacancies are introduced, the atomic hydrogen is stabilized on

the S-vacancies and results in the decrease of ∆GH as shown in Fig. 6I. Then, they explored the

effect of the strained 2H-MoS2. Regardless of the S-vacancies concentration, the elastic strains

lead to more negative ∆GH by strengthening hydrogen adsorption [75, 91]. The decrease of ∆GH by

introducing S-vacancies and elastic strains can be explained through investigating its electronic

structure. As S-vacancies are generated, new electronic bands are created in the gap close to the

(22)

closer to the Fermi level. In order to identify the results from DFT calculations, they synthesized

monolayer 2H-MoS2. Then, tensile strain was applied by capillary force using a patterned Au

nanocone and the S-vacancies were introduced on basal plane of 2H-MoS2 through exposure to

Ar plasma. Fig. 6J displays linear sweep voltammetry (LSV) curves of Au electrode, Pt electrode,

Pristine MoS2 without strain and S-vacancy, strained MoS2 without S-vacancy (S-MoS2),

S-vacant MoS2 without strain (V-MoS2), and strained MoS2 with S-vacancies (SV-MoS2). Firstly,

the Au electrode exhibits little HER activity and the Pt electrode has the highest HER activity as

in previous studies [34, 73, 92-94]. The pristine MoS2 has very low current density which well

coincides with the nature of inert basal plane. The HER activity is slightly increased for S-MoS2

while the HER activity is considerably increased for V-MoS2 in accordance with the prediction

for ∆GH. Moreover, the SV-MoS2 exhibits much higher current density at fixed potential. These

strategies of creating S-vacancies with elastic strain can be a universal post-treatment method for

other 2H-TMD catalysts that could maximize catalytic activity by activating inert basal planes of

2H-TMDs. Cheng et el. [95] also created S-vacancies to activate the basal plane of MoS2

monolayer using remote hydrogen plasma. They experimentally and theoretically examined the

effect of S-vacancies on the basal planes of the MoS2 monolayer. They demonstrated that the

HER activity of the MoS2 monolayer can be controlled by tuning the period of hydrogen plasma.

By using hydrogen plasma, the stoichiometry of the MoS2 monolayer can be controlled from

MoS1.97 to MoS1.20, providing relatively wide range of Mo-S atomic ratios with preserving the

original basal plane of MoS2 monolayer. Yin et al. [96] fully explored the catalytic nature of

MoS2 for HER in terms of the phase, edge exposure, and sulfur vacancies. They used a series of

representative samples including 2H and 1T phase, porous 2H and 1T phase, and

sulfur-compensated 2H phase MoS2 nanosheets. Among them, they reported that the excellent HER

(23)

and S-vacancies than pristine 1T-MoS2. This result is not only because the crystal phase, as

mentioned in Chapter 2.4, plays a major role in HER properties, but also because the synergistic

effects of both increased edge sites and S-vacancies contributed greatly to the catalytic activity of

porous 1T-MoS2 nanosheets.

2.6 By heterojunctions

Heterojunction is one of the promising approaches to make the best use of TMDs as

electrocatalyst. Xiang et al. [97] synthesized TiO2/MoS2/graphene heterojunctions (see Fig. 7A)

via two-step hydrothermal process. Fig. 7B shows TEM images of the synthesized

TiO2/MoS2/graphene heterostructure composite, which represents uniform distribution of TiO2

nanoparticles on MoS2/graphene support. The photocatalytic activity of TiO2 nanoparticles

without any support exhibited a negligible photocatalytic activity due to the fast recombination

of photogenerated electron-hole pairs. When the TiO2 nanoparticle make heterojunction with

MoS2, the TiO2/MoS2 composite showed a modest photocatalytic hydrogen production rate of

36.8 µmol hmol h−1. This relatively low photocatalytic activity is attributed to the mismatched

heterostructure of TiO2/MoS2 that form Schottky barrier at the interface [98], hindering the

transfer of photogenerated charge. With introduction of graphene, the photocatalytic hydrogen

production activity was significantly enhanced to 165.3 µmol hmol h−1

as shown in Fig. 7C. Since the

redox potential of graphene is located between the conduction band minimum of TiO2 and H+ /H2

potential, the photogenerated electrons in the TiO2 nanoparticles could be easily transferred to

MoS2 nanosheets through the conductive graphene “highway”. Then, the transferred electrons

could efficiently reduce H+

ions to H2 at the active edge site of MoS2 nanosheets. This synergetic

effect of TiO2/MoS2/graphene heterostructure is illustrated in Fig. 7A. Zong et al. [99] reported

(24)

The MoS2/CdS heterostructure showed even higher rate of hydrogen evolution than Pt/CdS

heterostructure. This result originates in the structural and electronic coupling of MoS2/CdS

heterostructure. Chang et al. [100] demonstrated highly efficient CdS/MoS2/graphene

heterostructure photocatalyst. The introduction of graphene prevents the aggregation of MoS2,

resulting in the few layer MoS2/graphene heterostructure. Furthermore, the incorporation of

graphene enable the rapid charge transfer of electrons which were generated from CdS. Thus, the

photogenerated electrons could be directly transferred from CdS to the edge sites of MoS2 or via

graphene to reduce H+

ion to H2. The schematic illustration of CdS/MoS2/graphene

heterostructure is shown in Fig. 7E and the corresponding HRTEM image is displayed in Figure

7D. The photocatalytic activity of CdS/MoS2/graphene heterostructure was greatly enhanced in

lactic acid solution. As shown in Fig. 7F, the hydrogen production rate of the

CdS/MoS2/graphene heterostructure system reached 2.0mmol h -1

which was higher value than

that of Pt/CdS heterostructure system (0.3 mmol h-1

). This result suggests that the charge transfer

is enhanced and the recombination of photogenerated electron-hole pairs is suppressed with

(25)

3. Conclusion and perspectives

Although the 2D TMDs have been extensively focused as a catalyst for evolving hydrogen,

the catalytic activity of 2D TMDs are still not enough to replace the Pt catalyst because of its

catalytic inactive in their basal planes. Both computational and experimental studies revealed

that the edge sites of 2H-TMDs exhibit the good catalytic activity for the HER, while the basal

planes were catalytically inert due to weak hydrogen binding capability. In this review, we have

categorized the tailoring methods to improve the catalytic activity in 2D TMDs based on their

dimensionalities and atomic structures. We have summarized the major approaches to improve

the catalytic activity of 2D TMDs in EC and PEC catalysis including i) core-shell nanostructures,

ii) substitutional anion and cation doping, iii) phase transformation from 2H to 1T, iv) effective

sulfur vacancies and the strain on its surface, and v) the heterostructure with other materials.

These tailoring methods have been reported to exhibit the enhanced catalytic activity in HER,

but the several issues still remained for applying these techniques to the practical application. To

increase the likelihood of industrialization, not laboratory-level performance, the TMD-based

catalysts have to be synthesized in the largescale via cost-competitive method. Also, the

long-term stability in actual application should be ensured. To date, the research has focused only on

edge sites and defects as active sites for 2H-TMDs.

One approach of our suggestions to achieve the highly efficient catalysts is heterojunction

between atomically thin TMD films. This approach using atomically thin TMD pn junctions to

reduce the overpotential of HER provides a very innovative way for realizing a highly efficient

and stable PEC water splitting because it allows us to overcome the fundamental limitations of

the conventional 2H-TMDs. According to the idealized volcano curve shown in Fig. 8, the

exchange current density of existing catalysts is determined intrinsically by the Gibbs free

(26)

each other to make a junction, carriers diffuse to reach the equilibrium and thus space charge

(depletion) region spatially confined in the 2D atomic planes of the junction is formed. Since the

depletion width is atomically thin, the induced electrical field would be extremely strong. Such a

field significantly promotes the transfer of injected carriers from p-Si into water. In other words,

the TMD pn junction can act as an interface dipole layer to preferentially adjust energy level

alignment at the electrode/electrolyte interface, as analogous to the n+p Si which has internal

built-in potential to modify the electrode/electrolyte junction. That is, an atomically thin TMD pn

junction plays a role as an electrostatic catalyst, which allows us to overcome the limitations of

the catalytically inert basal plains of TMDs, enabling the realization of dipole-induced

self-activating water splitting by TMD pn junctions. In addition, the TMD pn junctions should enable

the ultrafast exciton dissociation [101, 102], which helps transfer and separation of

photogenerated carriers injected from underlying light absorbing semiconductor without notable

losses. Therefore, the tailored TMD monolayer (ML) pn junctions can be highly efficient HER

catalysts. Furthermore, since diverse TMD materials with sizable band gaps and work functions

are available, there are a large number of untested combinations to form pn junctions. In order to

realize this TMD pn junction electrocatalyst, extensive studies in various combinations of TMD

and their alloys should be performed to identify a clear descriptor for HER, rather than being

(27)

Acknowledgements

This work was supported by the Samsung Research Funding Center of Samsung Electronics.

Ki Chang Kwon acknowledges the Global Ph. D Fellowship Program of the National Research

Foundation of Korea funded by the Ministry of Education.

References

[1] T. Hisatomi, J. Kubota, K. Domen, Chem. Soc. Rev. 43 (2014) 7520-7535.

[2] M.R. Gholipour, C.-T. Dinh, F. Béland, T.-O. Do, Nanoscale 7 (2015) 8187-8208. [3] J.M. Bockris, Int. J. Hydrogen Energy 27 (2002) 731-740.

[4] J.R. Rostrup•Nielsen, Catalysis reviews 46 (2004) 247Nielsen, Catalysis reviews 46 (2004) 247-270.

[5] J. Greeley, T.F. Jaramillo, J. Bonde, I. Chorkendorff, J.K. Nørskov, Nat. Mater. 5 (2006) 909-913.

[6] D. Voiry, J. Yang, M. Chhowalla, Adv. Mater. 28 (2016) 6197-6206.

[7] M.G. Walter, E.L. Warren, J.R. McKone, S.W. Boettcher, Q. Mi, E.A. Santori, N.S. Lewis, Chem. Rev. 110 (2010) 6446-6473.

[8] K. Sivula, R. Van De Krol, Nature Reviews Materials 1 (2016) 15010.

[9] D.M. Andoshe, J.-M. Jeon, S.Y. Kim, H.W. Jang, Electron. Mater. Lett. 11 (2015) 323-335. [10] R. De Levie, J. Electroanal. Chem. 476 (1999) 92-93.

[11] M. Grätzel, Nature 414 (2001) 338-344.

[12] S. Cobo, J. Heidkamp, P.-A. Jacques, J. Fize, V. Fourmond, L. Guetaz, B. Jousselme, V. Ivanova, H. Dau, S. Palacin, M. Fontecave, V. Artero, Nat. Mater. 11 (2012) 802-807.

[13] L. Ji, M.D. McDaniel, S. Wang, A.B. Posadas, X. Li, H. Huang, J.C. Lee, A.A. Demkov, A.J. Bard, J.G. Ekerdt, E.T. Yu, Nat. nanotechnol. 10 (2015) 84-90.

[14] J. Kye, M. Shin, B. Lim, J.-W. Jang, I. Oh, S. Hwang, ACS nano 7 (2013) 6017-6023. [15] D.M. Andoshe, S. Choi, Y.-S. Shim, S.H. Lee, Y. Kim, C.W. Moon, D.H. Kim, S.Y. Lee, T. Kim, H.K. Park, M.G. Lee, J.-M. Jeon, K.T. Nam, M. Kim, J.K. Kim, J. Oh, H.W. Jang, J. Mater. Chem. A 4 (2016) 9477-9485.

[16] Q. Peng, B. Kalanyan, P.G. Hoertz, A. Miller, D.H. Kim, K. Hanson, L. Alibabaei, J. Liu, T.J. Meyer, G.N. Parsons, J.T. Glass, Nano Lett. 13 (2013) 1481-1488.

[17] M. Gong, Y. Li, H. Wang, Y. Liang, J.Z. Wu, J. Zhou, J. Wang, T. Regier, F. Wei, H. Dai, J. Am. Chem. Soc 135 (2013) 8452-8455.

[18] D. Friebel, M.W. Louie, M. Bajdich, K.E. Sanwald, Y. Cai, A.M. Wise, M.-J. Cheng, D. Sokaras, T.-C. Weng, R. Alonso-Mori, R.C. Davis, J.R. Bargar, J.K. Nørskov, A. Nilsson, A.T. Bell, J. Am. Chem. Soc. 137 (2015) 1305-1313.

[19] Q. Xiang, J. Yu, M. Jaroniec, Chem. Soc. Rev. 41 (2012) 782-796.

[20] J. Luo, J.-H. Im, M.T. Mayer, M. Schreier, M.K. Nazeeruddin, N.-G. Park, S.D. Tilley, H.J. Fan, M. Grätzel, Science 345 (2014) 1593-1596.

[21] Y. Hou, M.R. Lohe, J. Zhang, S. Liu, X. Zhuang, X. Feng, Energy Environ. Sci. 9 (2016) 478-483.

[22] Q. Ding, B. Song, P. Xu, S. Jin, Chem 1 (2016) 699-726.

(28)

[24] B. Peng, P.K. Ang, K.P. Loh, Nano Today 10 (2015) 128-137.

[25] B. Hinnemann, P.G. Moses, J. Bonde, K.P. Jørgensen, J.H. Nielsen, S. Horch, I. Chorkendorff, J.K. Nørskov, J. Am. Chem. Soc. 127 (2005) 5308-5309.

[26] F. Wang, Z. Wang, Q. Wang, F. Wang, L. Yin, K. Xu, Y. Huang, J. He, Nanotechnology 26

[29] Z. Zeng, C. Tan, X. Huang, S. Bao, H. Zhang, Energy Environ. Sci. 7 (2014) 797-803. [30] H.L. Zhuang, R.G. Hennig, J. Phys. Chem. C 117 (2013) 20440-20445.

[31] A.B. Laursen, S. Kegnæs, S. Dahl, I. Chorkendorff, Energy Environ. Sci. 5 (2012) 5577-5591.

[32] T.F. Jaramillo, K.P. Jørgensen, J. Bonde, J.H. Nielsen, S. Horch, I. Chorkendorff, Science 317 (2007) 100-102.

[33] P. Joensen, R. Frindt, S.R. Morrison, Mater. Res. Bull. 21 (1986) 457-461.

[34] M.A. Lukowski, A.S. Daniel, F. Meng, A. Forticaux, L. Li, S. Jin, J. Am. Chem. Soc. 135 (2013) 10274-10277.

[35] J.N. Coleman, M. Lotya, A. O’Neill, S.D. Bergin, P.J. King, U. Khan, K. Young, A. Gaucher, S. De, R.J. Smith, I.V. Shvets, S.K. Arora, G. Stanton, H.-Y. Kim, K. lee, G.T. Kim, G.S. Duesberg, T. Hallam, J.J. Boland, J.J. Wang, J.F. Donegan, J.C. Grunlan, G. Moriaty, A. Shmeliov, R.J. Nicholls, J.M. Perkins, E.M. Grieveson, K. Theuwissen, D.W. McComb, P.D. Nellist, V. Nicolosi, Science 331 (2011) 568-571.

[36] J. Kibsgaard, Z. Chen, B.N. Reinecke, T.F. Jaramillo, Nat. Mater. 11 (2012) 963-969.

[37] H. Tributsch, J. Bennett, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 81 (1977) 97-111.

[38] J. Bonde, P.G. Moses, T.F. Jaramillo, J.K. Nørskov, I. Chorkendorff, Faraday Discuss. 140 (2009) 219-231.

[39] T.F. Jaramillo, J. Bonde, J. Zhang, B.-L. Ooi, K. Andersson, J. Ulstrup, I. Chorkendorff, J. Phys. Chem. C 112 (2008) 17492-17498.

[40] D. Merki, S. Fierro, H. Vrubel, X. Hu, Chemical Science 2 (2011) 1262-1267.

[41] Z. Chen, D. Cummins, B.N. Reinecke, E. Clark, M.K. Sunkara, T.F. Jaramillo, Nano Lett. 11 (2011) 4168-4175.

[42] Y. Li, H. Wang, L. Xie, Y. Liang, G. Hong, H. Dai, J. Am. Chem. Soc. 133 (2011) 7296-7299.

[43] F. Meng, J. Li, S.K. Cushing, M. Zhi, N. Wu, J. Am. Chem. Soc. 135 (2013) 10286-10289. [44] Q. Ding, F. Meng, C.R. English, M. Cabán-Acevedo, M.J. Shearer, D. Liang, A.S. Daniel, R.J. Hamers, S. Jin, J. Am. Chem. Soc. 136 (2014) 8504-8507.

[45] H. Liang, H. Shi, D. Zhang, F. Ming, R. Wang, J. Zhuo, Z. Wang, Chem. Mater. 28 (2016) 5587-5591.

[46] D. Kong, H. Wang, J.J. Cha, M. Pasta, K.J. Koski, J. Yao, Y. Cui, Nano Lett. 13 (2013) 1341-1347.

[47] S. Trasatti, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 39 (1972) 163-184.

(29)

[49] X. Zhang, F. Meng, S. Mao, Q. Ding, M.J. Shearer, M.S. Faber, J. Chen, R.J. Hamers, S. Jin, Energy Environ. Sci. 8 (2015) 862-868.

[50] Q. Ding, J. Zhai, M. Cabán•Acevedo, M.J. Shearer, L. Li, H.C. Chang, M.L. Tsai, D. Ma, Acevedo, M.J. Shearer, L. Li, H.C. Chang, M.L. Tsai, D. Ma, X. Zhang, R.J. Hamers, J.-H. He, S. Jin, Adv. Mater. 27 (2015) 6511-6518.

[51] S.W. Boettcher, E.L. Warren, M.C. Putnam, E.A. Santori, D. Turner-Evans, M.D. Kelzenberg, M.G. Walter, J.R. McKone, B.S. Brunschwig, H.A. Atwater, N.S. Lewis, J. Am. Chem. Soc. 133 (2011) 1216-1219.

[52] D. Bae, T. Pedersen, B. Seger, M. Malizia, A. Kuznetsov, O. Hansen, I. Chorkendorff, P.C. Vesborg, Energy Environ. Sci. 8 (2015) 650-660.

[53] P. Xiao, M.A. Sk, L. Thia, X. Ge, R.J. Lim, J.-Y. Wang, K.H. Lim, X. Wang, Energy Environ. Sci. 7 (2014) 2624-2629.

[54] J. Kibsgaard, T.F. Jaramillo, Angew. Chem. Int. Ed. 53 (2014) 14433-14437. [55] Z. Xing, Q. Liu, A.M. Asiri, X. Sun, Adv. Mater. 26 (2014) 5702-5707.

[56] R. Ye, P. del Angel•Vicente, Y. Liu, M.J. Arellano•Jimenez, Z. Peng, T. Wang, Y. Li, B.I. Vicente, Y. Liu, M.J. Arellano•Jimenez, Z. Peng, T. Wang, Y. Li, B.I. Yakobson, S.H. Wei, M.J. Yacaman, J.M. Tour, Adv. Mater. 28 (2015) 1427-1432.

[57] P.E. Blöchl, Physical review B 50 (1994) 17953.

[58] G. Kresse, D. Joubert, Physical Review B 59 (1999) 1758.

[59] J.P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett. 77 (1996) 3865. [60] G. Kresse, J. Hafner, Physical Review B 47 (1993) 558.

[61] G. Kresse, J. Furthmüller, Physical review B 54 (1996) 11169.

[62] K.C. Kwon, S. Choi, J. Lee, K. Hong, W. Sohn, D.M. Andoshe, K.S. Choi, Y. Kim, S. Han, S.Y. Kim, H.W. Jang, J. Mater. Chem. A (2017).

[63] M. Cabán-Acevedo, M.L. Stone, J. Schmidt, J.G. Thomas, Q. Ding, H.-C. Chang, M.-L. Tsai, J.-H. He, S. Jin, Nat. Mater. 14 (2015) 1245-1251.

[64] X. Huang, Z. Zeng, S. Bao, M. Wang, X. Qi, Z. Fan, H. Zhang, Nat. Commun. 4 (2013)

[70] M. Kertesz, R. Hoffmann, J. Am. Chem. Soc. 106 (1984) 3453-3460. [71] A. Ambrosi, Z. Sofer, M. Pumera, Chem. Commun. 51 (2015) 8450-8453.

[72] M.A. Lukowski, A.S. Daniel, C.R. English, F. Meng, A. Forticaux, R.J. Hamers, S. Jin, Energy Environ. Sci. 7 (2014) 2608-2613.

[73] D. Voiry, M. Salehi, R. Silva, T. Fujita, M. Chen, T. Asefa, V.B. Shenoy, G. Eda, M. Chhowalla, Nano Lett. 13 (2013) 6222-6227.

[74] H. Wang, Z. Lu, S. Xu, D. Kong, J.J. Cha, G. Zheng, P.-C. Hsu, K. Yan, D. Bradshaw, F.B. Prinz, Y. Cui, Proceedings of the National Academy of Sciences 110 (2013) 19701-19706. [75] D. Voiry, H. Yamaguchi, J. Li, R. Silva, D.C. Alves, T. Fujita, M. Chen, T. Asefa, V.B. Shenoy, G. Eda, M. Chhowalla, Nat. Mater. 12 (2013) 850-855.

[76] G. Gao, Y. Jiao, F. Ma, Y. Jiao, E. Waclawik, A. Du, J. Phys. Chem. C 119 (2015) 13124-13128.

(30)

[78] H. Vrubel, D. Merki, X. Hu, Energy Environ. Sci. 5 (2012) 6136-6144.

[79] J. Xie, H. Zhang, S. Li, R. Wang, X. Sun, M. Zhou, J. Zhou, X.W.D. Lou, Y. Xie, Adv. H. Terrones, M. Terrones, B.K. Tay, J. Lou, S.T. Pantelides, Z. Liu, W. Zhou, P.M. Ajayan, Nat. Mater. 13 (2014) 1135-1142.

[83] A. Azizi, X. Zou, P. Ercius, Z. Zhang, A.L. Elías, N. Perea-López, G. Stone, M. Terrones, B.I. Yakobson, N. Alem, Nat. Commun. 5 (2014).

[84] N. Kang, H.P. Paudel, M.N. Leuenberger, L. Tetard, S.I. Khondaker, J. Phys. Chem. C 118 (2014) 21258-21263. H.C. Manoharan, F. Abild-Pedersen, J.K. Nørskov, X. Zheng, Nat. Mater. 15 (2016) 48-53. [90] J.K. Nørskov, T. Bligaard, A. Logadottir, J. Kitchin, J.G. Chen, S. Pandelov, U. Stimming, J. Electrochem. Soc. 152 (2005) J23-J26.

[91] P. Strasser, S. Koh, T. Anniyev, J. Greeley, K. More, C. Yu, Z. Liu, S. Kaya, D. Nordlund, H. Ogasawara, M.F. Toney, A. Nilsson, Nature chemistry 2 (2010) 454-460.

[92] D. Gopalakrishnan, D. Damien, M.M. Shaijumon, ACS nano 8 (2014) 5297-5303.

[93] D.J. Li, U.N. Maiti, J. Lim, D.S. Choi, W.J. Lee, Y. Oh, G.Y. Lee, S.O. Kim, Nano Lett. 14 (2014) 1228-1233.

[94] J.F. Callejas, J.M. McEnaney, C.G. Read, J.C. Crompton, A.J. Biacchi, E.J. Popczun, T.R. Gordon, N.S. Lewis, R.E. Schaak, ACS nano 8 (2014) 11101-11107.

[95] C.-C. Cheng, A.-Y. Lu, C.-C. Tseng, X. Yang, M.N. Hedhili, M.-C. Chen, K.-H. Wei, L.-J. Li, Nano Energy 30 (2016) 846-852.

[96] Y. Yin, J. Han, Y. Zhang, X. Zhang, P. Xu, Q. Yuan, L. Samad, X. Wang, Y. Wang, Z. Zhang, P. Zhang, X. Cao, B. Song, S. Jin, J. Am. Chem. Soc. 138 (2016) 7965-7972.

[97] Q. Xiang, J. Yu, M. Jaroniec, J. Am. Chem. Soc. 134 (2012) 6575-6578. [98] K. Chang, X. Hai, J. Ye, Adv. Energy Mater. 6 (2016) 1502555.

[99] X. Zong, H. Yan, G. Wu, G. Ma, F. Wen, L. Wang, C. Li, J. Am. Chem. Soc. 130 (2008) 7176-7177.

[100] K. Chang, Z. Mei, T. Wang, Q. Kang, S. Ouyang, J. Ye, ACS nano 8 (2014) 7078-7087. [101] C.-H. Lee, G.-H. Lee, A.M. Van Der Zande, W. Chen, Y. Li, M. Han, X. Cui, G. Arefe, C. Nuckolls, T.F. Heinz, J. Guo, J. Hone, P. Kim, Nat. nanotechnol. 9 (2014) 676-681.

(31)

Fig. 1. (A) Schematic of core-shell nanowires with substiochiometric MoO3 cores of 20–50 nm and conformal MoS2 shells of 2–5 nm for the PEC hydrogen generation. (B) External morphology of synthesized core (MoO3) – shell (MoS2) structure at 200 °C. The inset figure shows the 2 × 3 cm2

samples on the SiO2 substrate. (C) Internal morphology of nanowires sulfurized at 200 °C. The inset image shows the magnified TEM images. (D) The PEC performance of the nanowires sulfurized at 200 and 300 °C on the TEC15FTO glass [41]. Copyright © 2011, American Chemical Society. Schematic solvothermal synthesis (E) with GO sheets to afford the MoS2/RGO hybrid and (F) without any GO sheets, free MoS2 particles. (G) High-resolution TEM images of folded edges of MoS2 particles on RGO in the hybrid. (H) Tafel plots recorded on glassy carbon electrodes with a catalyst loading of 0.28 mg/cm3

(32)

Fig. 2. (A) Schematic of the synthesis setup in a horizontal tube furnace and the volume-rendered reconstructed TEM tomogram of a MoS2 film grown by rapid sulfurization, which resembles the ideal edge-terminated structure. The TEM image of (B) MoS2 and (C) MoSe2 thin film produced by rapid sulfurization, clearly showing exposed edges. (D) The polarization curves of edge-terminated MoS2 and MoSe2 films as well as a blank glassy carbon substrate showing H2 evolution [46]. Copyright © 2013, American Chemical Society. (E) The schematic diagram of simple thermolysis method by using thermal chemical vapor deposition system. (F) The TEM image of the synthesized MoS2 thin film with partially vertical stacked c-domains (001) and Moiré fringes. (G) PEC performances shown as polarization J–V curves of MoS2 films of

different thicknesses. The 13 nm-thick MoS2/p-Si photocathode shows 24.6 mA/cm2 at 0 V and a

significant potential value shift at 10 mA/cm2 from0.69 V for

p-Si to 0.082 V [48]. Copyright ©

(33)

Fig. 3. (A) Schematic of MoQxCly catalysts coated on n+ pp+

(34)

(35)

Fig. 5. (A) Edge-oxidized MoS2 nanosheets. High-resolution TEM and HAADF STEM of edge-oxidized MoS2 nanosheets showing corrugated edges caused by the chemical oxidization. Scale bars: 5 and 1 nm, respectively. And schematic of the oxidation process and partial restoration of the nanosheet edges after several voltammetric cycles. (B) Polarization curves of 1T and 2H MoS2 nanosheet electrodes before and after edge oxidation. iR-corrected polarization curves from 1T and 2H-MoS2 are shown by dashed lines. (C) Corresponding Tafel plots obtained from the polarization curves. Tafel slopes of ∼40 and 75−85 mV/dec have been measured for 1T and 2H MoS2, respectively. After oxidation, the Tafel slopes of 45 and 186 mV/dec for 1T and 2H-MoS2, respectively, were obtained [73]. Copyright © 2013, American Chemical Society. (D) Schematic of the battery testing system. The cathode is MoS2 nanofilm with molecular layer perpendicular to the substrate, where the green and yellow colors represent the edge sites and the terrace sites, respectively. The anode is the Li foil. (E) TEM image of MoS2 by Li electrochemical intercalation to 1.1 V vs. Li+

(36)

(37)

(38)

(39)

Highlights

Transition metal disulfides (TMDs) have received tremendous attention as hydroge n evolution reaction (HER) catalysts

Recent progress in tailoring catalytic activities of TMDs for electrochemical and p hotoelectrochemical hydrogen production is reviewed.

Obstacles for practical applicability are discussed to mention challenges to be mad e for realizing a catalyst comparable to Pt.

(40)

Gambar

Fig. 1. morphology of synthesized core (MoOshows the 2 × 3 cm (A) Schematic of core-shell nanowires with substiochiometric MoO3 cores of 20–50 nm and conformal MoS2 shells of 2–5 nm for the PEC hydrogen generation
Fig. 2. (A) Schematic of the synthesis setup in a horizontal tube furnace and the volume-resembles the ideal edge-terminated structure
Fig. 3. hydrogen generation. (B) MoS (A) Schematic of MoQxCly catalysts coated on n+pp+ Si micropyramids (MPs) for PEC xCly/Si MPs (top) and MoSexCly/Si MPs (bottom)
Fig. 4.  (A) The HAADF-STEM image of Pt-MoS2 nanosheet. The green, yellow, and blue balls represent Mo, S, and Pt, respectively, in the inset image
+5

Referensi

Dokumen terkait

Klarifikasi dan pembuktian kualifikasi wajib dihadiri oleh direktur atau personil yang diberikan kuasa dengan menunjukkan surat kuasa serta membawa seluruh dokumen

- Pihak lain yang bukan Direktur Utama/Pimpinan Perusahaan atau yang namanya tidak disebutkan dalam Akta Pendirian/ Anggaran Dasar, dapat menandatangani Berita Acara

[r]

[r]

[r]

KELOMPOK KERJA DISNAK PROV TAHUN 2012. Sekretariat

Penelitian ini untuk menguji aktivitas repelan dan mengetahui daya proteksi dari lotion ekstrak etanol kulit buah langsat terhadap nyamuk Aedes aegypti.. Ekstrak

Nila;loial HFS : Rp. Meftiliki srral tzin ti$ha rcrda:Bn:ta. {SII]P) K€cil F\g diterbitkBn oleh remcdn!,h.. kbnpatcr/tui.l iempat domisili p.n)-cdia barana dan nasih