Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Cell TMDs in LiNa Ion Battery

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Review

Design Strategies for Development


of TMD-Based Heterostructures
in Electrochemical Energy Systems
P. Prabhu,1 Vishal Jose,1 and Jong-Min Lee1,*

Transition metal dichalcogenides (TMDs) are promising materials for use in elec- Progress and Potential
trocatalytic and electrochemical energy-storage systems owing to their excep- In realizing clean and viable
tional physicochemical properties, including large surface area, remarkable electrochemical technologies for
mechanical properties, high catalytic activity, chemical stability, and low cost. chemical conversions or energy
In further improving material properties tailored to meet application-specific storage, it is of utmost importance
requirements, heterostructure construction holds significant advantages, to develop economical and high-
benefiting from the synergistic effect between constituents involved. TMD- performance electrode materials.
based heterostructures have been widely explored recently, giving rise to Transition metal dichalcogenides
diverse materials with desirable characteristics such as significantly increased (TMDs), in particular, are attractive
interfacial contact of low resistance for efficient electron transfers, constitu- candidates on account of their
ent-dependent electronic structure, tunable layer distances facilitating easily exceptional physicochemical
intercalation of redox species, and increased surface area for effective interac- properties. To further enhance
tion with electrolyte. In this review, TMD-based heterostructures are assessed their performances, the
for performance in electrocatalytic conversion (hydrogen evolution reaction) construction of TMD-based
and electrochemical energy-storage systems (NiB/LiB/supercapacitors). The im- heterostructures has been
pactful strategies employed in overcoming key challenges are evaluated, and identified to yield a new class of
finally, future directions for TMD-based heterostructure construction are pre- promising materials with enhanced
sented. functionality, stability, and
efficiencies, even surpassing that of
noble metal benchmarks. In this
INTRODUCTION
review, we highlight recent
The constant pursuit for development and modernization necessitates extensive and progress in the development of
ever-increasing global energy consumption, imposing strain on our current availabil- TMD-based heterostructures for
ity of non-renewable fossil fuel resources and raising concern about eventual deple- electrocatalytic conversion systems
tion.1,2 Furthermore, the emission of greenhouse gases and pollutants with time (hydrogen evolution reaction) and
greatly induces harm on the environment in the form of ozone depletion, eutrophi- electrochemical energy-storage
cation, smog, global warming, acid rain, and many other related effects.1,2 These systems (LiB/NiB/supercapacitors).
perspectives question a sustainable future, influencing the continued search for Vital strategies for rational design
viable technologies capable of effectively resolving our global environmental and of these heterostructures are
energy dilemmas.3 Among the methodologies identified, electrochemical ap- discussed, and future directions in
proaches are seemingly promising solutions,4 offering a means to harness intermit- overcoming key limitations are
tent renewable energy sources, including wind and solar, for electricity generation provided. We hope that the
which, in turn, can be channeled through electrocatalytic conversion technologies perspectives presented will
for a clean and sustainable means of driving specific chemical conversions while facilitate advancements in
limiting production of undesired products with negative environmental effects. attaining new and improved TMD-
The electricity generated by renewable sources can be channeled through en- based heterostructures better
ergy-storage systems for later use, allowing high power over a short duration and suited for the respective
point-of-need utilization of the stored energy.5 To make possible these technologies applications.
on the industrial scale, the intrinsic physicochemical properties of electrode mate-
rials is of utmost importance, directly correlating with the performance and

526 Matter 2, 526–553, March 4, 2020 ª 2020 Elsevier Inc.


efficiencies of processes which may be derived.6 From the materials aspect, the
numerous inorganic materials, including Earth-abundant transition metal hydrox-
ides, sulfides, carbides, phosphides, nitrides, and others, encompassed into elec-
trode designs demonstrate high redox activities but suffer in terms of electrical con-
ductivity due to their characteristic semiconducting nature.7–9 Conversely,
electrodes derived from purely carbon-based material such as graphene and carbon
nanotubes demonstrate excellent electrical conductivity but, unfortunately exhibit
poor redox activities. Considering this challenge in attaining superior electrochem-
ical performance, the careful selection and combination of multiple solid-state ma-
terials into formation of heterostructures has become an appealing concept for the
associated constituents to institute an advantageous synergistic effect that endows
the material with enhanced and favorable physicochemical properties tailored to
satisfy the application-specific requirements.10

In particular, design and fabrication of active heterostructure electrodes involving


nanomaterials is appealing. On the basis of dimensions, the nanomaterial constitu-
ents may be classified into 0D, which comprises nanoparticles or quantum dots, 1D,
as in nanowires or nanotubes, and lastly, 2D layered nanomaterials such as nano-
sheets (NSs) and nanoflakes. Among these, the 2D layered nanomaterials are typi-
cally prospective candidates for the fabrication of heterostructure electrodes on ac-
count of their unique and remarkable structural properties, having atomic-layer
thickness and largest exposed surface area, effectuating the greatest surface redox
reactions.11 Furthermore, since the advent of graphene in 2004, a wide spectrum of
2D nanomaterials has been developed for consideration, including transition metal
dichalcogenides (TMDs), MXenes, halides, oxides, and hydroxides, among
others.12,13

TMDs, in particular, having the chemical formula MX2, where M denotes transition
metal atoms (M = Mo, W, and others) and X represents a chalcogen atom (X = S,
Se, or Te), are very attractive candidates because of their exceptional physico-
chemical properties, including large surface area, high mechanical properties,
good semiconducting ability, high catalytic activity, high chemical stability, low
cost, and easy synthesis.14–18 Numerous research works have also highlighted
their tremendous potential for applications in energy-conversion and storage sys-
tems, with performances comparable with benchmarks set by noble metal-based
catalysts including Pt, IrO2, and RuO2.19 Given these desirable characteristics of
TMDs, extensive research has been devoted to the exploration of improved
TMD-based materials. Through incorporating various design strategies and con-
structing heterostructures, a promising class of material can be developed,
benefiting from synergistic effects of each constituent and consequently giving
rise to enhanced functionality, stability, and efficiency, even surpassing that of
benchmarks.20

To this end, it is of utmost importance to discuss collectively key findings and


recent progress to better assess directions for the future. In the past there
were some good review articles focused specifically on G/TMD-based hetero-
structures with respect to other fields of research, such as photovoltaics and pho-
toelectrochemical solar energy conversions.21 Also, there were reviews focusing
particularly on hydrogen evolution and CO2 reduction with consideration over a 1Schoolof Chemical and Biomedical
broad spectrum of heterostructures.22,23 Even just recently, inorganic-based ma- Engineering, Nanyang Technological University,
terials, specifically MXenes and other low-dimensional nanostructures, were as- 62 Nanyang Drive, Singapore 637459, Singapore
sessed for energy-related applications.24 However, the important advances and *Correspondence: jmlee@ntu.edu.sg
design considerations for TMD-based heterostructures were not the key focus, https://doi.org/10.1016/j.matt.2020.01.001

Matter 2, 526–553, March 4, 2020 527


especially so in electrochemical energy systems comprising both energy-conver-
sion fields and energy-storage systems. Here, we present a critical review of
recent developments, discussing the latest impactful work concerning TMD-
based heterostructures, taking into account specific examples of electrocatalytic
conversion systems (hydrogen evolution reaction [HER]) and electrochemical en-
ergy-storage systems (Li-ion batteries [LiB]/Na-ion batteries [NiB]/supercapaci-
tors). Strategies employed in the rational design of these materials are discussed.
Finally, the challenges ahead and directions for future TMD-based heterostructure
construction are presented.

TMD-BASED HETEROSTRUCTURES FOR ELECTROCATALYTIC


CONVERSION SYSTEMS (HER)
As discussed previously, electrochemical approaches are seemingly advanta-
geous solutions, offering a means to harness intermittent but renewable energy
sources (i.e., solar energy) for electricity generation, which in turn can be chan-
neled through electrocatalytic conversion technologies for a clean and sustain-
able means of driving chemical conversions into fuel resources.4 In particular,
electrochemical water splitting utilizing solar-derived electricity holds much prom-
ise, owing to natural abundance of water and formation of high gravimetric en-
ergy density hydrogen, for utility in fuel cells and electricity generation with no
carbon footprint.25–29 An efficient electrochemical water-splitting system,
comprising the two fundamental redox processes of oxygen evolution reaction
(OER) and HER taking place at the anode and cathode, respectively, requires
high activity for both half-cell reactions.8 Notably, however, HER and OER activity
hinders overall water splitting owing to excessive overpotentials and sluggish ki-
netics.7 Through incorporating catalysts as electrodes, the overpotentials neces-
sary to drive redox processes can thereby be lowered, enhancing efficiencies
and kinetics.30 Noble metal Pt catalyst for HER and IrO2 and RuO2 catalyst for
OER, thus far, remain benchmarks offering overpotentials close to zero.31–33 How-
ever, cost concerns and limited availability on Earth greatly hinder their wide-
spread utilization and potential for large-scale consideration in commercial
setups. Nevertheless a variety of other Earth-abundant transition metal electroca-
talysts, including hydroxides, sulfides, carbides, phosphides, nitrides, and others,
have been explored.7–9,34 Many of these electrocatalysts are plagued by limita-
tions including poor electrical conductivities that limit electron mobilities to
catalytic active sites, high overpotentials to drive redox reactions, low working
surface areas, degradation of catalyst with repeated cycling, and cost con-
cerns.35,36 To overcome these limitations, heterostructure construction utilizing
Earth-abundant elements represents a promising strategy to obtain catalysts pos-
sessing remarkable performance.37 In this regard, the advantageous heterostruc-
ture construction is iterated, focusing particularly on HER. Recently, TMDs have
been of great interest and extensively investigated in various strategies of nano-
structure design,38,39 including presence of defects,40,41 mixed metal (alloy),42,43
heteroatom doping,44,45 compositional effects, phase differences,15,46 carbon-
material-containing heterostructures,47,48 and unique morphological characteris-
tics.49 Readers are also referred to Table 1 for a detailed review of performance
results of HER obtained for various electrocatalysts.

Influence of Nanoarchitectures
TMDs possess two types of surface sites, terrace sites at the basal plane and edge
sites at the sides of the nanomaterial.57 As confirmed through theoretical insights
and experimental observations, terrace sites terminate with notably much smaller

528 Matter 2, 526–553, March 4, 2020


Table 1. Performance Results Obtained from Various Heterostructured Electrocatalysts for HER
Heterostructured Overpotential (mV) Tafel Slope Electrolyte Reference
Electrocatalyst (at 10 mA cm2
Current Density)
MoS2/MoSe2 154 46 mV dec1 0.5 M H2SO4 Yang et al.50
1
MoS2/rGO 230 105 mV dec 0.5 M H2SO4 Shizheng et al.51
MoS2/VS2 291 58.1 mV dec1 0.5 M H2SO4 Du et al.52
1
MoS2/WS2 129 72 mV dec 0.5 M H2SO4 Vikraman et al.53
1T/2H MoS2 250 65 mV dec1 1 M KOH Wang et al.46
1
1T/2H MoS2 156 47.9 mV dec 0.5 M H2SO4 Wang et al.54
MoS2/G 421 84 mV dec1 0.5 M H2SO4 Gnanasekar et al.10
1
WxMo1xS2/rGO 96 38.7 mV dec 0.5 M H2SO4 Lei et al.55
(MoS2)x(SnO2)1x/ 263 50.8 mV dec1 0.5 M H2SO4 Ravula et al.56
rGO

fractions of dangling bonds, and hence possess negligible activity for catalytic reac-
tions.57–60 On the other hand, the edge sites predominantly contribute as catalyti-
cally active sites, driving transformations to adsorbed species through electron
transfers and redox processes. In this regard, the fraction of edge to basal sites holds
direct correlation with HER activity. Besides the renowned strategies of employing
low-dimensional nanomaterials (0D, 1D, 2D) and reducing their size dimensions
further, several other key techniques have been explored to increase fraction and
exposure of edge sites for superior HER performance, such as designing suitable ar-
chitectures and engineering of defects.61–63 The careful orientation and formation of
edge standing sheets (nanosheet arrays [NAs]), perpendicularly over a substrate, is
identified as an effective strategy to increase the presence of edge sites and, conse-
quently, working surface area for H+ adsorption.64–66 For example, MoSe2 NAs were
perpendicularly aligned against MoS2 layers and MoS2/MoSe2 heterostructures by
Yang et al., prepared through a facile and inexpensive solvothermal growth
approach (Figure 1A).50 As evidenced from scanning electron microscopy (SEM)
and transmission electron microscopy (TEM) images (Figure 1B), MoSe2 NAs
protrude outward homogenously from the surface of MoS2, providing for greater
specific surface area and presence of active sites with greater exposure. Further-
more, energy-dispersive X-ray spectroscopy (EDX) mapping (Figure 1C) shows
well-dispersed constituent elements over the heterostructure, identifying the abun-
dance of heterojunctions for electron transfers. In addition, the wrinkled morphology
of individual MoSe2 NSs gives rise to porous features (20–50 nm), functioning as
channels facilitating seamless mass transport of ions to active sites. Consequent
to these structural contributions, the MoS2/MoSe2 heterostructures delivered
remarkable HER activity (significantly low overpotential of 154 mV at 10 mA cm2),
far superior in comparison with individual constituents, pure MoS2 (overpotential
of 338 mV), and pure MoSe2 (overpotential of 240 mV) (Figure 1D). Similarly, in
the work of Zheng et al., MoS2 NAs/reduced graphene oxide (rGO) hollow-sphere
heterostructures were synthesized via a dual-templating strategy.51 Via electrostatic
attraction, GO was first coated onto positively charged SiO2 spheres, following
which a hydrothermal process was conducted to grow thin layers of MoS2 NSs
vertically over a GO coating. In the final step, chemical etching was carried out,
effectuating the removal of SiO2 templates. The as-obtained heterostructure cata-
lyst delivered remarkable HER activity attributed to highly exposed S-Mo-S edge
sites of MoS2 NAs while hollow structural features also significantly increased elec-
trochemically active surface areas by a factor of 4.6-fold compared with pure
MoS2, thereby facilitating greater interaction of electrode with electrolyte.

Matter 2, 526–553, March 4, 2020 529


Figure 1. Improving Activity through Structural Design
(A) Schematic illustrating the synthesis of MoS 2 /MoSe 2 heterostructures.
(B) SEM images at low and high magnification, respectively.
(C) TEM image with corresponding elemental mapping for Mo, S, and Se.
(D) Electrochemical analysis data. LSV curves (left) and Tafel plots (right). Adapted with permission
from Yang et al. 50 Copyright 2017, American Chemical Society.

Coupling of TMDs
In heterostructure construction, conjugation of TMD/TMDs is exciting due to the
structural similarity of either component, readily facilitating chemical interactions.
The secondary constituent involved serves as a substrate or support, necessarily
possessing high electrical conductivity, offering high electron mobility and lower
interfacial resistance, facilitating electron transfers across heterojunctions. In this re-
gard, well-dispersed MoS2 nanodots (NDs) decorated over layered thin metallic VS2
NSs (MoS2 NDs/VS2 NSs) were fabricated by Du et al. via a facile in situ hydrothermal

530 Matter 2, 526–553, March 4, 2020


reaction.52 Through optimizing (increasing) the MoS2 ND loading, the authors mini-
mized aggregation of VS2 NSs, giving rise to formation of thin lamellar structures
with enhanced exposure and presence of active edge sites for HER activity. On
the other hand, the incorporated metallic VS2 NSs served as a highly conductive sub-
strate offering excellent electron mobility while also the thin layered structure
provided for enhanced mass transport. As a result of these crucial synergistic merits,
the MoS2 NDs/VS2 NSs heterostructures demonstrated superior acidic HER activity,
earlier onset potentials (low overpotential of 291 mV at 10 mA cm2), and much
smaller Tafel slopes (58.1 mV dec1) in comparison with pure MoS2 NDs and VS2
NSs. Recently, Vikraman et al. reported synthesis of layer-over-layer stacked
MoS2/WS2 heterostructures via sequential procedures starting with a chemical
bath deposition process for initial placement of MoS2 surface on fluorine-doped
tin oxide substrate followed by homogeneous dispersion of tungsten (W) element
over MoS2 layer through radiofrequency sputtering and, lastly, sulfurization, forming
the stacked layer assembly.53 The maximized interfacial contact with hybridization
allows for facile charge transfer across interfaces and utilization at redox active sites
for HER activity. As a result, the obtained catalyst achieved HER performance of
129 mV at 10 mA cm2 and Tafel slopes of 72 mV dec1.

Phase-Conversion Effects
Phase conversion also provides for an effective strategy in heterostructure construc-
tion to induce metallic characteristics, promote charge mobilities, and increase con-
centrations of active sites.67–69 Numerous experimental and theoretical studies have
previously noted that catalytic activity originates from the presence of unsaturated
sulfur atoms found at edge sites while the basal plane of MoS2 or WS2 possesses
inactivity for adsorption/desorption of H+ ions and redox reactions (‘‘inert’’).63,70
Recently, Chen et al. provided more mechanistic insight through density functional
theory (DFT) calculations, highlighting the advantageous characteristics, catalyti-
cally active basal plane and edge sites of 1T converted from 2H phases.71 Further-
more, the phase transition from the semiconducting 2H (trigonal prismatic) to
metallic 1T (octahedral) phase endows the structure with increased electrical con-
ductivity. To date, chemical exfoliation, annealing, modulating strain, plasmonic
hot electrons,67,72 metallic doping, and intercalation methods have been known
for engineering the phase conversions. However, many of these methods necessi-
tate extreme and intensive reaction conditions. Wang et al. recently synthesized
1T-2H MoS2 heterostructures through a simple one-pot annealing method utilizing
Ar and phosphorus gaseous mixture.46 The intercalation of phosphorus resulted in
increased interlayer spacing and weaker layer-to-layer interactions. Furthermore,
the authors also stressed that phosphorus embedded into the lattice through forma-
tion of P-S bonds, evidenced through X-ray photoelectron spectroscopy (XPS)
studies, resulted in the lateral shift of S atomic planes into formation of mixed
1T-2H phases (Figures 2A and 2B). The as-obtained in-plane heterostructures
demonstrated increased electrical conductivity by a factor of more than 500-fold
in comparison with bulk and 2H MoS2 (Figure 2C). In addition, a favorable hydrophil-
ic property was observed as noted from decreased contact angle. Finally, the density
of active sites present in the heterostructures was drastically greater by at least 3-fold
relative to bulk and 2H MoS2. In view of these advantageous effects, the HER activity
in alkaline electrolyte was significantly improved (Figure 2D). In contrast to the
in-plane heterostructures with distinct domains of 1T and 2H phases, hierarchical
heterostructures with homogenized interfacing mixed phases effectuate enhance-
ments in catalytic activity attributed to improved synergistic effects from electrical
conductivity and active sites. Wang et al. developed a facile hydrothermal method
to synthesize hierarchical MoS2 nanorods in situ, composed of mixed 1T and 2H

Matter 2, 526–553, March 4, 2020 531


Figure 2. Inducing Metallic Characteristics through Phase Conversion
(A) Schematic illustrating phosphorus gas-induced phase transition from 2H to 1T MoS 2 .
(B) HRTEM image of mixed 1T-2H in-plane heterostructures with distinct 1T and 2H domains noted.
(C) Comparison of electrical conductivity for 1T-2H in plane MoS 2 heterostructures, 2H bulk MoS 2 ,
and 2H MoS 2 NSs.
(D) Electrochemical analysis data. TOF values, LSV curves, and Tafel plots (left to right,
respectively). Adapted with permission from Wang et al. 46 Copyright 2018, Wiley-VCH Verlag
GmbH & Co. KGaA, Weinheim.

phases.54 Similarly, electrical conductivity and density of active sites was attributed
to the presence of 1T phases. Furthermore, the nanorod morphology of the hetero-
structure provided higher specific surface area by approximately 10-fold relative to
pure 2H MoS2, evidenced from Brunauer-Emmett-Teller (BET) analysis, while also
serving as channels for mass transport and diffusion of H+ to abundant active
edge sites. More importantly, the authors highlighted the formation of Schottky
junctions between metallic 1T and semiconducting 2H interfaces, advantageous
for effective charge separation across heterojunctions and electron mobility. Conse-
quently, the heterostructure electrocatalyst demonstrated remarkable HER activity
in acidic electrolyte, low overpotential of 156 mV at 10 mA cm2, and a small Tafel
slope of 47.9 mV dec1.

Incorporation of Conductive Supports


In bulk TMDs, the necessity for charge-carrier hopping across highly resistive in-
dividual layers gives rise to poor interlayer electron transport, which significantly
hampers catalyst performance. As such, it is highly desirable to rationally

532 Matter 2, 526–553, March 4, 2020


Figure 3. Conjugating with Carbon-Based Substrates
(A) Schematic illustrating experimental setup in CVD furnace used to synthesize MoS 2 NS.
(B) FESEM images of MoS 2 /G heterostructures.
(C) Electrochemical analysis data. LSV curves, graphical illustration of turn-on and overpotentials for respective samples, and corresponding Tafel plots
(left to right, respectively). Adapted with permission from Gnanasekar et al. 10 Copyright 2019, The Royal Society of Chemistry.

combine TMDs with other conductive materials with suitable band alignments
and significantly reduced resistive pathways, so as to effectively realize electron
transfers across interfaces. In this regard, TMD monolayers are often assembled
over carbon-based materials, such as graphene and carbon nanotubes (CNTs),
as a means to take advantage of constituent-dependent properties, namely the
high intrinsic catalytic activity of TMD and superior electrical conductivity of car-
bon substrates.73 Furthermore, the spatial separation of charges and large current
densities that ensue allow for easy migration of electrons toward catalytic active
sites, hence enhancing electron-transfer efficiencies. In doing so, the synergistic
effects endow the resultant heterostructures with improved functionality and elec-
trochemical performance.

Recently, Gnanasekar et al. for the first time fabricated MoS2/G heterostructures,
composed of vertically aligned MoS2 NSs directly grown on graphene support.10
The heterostructure was obtained via a simple chemical vapor deposition (CVD)
method through modulating gas flow and temperature while simultaneously ori-
enting the substrate perpendicularly (Figure 3A). Field-emission SEM (FESEM) im-
ages reveal uniform growth and protrusions of MoS2 NSs outward from the gra-
phene substrate (Figure 3B). More importantly, the vertical alignment effects of
NSs can be noted from the HER activity enhanced by 8-fold (noted from
maximum current densities reached) in comparison with the parallel configura-
tion, attributed to the increased edge site exposure (Figure 3C). Also, the heter-
ostructure catalyst demonstrated earlier onset potentials and moderate overpo-
tentials of 421 mV at 10 mA cm2. The reaction mechanism for each sample
was understood through Tafel slope calculations (h = b log(j) + a, where h, b,
j, and a denote overpotential, Tafel slope, current density, and constant,

Matter 2, 526–553, March 4, 2020 533


respectively). Notably, the heterostructures with either vertically or parallel
aligned MoS2 NSs over graphene substrate, 84 and 104mV dec1 Tafel slopes,
respectively, demonstrated the Volmer-Heyrovsky mechanism with a coupled
Volmer H+ adsorption step (H+ + e = Hads) and a subsequent protonation of
Hads, in accordance with the Heyrovsky mechanism (Hads + H+ + e = H2). To
reap superior catalytic activity, an electrocatalyst exhibiting Volmer-Tafel reaction
mechanism is highly desired with characteristically low Tafel slopes of approxi-
mately 30 mV dec1. However, to date, despite numerous heterostructure cata-
lysts being synthesized, noble metal Pt remains renowned for its behavior
following that of the Volmer-Tafel mechanism. In addition, the optimal binding
of H+ over Pt surfaces accounts for overpotentials close to zero. It is noteworthy
that optimizing binding energies of active catalytic sites with reactant/intermedi-
ate species could significantly alleviate excessive overpotentials, according to the
Sabatier principle. To this end, several works have been undertaken to modulate
the electronic structure of active sites on catalyst and favor suitable adsorption
interactions with reactant species (discussed in the next section).

Alloying/Doping Effects
Among the strategies identified, alloying transition metal or chalcogen elements
such as MoxW1xS242,74 or MoS2(1x)Se2x,43 respectively, or doping secondary
element, non-metals such as N,75 F,76 and P77–79 or metals including Ni,80–82
Co,83–85 Pd,86 and Rh87 into TMD lattice structures holds much promise in improving
the reactivity of edge sites and activating the inert basal plane of TMDs. Lei et al.
fabricated WxMo1xS2/rGO thin films via a facile and scalable wet chemistry
approach (Figure 4A).55 The heterostructures were obtained through varying the
composition of precursor elements in the initial aqueous solution, spin coating,
and, in the final step, annealing at low temperatures. SEM images noted a den-
dritic-like morphology while EDX elemental mapping identified homogeneous dis-
tribution of both transition metal elements, W and Mo, indicative of successful alloy-
ing over the heterostructures. These results were further confirmed by Raman
spectra studies, noting peaks in alloy heterostructures attributed to that of MoS2
and WS2, respectively. Most importantly, the alloying effects were observed both
experimentally and theoretically with the help of DFT calculations. Scanning TEM im-
ages verified the presence of planar defects, in the form of triangular vacancies, and
compositional variations in local environments (Figure 4B). In addition, strain effects
were evidenced by deviations in lattice spacing in comparison with pure MoS2 and
WS2. Consequently, the increased edge sites enhanced activity over the basal plane
while tailoring the local environment toward optimized H+ adsorption.88,89 DFT cal-
culations also revealed that alloying with an optimized ratio of transition metal ele-
ments significantly decreased the energy barrier for HER (Figure 4C). On the basis of
these advantageous contributions as well as the synergistic effects endowed by
coupling with graphene-based substrate, the alloy heterostructure delivered excel-
lent HER performance, significantly low overpotential of 96 mV at 10 mA cm2, and
small Tafel slopes of 38.7 mV dec1, comparable with even that of noble metal Pt. It
is important to note that electrocatalytic activity relies on the adsorption profiles of
reaction intermediates onto catalytic active sites, in accordance with Sabatier’s prin-
ciple. Although the focus in this section pertains specifically to HER, an extension of
the advantageous alloying/doping strategy to other applications may be empha-
sized. Through rational selection of a secondary constituent element and tailoring
the respective composition of heterostructures, one could modulate the electronic
structure and binding strength of intermediates, giving rise to functionality across
a wide range of applications (OER, oxygen reduction reaction, CO2 reduction reac-
tion, and others).84,90

534 Matter 2, 526–553, March 4, 2020


Figure 4. Introducing Secondary Elements
(A) Schematic illustrating wet chemical approach for synthesizing rGO/W x Mo 1x S 2
heterostructures.
(B) High-angle annular dark-field scanning TEM images obtained for heterostructures at high
magnification, with observed defects, triangular domains of Mo, triangular domains of S, and
triangular vacancies (left to right, respectively).
(C) Theoretical insights obtained from DFT calculation noting activation energy barriers of Volmer
reaction (left) and Heyrovsky reaction mechanism (right) for samples with different alloying
compositions. Adapted with permission from Lei et al. 55 Copyright 2017, American Chemical
Society.

Apart from the incorporation of a secondary transition metal element in formation of


catalytically active alloys, the hybridization of multiple constituents at the nanoscale
has also been encompassed in research works, with notable positive effects
attributed to co-catalytic effects and electronic interactions between individual
constituents. For example, Ravula et al. synthesized ternary (MoS2)x(SnO2)1x/rGO
heterostructures, composed of homogenously dispersed (MoS2)x(SnO2)1x nano-
particles (average diameter of approximately 7 nm) anchored over rGO substrate,
via a hydrothermal method.56 The authors employed cationic ionic liquids as
solvents, which aided the synthesis of heterostructures through neutralizing the
negatively charged surfaces of GO precursor, thus preventing ready aggregation
of exfoliated GO. In addition, anionic charged MoO42 precursors could easily
interact with GO surfaces for growth and nanoparticle formation. Furthermore,
Mo species, possessing less oxophilic properties in comparison with Sn, notably
scavenged sulfur from decomposed thiourea, leading to formation of discrete
MoS2 and SnO2 phases. Most importantly, the advantageous effect of multiple
constituent nanohybrids on HER activity could be evidenced from comparison of
Tafel slopes, notably 50.8 mV dec1 for (MoS2)x(SnO2)1x/rGO and 105.5 mV

Matter 2, 526–553, March 4, 2020 535


dec1 for MoS2/rGO heterostructures. In addition, electrochemical impedance
spectroscopy analysis similarly confirmed significantly lower interfacial charge-trans-
fer resistance over nanohybrids in contrast to MoS2/rGO or SnS2/rGO heterostruc-
tures. To this end, it could be concluded that higher conductivity and ease in charge
accumulation at MoS2 edge sites, due to co-catalytic effect of SnO2 and electronic
contributions of rGO substrate, effectuated remarkable HER activity of 263 mV over-
potential at 10 mA cm2.

TMD-BASED HETEROSTRUCTURES IN ELECTROCHEMICAL


ENERGY-STORAGE SYSTEMS (LiBs, NiBs, SUPERCAPACITORS)
Energy-storage systems including lithium/sodium ion batteries (LiBs and NiBs) have
long since taken center stage as energy sources, fueling our daily technological
needs, powering of portable gadgets as in smartphones and laptops.91 However,
currently employed graphitic anodes in commercial battery systems unfortunately
exhibit low theoretical capacities of only 372 mA g h1, which greatly impairs utility
for larger, more demanding applications, such as electric vehicles.92–96 Recently,
another class of advanced electrochemical energy-storage system, supercapacitors,
has also gained significant interest from scientific communities across the globe in
view of advantageous high power densities, superior charging-discharging rates,
and prolonged cycling lifetimes.5,97,98 On the downside, however, supercapacitors
compromise on energy densities.99 The superior function of these energy-storage
systems largely depends on the ability of employed electrodes to deliver high power
and energy densities, such that rapid, extensive, and stable insertion and extraction
processes of ions can be achieved for point-of-need utilization.100 Also, these pro-
cesses typically induce volumetric expansions and, consequently, mechanical stress,
pulverizing and degrading the electrodes, resulting in diminishing capacities over
extended use. In this regard, a remarkable electrode material characteristically
also resists wear and tear incurred, keeping it to a minimum. In realizing next-gener-
ation rechargeable batteries and supercapacitors, it is of utmost importance to first
devise high-performance materials with greater capabilities in overcoming the
aforementioned shortcomings. To this end, scientific endeavors have been focused
on the design and development of advanced electrode materials with a wide spec-
trum of materials, oxides, hydroxides, sulfides, and others being explored.101,102

In the assessment of electrode materials, TMDs are promising candidates due to


their desirable intrinsic properties, including van der Waals interacting layered
morphology for facile intercalation of ions for redox reaction, maximized surface
area, and relatively higher theoretical capacities.103,104 To further enhance perfor-
mance, various strategies have been incorporated, yielding heterostructures en-
dowed with greater functionality. Here, recent developments and strategies encom-
passed into fabrication of these heterostructures are discussed. Readers are also
referred to Table 2 for a detailed review of performance results obtained from
various electrodes.

Influence of Nanoarchitectures
The exfoliation of bulk TMDs into individual NSs introduces unstable surfaces with
high energetics. Aggregation effects dominate, resulting in restacking phenomena
with ready formation of interlayer van der Waals interactions.115 The intimate stack-
ing and close contact between NSs greatly inhibit metal ion diffusion into and trans-
port within interspaces, hence reducing performance and preventing rapid and
extensive storage of ions. To this end, homogeneous anchoring onto carbon sub-
strates and formation of hierarchical nanostructures have been widely demonstrated

536 Matter 2, 526–553, March 4, 2020


Table 2. Performance Results Obtained from Various Heterostructured Electrodes for Energy Storage (NiB/LiB/Supercapacitors)
Heterostructured Electrode Current Density (A g1) Reversible Capacity (for batteries)/Specific Reference
Capacitance (for Supercapacitors)
Exfoliated graphene/MoS2 (EG-MoS2) 1 1,250 mA h g1 after 150 cycles in LiB Wang et al.64
509 mA h g1 after 250 cycles in NiB
MoS2/N-doped graphene (MoS2/NG) 1 198 mA h g1 after 1,000 cycles in NiB Xu et al.105
Yolk-shell MoS2/C 0.2 1,016 mA h g1 after 50 cycles in NiB Pan et al.106
1
MoS2/NG 0.1 1,060 mA h g after 150 cycles in LiB Wu et al.107
3D 1T MoS2/graphene foam 0.05 313 mA h g1 after 200 cycles in NiB Geng et al.108
1
MoS2/C/3DPGF-NS 5 z200 mA h g after 1,000 cycles in NiB Jiang et al.109
1T MoS2/G 0.5 907 mA h g1 after 60 cycles in LiB Zheng et al.110
1
MoS2/CNS 1 381 F g retaining 92% of capacitance Chen et al.111
after 3,000 cycles of charging/discharging
Hollow sandwiched structure 0.1 248 F g1 retaining 85% of capacitance Quan et al.112
of C/MoS2/C after 3,000 cycles of charging/discharging
3D MoS2@CNT/rGO 0.1 129.6 mF g1 retaining 94.6% of capacitance Wang et al.113
after 10,000 cycles of charging/discharging
Hetero-phased 1T-2H MoS2/rGO 1 416 F g1 retaining >90% capacitance Gigot et al.114
aerogel after 50,000 cycles of charging/discharging

to alleviate instability issues.116–119 From the structural perspective, the control over
nanoarchitectures with differing morphologies governs electrochemical activity, by
virtue of improving interfacial contact with electrolyte and favoring the easy mass
transport of metal ions.64,120,121 Amid these nanoarchitectures, the vertical orienta-
tion of TMD sheets perpendicularly over a substrate surface (edge standing) realizes
shortened diffusion pathways for ion transport as opposed to the horizontal align-
ments (basal plane stacking). In addition, each sheet directly contacts a conductive
substrate for electron transfers across the interface and rapid migration in the coaxial
direction. To this end, Wang et al. conjugated constituents with differing hydropho-
bicity properties, hydrophobic EG (electrochemically exfoliated graphene) of high
quality and carbon content, and hydrophilic MoS2, to obtain vertically oriented
EG-MoS2 heterostructures, as shown in Figure 5A.64 SEM and TEM images evidence
the presence of dense edge standing sheets, protruding outward from the graphene
substrate surface. In view of the structural geometry—vertically aligned relative to
laterally stacked sheets—high mass loading and utilization of active MoS2 (95 wt
%) on graphene substrate was achieved. Also, the graphene substrate endowed
the heterostructure with greater stability against volume-change-induced mechani-
cal stress while also allowing rapid electron transfer across the interface and fast
transport along the distal direction of MoS2 sheets. The as-obtained heterostruc-
tures demonstrated remarkable performance for use as anodes in LiBs, excellent
specific capacitance of 1,250 mA h g1 at 1 A g1 after 150 cycles, and in NiBs
high specific capacitance of 509 mA h g1 at 1 A g1 after 250 cycles. Li et al. further
investigated the geometrical effects, more specifically, areal density and height of
vertically aligned TMD sheets, on the corresponding performance for NiBs.122
Through tuning of heating rates, the authors varied the durations of nucleation
and growth of TMD, achieving control over morphologies of resultant MoS2/
N-rGO heterostructures (Figure 5B). Most importantly, with a smaller ramping
rate, denser and shorter MoS2 sheets were obtained, and vice versa. X-ray diffraction
(XRD) spectra revealed a characteristic peak corresponding to c axis of MoS2 with
interlayer spacing of 6.48 Å. As such, it is worth noting that the interlayer spacings
were largely unmodified. The highly dense packing configuration of MoS2 conse-
quently provided for increased utilization of active species. On the other hand,

Matter 2, 526–553, March 4, 2020 537


Figure 5. Improving Activity through Structural Design
(A) Schematic illustrating synthesis of EG/MoS 2 heterostructures (left) and SEM and TEM images
obtained for heterostructures (right). From Wang et al. 64 Copyright 2017, Wiley-VCH Verlag GmbH
& Co. KGaA, Weinheim.
(B) Schematics illustrating synthesis of MoS 2 /N-rGO heterostructures with control over areal
density and sheet height through varying heating rates. From Li et al.122 Copyright 2018, Wiley-VCH
Verlag GmbH & Co. KGaA, Weinheim.
(C) Schematics illustrating synthesis of MoS 2 /NG heterostructures through control of utilized
precursors and solvent in experimental setup. From Xu et al.105 Copyright 2018, Wiley-VCH Verlag
GmbH & Co. KGaA, Weinheim.

the short layer heights effectuated easy mass transport of ions over shortened diffu-
sion paths, aiding in energy storage. Recently, more elaborate control of morphol-
ogies for MoS2/N-doped graphene (NG) heterostructures was demonstrated by Xu
et al.105 Through varying precursors and solvent components in reaction setups, the
authors influenced size, number of layers, and interlayer spacing of anchored MoS2

538 Matter 2, 526–553, March 4, 2020


NSs (Figure 5C). Notably, the anchorage of ultrasmall MoS2, with 1–3 layer couplings
and increased interlayer spacing to between 6.8 and 6.9 Å, onto NG delivered the
most promising performance as anode in NiBs, 198 mA h g1 at 1 A g1 after
1,000 cycles.

Besides perpendicular orientation on substrate surfaces, various other intricate


nanoarchitectures and morphologies have been demonstrated with regard to en-
ergy-storage applications. For example, Chen et al. synthesized heterostructures
composed of MoS2 and carbon nanosheets (CNS) via a facile hydrothermal method
(Figure 6A).111 Surface treatment with concentrated nitric acid endowed the CNS
surfaces with O-functional units, which served as nucleation sites for homogeneous
anchorage and growth of subsequent MoS2. SEM and TEM images revealed a
nanoflower morphology with curved MoS2 layers (Figure 6B). In addition, the CNS
structures were well interconnected, allowing for facile electron transport and
conductivity across interfaces and aiding in the charge-discharging processes.
BET analysis identified greater specific surface area for electrochemical activity of
approximately 20% over MoS2/CNS heterostructures in contrast to pure MoS2.
Consequently, superior performances were obtained, reaching a maximum specific
capacitance of 381 F g1 in KOH electrolyte while retaining 92% of capacitance over
3,000 cycles (pure MoS2 achieved 255 F g1 specific capacitance and retention of
80%). Furthermore, void or hollow structures are appealing, serving as volume-
change buffers with charging-discharging processes.112,123,124 As a result, the
electrodes are protected for prolonged cycling and greater stability characteris-
tics.123,124 To this end, Pan et al. synthesized yolk-shell heterostructures (MoS2/C)
composed of an inner MoS2 yolk surrounded by a carbon shell with internal voids
in between.106 The authors utilized etching procedures with differing concentrations
of hydrogen peroxide to finely control the internal void spaces. In the absence of
voids, the mechanical stress induced with lithiation resulted in breakage of the exte-
rior carbon shell framework, compromising on electrical conductivity. It is worth
noting that the introduced voids effectively served as a buffer against volume expan-
sion, providing for greater stability of structures. Consequent to these structural con-
tributions, the MoS2/C heterostructure, as anodes in LiBs, delivered high reversible
capacities of 1,016 mA h g1 at 200 mA g1 current density after 50 cycles. Recently,
hollow sandwiched structures of C/MoS2/C heterostructures were obtained by Quan
et al.112 The authors carried out a Gibbsite template-based strategy (Figure 6C) to
coat alternating layers of polydopamine (a carbon precursor) and MoS2, after which
the samples were annealed at elevated temperatures. In the final step, etching was
conducted to remove templates, forming the hollow sandwiched structures, as
evident in TEM images (Figure 6D). Interestingly, BET analysis noted a remarkably
high electrochemical active area of 543 m2 g1 for the hexagonal-shaped
C/MoS2/C hollow nanoplates. It is worth noting that the hollow spaces effectively
contributed to increased specific surface areas. Also, the unique morphology
facilitated diffusion of electrolyte through nanosized pores on the surface of the
nanoplates into hollow spaces, thereby increasing interfacial contact with electrolyte
and enhancing performance.123,124 Owing primarily to the morphological con-
tributions, the as-obtained heterostructure delivered excellent capacitance perfor-
mance, 248 F g1 at 0.1 A g1.

Incorporation of Conductive Materials


Poor electronic conductivity, attributed to the semiconducting nature of TMDs,
has been detrimental to electron mobilities and redox activities, significantly
compromising charging-discharging rates.125 In this regard, numerous research
works have coupled TMDs with highly conductive carbon materials, such as CNTs

Matter 2, 526–553, March 4, 2020 539


Figure 6. Complex Architectures in Nano Dimensions
(A) Schematic illustrating synthesis of MoS 2 /CNS heterostructures.
(B) SEM (left) and TEM (right) images of MoS 2 /CNS heterostructures. From Chen et al. 111 Copyright 2017, Elsevier BV. All rights reserved.
(C) Schematic illustrating templated synthesis of hollow C/MoS 2 C heterostructures.
(D) TEM images of C/MoS 2 C heterostructures at low (left) and high (right) magnification. From Quan et al. 112 Copyright 2019, Wiley-VCH Verlag
GmbH & Co. KGaA, Weinheim.

and graphene, to derive improved functionalities from rapid electron migration and
transfers.55,110,126–128 Furthermore, carbon materials offer high specific surface area,
excellent mechanical properties, and physicochemical stability.129–132 The hybridi-
zation with carbon materials in the formation of these TMD-based heterostructure
electrodes is beneficial to obtaining superior electrochemical performance charac-
teristics while maintaining stable cycling lifetimes.133–137 Several studies have also
encompassed conjugating active materials with 3D conducting porous structures,
taking advantage of interconnected networks for electron transport while pores
facilitate rapid diffusion of electrolyte. In addition, the remarkable mechanical prop-
erties of these scaffolds are harnessed to withstand significant volume changes with
charging-discharging processes, favoring energy-storage applications.138,139 In the
work of Jiang et al., the advantageous effects of 3D porous graphene framework
(PGF) was demonstrated with the anchoring of various typical electrode materials
onto N, S co-doped 3D PGF to note enhancements in energy densities for
NiBs.109 Besides endowing the heterostructures with greater stability against
volume variations, the robust graphene frameworks also offered several advantages,
including increased electrical conductivity due to interconnected networks, easy
ionic migration to active sites as a result of channels with shortened diffusion lengths,
and increased interfacial contact with electrolyte. Furthermore, it is worth noting that
the presence of heteroatom co-doping, N and S, further presented defect sites of
high reactivity contributing to Na storage. As such, MoS2/C/3D PGF-NS heterostruc-
tures presented enhanced performance as anodes in NiBs, achieving remarkable
capacities of 317 mA h g1 at 5 A g1 with retention of approximately 60% of its
initial capacity after 1,000 cycles. In another work, Wang et al. reported synthesis
of 3D MoS2@CNT/rGO heterostructures, as shown in Figure 7A.113 Based on SEM

540 Matter 2, 526–553, March 4, 2020


Figure 7. Conjugating with Carbon-Based Substrates
(A) Schematic illustrating synthesis of MoS 2 @CNT/rGO heterostructures.
(B) SEM images at low and high magnification (left and middle, respectively) and stability test results (right). From Wang et al. 113 Copyright 2017, Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim.
(C) Schematic illustrating synthesis of 3D MoS 2 /rGO aerogel heterostructures (left) and SEM images of MoS 2 /rGO coated with nickel foam (right). From
Xu et al. 140 Copyright 2018, Elsevier BV. All rights reserved.

images (Figure 7B), interconnected networks of CNTs with crosslinked structures are
evident, highlighting the facile electron transport to redox active sites and MoS2
dispersed over CNTs. Furthermore, the thin-layer rGO sheets further enhance the
contact between individual constituents serving as additional electron pathways.
The authors also noted pores of at least 30 nm, which could provide easy electrolyte
diffusion and increased contact with MoS2 species, contributing to enhanced perfor-
mance. Consequent to these structural and ternary composition contributions,
the MoS2@CNT/rGO heterostructure delivered a high specific capacitance of
129.6 mF g1 at 0.1 A g1. Also, it is worth noting that CNTs and rGO possess
remarkable mechanical properties of stress and strain resistance. As such, the
incorporation of these elements endows the heterostructure with greater applica-
bility for use as flexible electrodes possessing high stabilities. To this end, the
authors conducted stability testing through galvanostatic charging-discharging

Matter 2, 526–553, March 4, 2020 541


over MoS2@CNT/rGO, which demonstrated retention of 94.6% capacitance after
10,000 cycles. In addition, the electrodes under bending delivered similar capaci-
tance retention with negligible differences, highlighting promising utility as flexible
electrodes. Xu et al. constructed 3D MoS2/rGO aerogel hybrids via a hydrothermal
co-synthesis methodology (Figure 7C).140 The presence of O species in rGO as well
as defective sites facilitated homogeneous anchorage of MoS2 layered sheets over
the porous aerogel, preventing restacking.141 SEM images evidently show MoS2
sheets protruding from rGO aerogel substrate surfaces. More importantly, the au-
thors noted that through varying the composition (loading) of MoS2, an optimum
highest specific capacitance of 472 F g1 could be obtained. Stability testing over
aerogel composites revealed an excellent 98.9% capacitance retention after
10,000 cycles. These works signify that 3D scaffolds are advantageous in perfor-
mance and stability. It is noteworthy that well-defined control of pore sizes seem-
ingly holds a correlation with the electrochemical activity by means of optimizing
the effective working surface area. However, to date much of the pore-size distribu-
tions are wide ranging and uncontrolled, requiring further work to develop facile
strategies with greater influence on resultant structures.

Besides incorporating carbon-based supports, conjugation with other electrically


conductive and structurally similar metal oxides/sulfides has been demonstrated.
For example, MoO2 possesses metallic characteristics with low resistivity, presence
of multiple valence states, and a strong metallic Mo-Mo bonding that favors interfa-
cial charge transfer and redox reactions.142,143 Concerning MoO2, however, poor
capacitive retention has been demonstrated due to severe volume changes,
increasing resistivity, and degradation of structure with prolonged utilization.144
On the other hand, MoS2 possesses high ionic conductivity owing to facile interca-
lation of ionic species, as well as structural stability.145 In this regard, several research
works have devised heterostructures through combining both components.145,146
Tian et al. fabricated N-doped hollow carbon spheres covered by MoO2/MoS2
sheets.142 As shown in the synthesis schematics, polyaniline was first coated onto
polystyrene sphere (PS) templates, following which a hydrothermal growth proced-
ure was carried out to guide uniform distribution of MoO2/MoS2 sheets onto exterior
surfaces. Finally, carbonization was conducted to remove PS templates while also
forming the microporous N-doped carbon shell. Most importantly, attributed to
the presence of metallic MoO2, the longest discharge time, remarkable maximum
specific capacitance of 569 F g1 at 1 A g1, and high Coulombic efficiencies
were attained. In the work by Mu et al., MoO2/MoS2 composites supported on
heat-treated polyaniline nanofibers were synthesized.147 Notably, the flexible 1D
nanofiber support present in the heterostructure facilitated cycling stability and
prevented severe volume change taking place with charging-discharging processes,
whereas the large specific surface area offered an increased active site for redox
processes and charge storage. Due to the synergistic effect from respective constit-
uents, the electrode delivered a high specific capacitance of 439.8 F g1 at 1 A g1.
Hollow core-shell NiCo2S4/MoS2 dodecahedron heterostructures were fabricated in
the work of Song et al. via an elaborate multistep synthesis procedure involving
ZIF-67 as starting template.148 Ternary NiCo2S4 endowed the resultant heterostruc-
tures with relatively greater redox activity and electrical conductivity, in contrast to
binary counterparts.149–151 Owing to hollow structures with increased specific sur-
face area for redox activity, and interaction between both phases, specific capaci-
tance and stability were improved, giving rise to electrodes with high performance
characteristics in supercapacitors. Huang et al. demonstrated a facile, single-step
hydrothermal synthesis of Ni3S4/MoS2 grown directly onto carbon paper.152 SEM
images revealed a flower-like morphology of MoS2 with nanoparticles of Ni3S4

542 Matter 2, 526–553, March 4, 2020


forming clusters. The authors noted that the Ni3S4 nanoparticles favored electron
mobility through offering conduction with low resistivity, hence enhancing activity.
Consequently, excellent specific capacitance of 1,296 F g1 was achieved at 1 A
g1 as well as superior stability characteristics, evidenced by 96.2% retention of
initial capacitance even after an extensive 5,000 cycles at 5 A g1.

Interlayer Expansion
As aforementioned, ions intercalate in between individual layers, migrating to redox
centers to initiate charge-storage mechanisms. The interlayer spacing also governs
the ease and kinetics of intercalation of ions and accessibility to redox centers, signif-
icantly increasing effective surface area for charge storage.153–155 In this regard, Wu
et al. synthesized MoS2/NG heterostructures via a poly(vinylpyrrolidone) (PVP)-assis-
ted hydrothermal route to effectuate interlayer expanded MoS2 nanoflowers,
assembled over NG substrate.107 PVP molecules were absorbed over (002) planes
of MoS2 sheets, thereby preventing restacking and agglomeration of MoS2 while
also contributing to interlayer expansions from characteristic 6.2 Å to 7.1 Å, as
confirmed by XRD spectra studies. When used as anode in LiB, superior electro-
chemical performance was attained due to facile intercalation of Li. Excellent revers-
ible capacity of 1,060 mA h g1 at 100 mA g1 current density after 150 cycles was
achieved, comparatively greater than with conventional MoS2/G electrodes (deliv-
ering reversible capacity of 416 mA h g1) without evident interlayer expansion.
Recently, Ji et al. further demonstrated individual effects of interlayer expansion
(IE) and edge orientation (EO) over MoS2/rGO heterostructures, obtained via a mi-
crowave-assisted solvothermal approach.156 Edge orientation resulted from differ-
ence in hydrophobicity between MoS2 and rGO constituents. As evidenced from
SEM images, uniformly distributed edge-oriented MoS2 are seen over rGO substrate
(Figure 8A). On the other hand, IE, through passing of pressurized nitrogen gas, to
9.4 Å from typical 6.15 Å in bulk MoS2 was confirmed through XRD studies based on
characterization of two new reflection peaks found within the low-angle range (Fig-
ure 8B). More importantly, the authors highlighted the positive electrochemical
behavior of IE/EO MoS2/rGO heterostructures over the EO MoS2/rGO counterpart
as observed from greater integrated area of cyclic voltammetry (CV) curves for
each scan rate (Figure 8C). Also, notably, with decreasing radii size of cations, spe-
cific capacitance increased drastically, indicative of facile intercalation and storage
(Figure 8D). In addition, the specific capacitance in aqueous HCl electrolyte over
IE/EO MoS2/rGO reaches a remarkable maximum value of 346.5 F g1, substantially
greater by at least 2-fold than 134.0 F g1 of EO MoS2/rGO at 1 mV s1. To this end,
the authors stressed the beneficial effects of IE. Several other works encompassing
doping with polymeric intercalants have also been reported. Chao et al. sandwiched
PANI into interlayers of MoS2 deriving expansion to 7.0 Å over MoS2/PANI/rGO het-
erostructures.157 The resultant electrodes delivered remarkably high specific capac-
itance of 520.0 F g1 at 1 A g1 current density. Moreover, the authors also stressed
that the conductive dopants further served as conductive channels across the lateral
surface of MoS2 sheets while limiting volumetric expansions and induced mechanical
stress. Finally, the remarkable stability of the heterostructures was noted from stabil-
ity test, whereby 81.9% of initial specific capacitance was retained after an extensive
40,000 cycles.

Phase Conversions
Phase transformations from trigonal prismatic 2H TMD to octahedral 1T phases are
well known to induce transition from semiconducting to metallic characteristics,
respectively,46,158–160 as a result effectively increasing the overall conductivity
of electrodes by a factor of 107 times, thereby improving electrochemical

Matter 2, 526–553, March 4, 2020 543


Figure 8. Expanding Interlayers for Ease in Ion Intercalation
(A) SEM image of EO&IE MoS 2 /rGO heterostructures.
(B) XRD spectra comparing EO&IE MoS 2 /rGO and EO MoS 2 /rGO heterostructures.
(C) CV results comparing EO&IE MoS 2 /rGO and EO MoS 2 /rGO heterostructures.
(D) Specific capacitance for EO&IE MoS 2 /rGO and EO MoS 2 /rGO heterostructures (left and right,
respectively) for electrolytes with differing cation species. Adapted with permission from Ji et al.156
Copyright 2018, American Chemical Society.

activity.159,160 Furthermore, the hydrophilic properties of 1T TMD sheets greatly


benefit interaction between electrodes with electrolytes.161 To date, methods for
driving these transformations include non-metal elemental doping in interlayers
including N, S, O, or P, organo-metal/metal intercalation, specifically alkali metals
such as Li, Na, or K, inducing strain, injecting charges, temperature effects,

544 Matter 2, 526–553, March 4, 2020


introducing defects such as S vacancies, and others.153,162–169 Several notable
studies have demonstrated that metallic 1T or mixed 1T-2H phases rather than
2H-phase TMDs are beneficial for performance in energy-storage applications,
including batteries and supercapacitors.159,170 Earlier work on 1T phased MoS2 elec-
trodes was achieved by Acerse et al, via an organolithium exfoliation approach to
obtain 70% 1T phase composition.159 Subsequently, the authors carried out a
filtration method and restacked MoS2 sheets into film electrodes, which delivered
a volumetric capacitance of between 400 and 700 F cm3. Most importantly, the
remarkable performance was attributed to remarkable electrical conductivity,
hydrophilic property, and dynamic intercalation of ions for storage. Geng et al. fabri-
cated 3D 1T MoS2/graphene foam heterostructures for use as anodes in NiBs.108 A
solvothermal approach, in autoclave at 200 C with MoO3 and thioacetamide precur-
sors as well as graphene foam in ethanol solvent, was employed to obtain uniform,
well-dispersed nucleation and growth of 1T MoS2 on both surfaces, interior and
exterior, of hollow graphene foam. The authors further noted through resistivity
testing that conjugation of 1T MoS2 with graphene gives rise to lower resistivity
even surpassing that of pure metallic MoS2. Also, with the unique sandwiched
configuration, high mass loading of active 1T MoS2 was realized. In addition, porous
graphene foam together with hollow tubing effectuated increased specific surface
area as well as interaction with electrolyte. Benefiting from these structural consid-
erations as well as excellent electrical conductivity and hydrophilic properties
of active 1T MoS2, the as-synthesized heterostructures delivered outstanding
reversible capacity of 313 mA h g1 at a small 0.05 A g1 current density even after
200 cycles, far superior compared with works on 2H MoS2/graphene heterostruc-
tures. Similarly, in another work by Zheng et al., a hydrothermal methodology was
used out to synthesize 1T MoS2/G heterostructure electrodes, composed of
thin, few layered graphene sheets with 1T MoS2 NSs grown in parallel with substrate
alignment.110 When used as anodes in LiBs, high reversible capacity of 907 mA h g1
at 500 mA g1 current density after 60 cycles was obtained. Gigot et al, conjugated
hetero-phased 1T-2H MoS2 with rGO aerogel substrate, obtaining electrodes with
promising utility in supercapacitors.114 The authors conducted a two-step synthesis
procedure, first hydrothermal reaction and subsequently freeze-drying, to obtain
heterostructure aerogels. Hetero-phases were confirmed through XPS, which noted
lower binding energy peaks corresponding to that of 1T MoS2 separated from that of
2H phases by 0.9 eV. HRTEM images also revealed the irregularity in interplanar
spacing, notable differences from interplanar distance of 6 Å as typically observed
in 2H MoS2, indicating the presence of 1T domains. The as-obtained electrodes
demonstrated a maximum specific capacitance of 416 F g1 in NaCl electrolyte while
maintaining repeated utility of electrodes for over 50,000 cycles, in contrast to pure
MoS2 achieving cycling of up to 13,000 cycles only. The authors further stressed that
the superior capacitance resulted from electrochemically active mixed phases
embedded near electrolyte rGO interfaces. In addition, the remarkable stability
demonstrated was due to the presence of rGO substrate protecting active
components from degradation and volume changes. Recently, Ke et al. elucidated
interfacial charge dynamics at boundaries between 1T and 2H and investigated
the consequent effects on energy storage through both experimental and theoret-
ical approaches.168 More importantly, the authors noted through DFT calculations
that the charge-density differences are significantly greater over boundaries, in
contrast to within pure phases, indicating easy charge transfers. Through varying
temperature, the authors yielded differing compositions of 1T to 2H in-plane
MoS2, higher temperatures stabilizing 2H MoS2 and, consequently, decreasing ratio
of 1T to 2H phases and increasing phase boundaries (Figure 9A).158,171 STEM im-
ages noted that post-annealing at 150 C introduced abundant distinct 1T phase

Matter 2, 526–553, March 4, 2020 545


Figure 9. Inducing Metallic Characteristics through Phase Conversion
(A) Temperature-dependent compositional synthesis of in-plane 1T-2H MoS 2 heterostructures.
(B) STEM images of exfoliated MoS 2 sheets and corresponding MoS 2 after exposure to annealing at 150  C.
(C) Graphical illustration of sample capacitance with exposure to varying annealing temperatures. From Ke et al. 168 Copyright 2019, Wiley-VCH Verlag
GmbH & Co. KGaA, Weinheim.

domains into the lattice of MoS2 (Figure 9B). The heterostructures with the presence
of greatest 1T-2H boundaries, maximum at 150 C annealing temperature, yielded a
highest specific capacitance of 438 F g1 (Figure 9C). For temperatures beyond
which 2H domains were preferentially stabilized, greatly reducing hetero-phase
boundaries and as a result, specific capacitance obtained was significantly reduced.
To this end, it is worth noting that in contrast to thermodynamically favored 2H
MoS2, the 1T phase demonstrates greater instability and hence necessitates favor-
able environmental conditions, temperature, and pH environments to maintain
integrity of electrodes over extended durations. Hence, further work and investiga-
tions are requiring to obtain stable metallic MoS2 for more intricate construction of
heterostructure designs.

CONCLUSION AND FUTURE PERSPECTIVES


In this review, the recent developments in construction of TMD-based heterostruc-
tures, particularly, in the fields of electrochemical conversion (HER) and energy
storage (supercapacitors) have been discussed. More specifically, the strategies
that have been encompassed into research works have been focused upon, iterating
the advantageous effects of deriving superior performance characteristics in respec-
tive applications. In HER, fundamental issues with current developments of low-cost

546 Matter 2, 526–553, March 4, 2020


transition metal-based catalysts include excessive overpotentials for driving trans-
formation, poor yields of hydrogen evolved, poor electrical conductivities and cur-
rent densities, low effective working surface areas, and degradation of catalyst
with extensive use.35,36 To this end, heterostructure construction has been estab-
lished as a prominent prospect for overcoming these limitations. Several effective
strategies incorporated into designing heterostructures for further promoting per-
formance were summarized, which include the following: tailoring structural config-
urations (EO), increasing the specific abundance of active sites (either edge sites or
inducing active sites through introducing defects on typically inert basal plane),
modulating electronic structures to preferentially adsorb H+ with optimal binding
(doping, alloying with secondary elements or compounds at the nanoscale), conju-
gation with conductive materials (metallic/superconducting TMDs, carbon-based
materials), and phase transformations (mixed phase 1T-2H TMD-based heterostruc-
tures). On the other hand, in LiB/NiB/supercapacitors, several key limitations were
also noted, namely, compromise between either energy or power densities, poor
electrical and ionic conductivities, low effective working surface areas, volume
expansion with charging-discharging, and degradation of catalyst with prolonged
cycling. In this regard, similarly it was established that designing nanoarchitectures
with suitable morphologies (hollow structures, nanoflower morphologies), interlayer
engineering for increasing ease in ion intercalation, conjugating with conductive ma-
terials (mixed 2H-1T or metallic 1T phase heterostructures or carbon-based mate-
rials),172 and doping with metal/non-metal elements173,174 are promising tech-
niques to derive heterostructure electrodes with enhanced function and
performance toward energy storage. The aforementioned areas of electrocatalysis
and energy storage have witnessed significant progress in recent years, but many
challenges still remain to be resolved before these technologies may be encom-
passed into wider practical applications.

Devising materials possessing both high catalytic activity and excellent electrical
conductivity is urgently needed. For example, recently, group V TMDs have drawn
much attention due to their promising superconducting properties, utilized mostly
as a conductive substrate, while their catalytic activity remains poor, requiring
much more exploration of strategies. Although heterostructures demonstrate prop-
erties attributed to one component typically demonstrating superior conductivity
versus another possessing catalytic activity, it should be noted that the utilization
of multiple constituents necessitates significant mass loading, less justifiable in
contrast to efficiencies obtained. From a large-scale perspective, the excessive con-
sumption of materials to derive performance could exert significant strain from the
economic viewpoint. To this end, it is highly desirable to construct heterostructures
while simultaneously inducing catalytic activity of constituents to effectively reap a
much larger specific abundance of active sites. Another point to note is that in
much of the work conducted thus far, multiple strategies are often incorporated
into designing heterostructures, giving rise to synergistic contributions and eleva-
tion of performances. However, the degree of impact in consideration of individual
factors is much less understood. For example, in energy-storage systems the inter-
layer distances in between stacked sheets of TMD primarily governs ion diffusion
to redox centers and, consequently, charging-discharging rates and power den-
sities.175 It is also worth noting that interlayer charge-carrier hopping holds an in-
verse relationship with increasing interlayer distances, evidently detrimental to elec-
trical conductivity and resultant electrochemical activities. To this end, there has
been little investigation, both experimental and theoretical, to elucidate which of
the two factors holds the dominant contribution to overall performance. These
gaps in understanding require attention to effectuate rational design of

Matter 2, 526–553, March 4, 2020 547


heterostructures with greater control over the nature of competing effects. To move
forward objectively and beyond the current arsenal of strategies, exploration of
alternative methods assisted by theoretical perspectives could shed light on much
needed future directions. Lastly, despite the broad and widely encompassing list
of TMDs available, much focus is still being placed on primarily MoS2, MoSe2,
WS2, and WSe2 TMDs. Perhaps the construction of heterostructures with other
possible selections would bring about distinct differences in physicochemical prop-
erties, possibly giving rise to materials better suited for these electrochemical sys-
tems. Venturing into these less explored TMD constituents could also, reveal
more insights into fundamental processes. Theoretical DFT calculations would
render much assistance for providing timely and extensive screening of material con-
jugations. Given the aforementioned challenges and the significant room for
improvement, much work still requires to be done prior to developing competent
electrochemical energy-storage and conversion systems. However, with progress
in understanding, the outlook for successful development of these technologies is
certainly promising.

ACKNOWLEDGMENTS
This work is supported by Singapore Maritime Institute Maritime Sustainability
(MSA) R&D Programme (grant number M4061829) and Academic Research Fund
(RG17/16) of the Ministry of Education in Singapore. The authors thank Mr. Sng
Yeow Poo at Trinity Offshore and Mr. Ravindra Pallaniappan for their support.

AUTHOR CONTRIBUTIONS
Conceptualization, P.P., V.J., and J.-M.L.; Writing – Original Draft, P.P. V.J., and
J.-M.L; Writing – Review & Editing, P.P., V.J., and J.-M.L; Funding Acquisition,
J.-M.L.; Supervision, J.-M.L.

REFERENCES
1. Chu, S., and Majumdar, A. (2012). 7. Anantharaj, S., Ede, S.R., Sakthikumar, K., G.-H., et al. (2017). Recent advances in
Opportunities and challenges for a Karthick, K., Mishra, S., and Kundu, S. (2016). ultrathin two-dimensional nanomaterials.
sustainable energy future. Nature 488, Recent trends and perspectives in Chem. Rev. 117, 6225–6331.
294–303. electrochemical water splitting with an
emphasis on sulfide, selenide, and phosphide 14. Novoselov, K.S., Mishchenko, A., Carvalho,
2. Jeffrey, C., Raymond, J.K., and Paul, R.P. catalysts of Fe, Co, and Ni: a review. ACS A., and Castro Neto, A.H. (2016). 2D materials
(2003). Energy resources and global Catal. 6, 8069–8097. and van der Waals heterostructures. Science
development. Science 302, 1528. 353, aac9439.
8. Li, Y., Yin, J., An, L., Lu, M., Sun, K., Zhao,
Y.-Q., Gao, D., Cheng, F., and Xi, P. (2018). 15. Jiao, Y., Hafez, A.M., Cao, D., Mukhopadhyay,
3. Cheng, Q., and Yi, H. (2017). A., Ma, Y., and Zhu, H. (2018). Metallic MoS2
FeS2/CoS2 interface nanosheets as efficient
Complementarity and substitutability: a for high performance energy storage and
bifunctional electrocatalyst for overall water
review of state level renewable energy policy energy conversion. Small 14, 1800640.
splitting. Small 14, 1801070.
instrument interactions. Renew. Sustain.
Energy Rev. 67, 683–691. 9. Zhang, Y., and Zou, X. (2015). Noble metal- 16. Bian, L., Sun, J., Han, M., Li, F., Gao, Z., Yang,
free hydrogen evolution catalysts for water Z.X., Song, A., Qu, Y., Gao, W., Shu, L., et al.
4. Laursen, A.B., Dahl, S., Chorkendorff, I., and splitting. Chem. Soc. Rev. 44, 5148–5180. (2018). Phosphorus-doped MoS2 nanosheets
Kegnæs, S. (2012). Molybdenum sulfides— supported on carbon cloths as efficient
efficient and viable materials for electro- and 10. Gnanasekar, P., Periyanagounder, D., and hydrogen-generation electrocatalysts.
photoelectrocatalytic hydrogen evolution. Kulandaivel, J. (2019). Vertically aligned MoS2 ChemCatChem 10, 1571–1577.
Energy Environ. Sci. 5, 5577–5591. nanosheets on graphene for highly stable
electrocatalytic hydrogen evolution reactions. 17. Xu, J., Huang, Z., Ji, H., Tang, H., Tang, G.,
5. Qi, D., Liu, Y., Liu, Z., Zhang, L., and Chen, X. Nanoscale 11, 2439–2446. and Jiang, H. (2019). g-C3N4 anchored with
(2017). Design of architectures and materials MoS2 ultrathin nanosheets as high
in in-plane micro-supercapacitors: current 11. Yun, Q., Li, L., Hu, Z., Lu, Q., Chen, B., and performance anode material for
status and future challenges. Adv. Mater. 29, Zhang, H. (2019). Layered transition metal supercapacitor. Mater. Lett. 241, 35–38.
1602802. dichalcogenide-based nanomaterials for
electrochemical energy storage. Adv. Mater. 18. Fu, G., Wang, Y., Tang, Y., Zhou, K.,
32, 1903826. Goodenough, J.B., and Lee, J.-M. (2019).
6. Elouarzaki, K., Kannan, V., Jose, V., Superior oxygen electrocatalysis on nickel
Sabharwal, H.S., and Lee, J.-M. (2019). Recent 12. Geim, A.K., and Grigorieva, I.V. (2013). Van indium thiospinels for rechargeable Zn-air
trends, benchmarking, and challenges of der Waals heterostructures. Nature 499, 419. batteries. ACS Mater. Lett. 1, 123–131.
electrochemical reduction of CO2 by
molecular catalysts. Adv. Energy Mater. 9, 13. Tan, C., Cao, X., Wu, X.-J., He, Q., Yang, J., 19. Wang, T., Pang, H., Xue, H., Yu, Y., and Chen,
1900090. Zhang, X., Chen, J., Zhao, W., Han, S., Nam, S. (2017). MoS2-Based nanocomposites for

548 Matter 2, 526–553, March 4, 2020


electrochemical energy storage. Adv. Sci. 4, 34. Jiao, Y., Zheng, Y., Jaroniec, M., and Qiao, enhanced hydrogen evolution reaction. Adv.
1600289. S.Z. (2015). Design of electrocatalysts for Energy Mater. 8, 1801345.
oxygen- and hydrogen-involving energy
20. Tan, C., Chen, J., Wu, X.-J., and Zhang, H. conversion reactions. Chem. Soc. Rev. 44, 47. Trogadas, P., Fuller, T.F., and Strasser, P.
(2018). Epitaxial growth of hybrid 2060–2086. (2014). Carbon as catalyst and support for
nanostructures. Nat. Rev. Mater. 3, 17089. electrochemical energy conversion. Carbon
35. Geng, X., Sun, W., Wu, W., Chen, B., Al-Hilo, 75, 5–42.
21. Li, C., Cao, Q., Wang, F., Xiao, Y., Li, Y., A., Benamara, M., Zhu, H., Watanabe, F., Cui,
Delaunay, J.-J., and Zhu, H. (2018). J., and Chen, T.-P. (2016). Pure and stable 48. Ma, W., Li, H., Jiang, S., Han, G., Gao, J., Yu,
Engineering graphene and TMDs based van metallic phase molybdenum disulfide X., Lian, H., Tu, W., Han, Y.-f., and Ma, R.
der Waals heterostructures for photovoltaic nanosheets for hydrogen evolution reaction. (2018). Facile synthesis of superstructured
and photoelectrochemical solar energy Nat. Commun. 7, 10672. MoS2 and graphitic nanocarbon hybrid for
conversion. Chem. Soc. Rev. 47, 4981–5037. efficient hydrogen evolution reaction. ACS
36. Wang, D., Zhang, X., Bao, S., Zhang, Z., Fei, Sustain. Chem. Eng. 6, 14441–14449.
22. Voiry, D., Shin, H.S., Loh, K.P., and Chhowalla, H., and Wu, Z. (2017). Phase engineering of a
M. (2018). Low-dimensional catalysts for multiphasic 1T/2H MoS2 catalyst for highly 49. Li, Y., Hao, S., Murthy, A.A., Distefano, J.G.,
hydrogen evolution and CO2 reduction. Nat. efficient hydrogen evolution. J. Mater. Chem. Hanson, E.D., Xu, Y., Wolverton, C., Dravid,
Rev. Chem. 2, 0105. A 5, 2681–2688. V.P., Chen, X., Majewski, M.B., et al. (2018).
Morphological engineering of winged
23. Deng, D., Novoselov, K.S., Fu, Q., Zheng, N., 37. Doan, M.-H., Jin, Y., Adhikari, S., Lee, S., Zhao, Au@MoS2 heterostructures for
Tian, Z., and Bao, X. (2016). Catalysis with two- J., Lim, S.C., and Lee, Y.H. (2017). Charge electrocatalytic hydrogen evolution. Nano
dimensional materials and their transport in MoS2/WSe2 van der Waals Lett. 18, 7104–7110.
heterostructures. Nat. Nanotechnol. 11, heterostructure with tunable inversion layer.
218–230. ACS Nano 11, 3832–3840. 50. Yang, J., Zhu, J., Xu, J., Zhang, C., and Liu, T.
(2017). MoSe2 nanosheet array with layered
24. Liu, Y.-T., Zhu, X.-D., and Pan, L. (2018). Hybrid 38. Sun, Y., Alimohammadi, F., Zhang, D., and MoS2 heterostructures for superior hydrogen
architectures based on 2D MXenes and low- Guo, G. (2017). Enabling colloidal synthesis of evolution and lithium storage performance.
dimensional inorganic nanostructures: edge-oriented MoS2 with expanded ACS Appl. Mater. Interfaces 9, 44550–44559.
methods, synergies, and energy-related interlayer spacing for enhanced HER catalysis.
applications. Small 14, 51. Nano Lett. 17, 1963–1969. 51. Shizheng, Z., Lijun, Z., Zhengyou, Z., Jian, C.,
Jianli, K., Zhulin, H., and Dachi, Y. (2018). MoS2
25. Schapbach, L., and Zuttel, A. (2001). 39. Feng, J.X., Ye, S.H., He, X.J., Xu, H., Tong, nanosheet arrays rooted on hollow rGO
Hydrogen-storage materials for mobile Y.X., Li, G.R., and Ding, L.X. (2015). spheres as bifunctional hydrogen evolution
applications. Nature 414, 353. Co(OH)2@PANI hybrid nanosheets with 3D catalyst and supercapacitor electrode. Nano
networks as high-performance Micro Lett. 10, 1.
26. Kreuter, W., and Hofmann, H. (1998). electrocatalysts for hydrogen evolution
Electrolysis: the important energy transformer reaction. Adv. Mater. 27, 7051–7057. 52. Du, C., Liang, D., Shang, M., Zhang, J., Mao,
in a world of sustainable energy. Int. J. J., Liu, P., and Song, W. (2018). In situ
40. Seh, Z.W., Kibsgaard, J., Dickens, C.F.,
Hydrogen Energy 23, 661–666. engineering MoS2 NDs/VS2 lamellar
Chorkendorff, I.B., Norskov, J.K., and heterostructure for enhanced electrocatalytic
27. Zhao, G., Rui, K., Dou, S.X., and Sun, W. (2018). Jaramillo, T.F. (2017). Combining theory and hydrogen evolution. ACS Sustain. Chem. Eng.
Heterostructures for electrochemical experiment in electrocatalysis: insights into 6, 15471–15479.
hydrogen evolution reaction: a review. Adv. materials design. Science 355, 146.
Funct. Mater. 28, 1803291. 41. Li, H., Tsai, C., Koh, A.L., Cai, L., Contryman,
53. Vikraman, D., Hussain, S., Akbar, K., Truong,
L., Kathalingam, A., Chun, S.-H., Jung, J., Park,
A.W., Fragapane, A.H., Zhao, J., Han, H.S.,
28. Dincer, I., and Acar, C. (2015). Review and H.J., and Kim, H.-S. (2018). Improved
Manoharan, H.C., Abild-Pedersen, F., et al.
evaluation of hydrogen production methods hydrogen evolution reaction performance
(2016). Activating and optimizing MoS2 basal
for better sustainability. Int. J. Hydrogen using MoS2-WS2 heterostructures by
planes for hydrogen evolution through the
Energy 40, 11094–11111. physicochemical process. ACS Sustain.
formation of strained sulphur vacancies. Nat.
Chem. Eng. 6, 8400–8409.
29. Nagpal, M., and Kakkar, R. (2018). An evolving Mater. 15, 48–53.
energy solution: intermediate hydrogen 54. Wang, C., Wang, H., Lin, Z., Li, W., Lin, B., Qiu,
42. Gan, X., Lv, R., Wang, X., Zhang, Z., Fujisawa,
storage. Int. J. Hydrogen Energy 43, 12168– K., Lei, Y., Huang, Z.-H., Terrones, M., and W., Quan, Y., Liu, Z., and Chen, S. (2019). In-
12188. situ synthesis of edge-enriched MoS2
Kang, F. (2018). Pyrolytic carbon supported
hierarchical nanorods with 1T/2H hybrid
alloying metal dichalcogenides as free-
30. McCrory, C.C.L., Jung, S., Ferrer, I.M., phases for highly efficient electrocatalytic
standing electrodes for efficient hydrogen
Chatman, S.M., Peters, J.C., and Jaramillo, hydrogen evolution. CrystEngComm 21,
evolution. Carbon 132, 512–519.
T.F. (2015). Benchmarking hydrogen evolving 1984–1991.
reaction and oxygen evolving reaction 43. Gong, Q., Cheng, L., Liu, C., Zhang, M., Feng,
electrocatalysts for solar water splitting Q., Ye, H., Zeng, M., Xie, L., Liu, Z., and Li, Y. 55. Lei, Y., Pulickal Rajukumar, L., Zhou, C.,
devices. J. Am. Chem. Soc. 137, 4347–4357. (2015). Ultrathin MoS2(1-x)Se2x alloy nanoflakes Kabius, B., Alem, N., Terrones, M., Pakhira, S.,
for electrocatalytic hydrogen evolution Iyiola, O.O., Mendoza-Cortes, J.L., Fujisawa,
31. Wang, P., Zhang, J., Yao, J., Huang, X., Zhang, reaction. ACS Catal. 5, 2213–2219. K., et al. (2017). Low-temperature synthesis of
X., Lu, G., Wan, S., and Guo, S. (2017). Precise heterostructures of transition metal
tuning in platinum-nickel/nickel sulfide 44. Zhang, J., Liu, S., Dong, R., Zhuang, X., Feng, dichalcogenide alloys (WxMo1-xS2) and
interface nanowires for synergistic hydrogen X., Wang, T., Liu, P., and Chen, M. (2016). graphene with superior catalytic performance
evolution catalysis. Nat. Commun. 8, 14580. Engineering water dissociation sites in MoS2 for hydrogen evolution. ACS Nano 11, 5103–
nanosheets for accelerated electrocatalytic 5112.
32. Lim, J., Park, D., Jeon, S.S., Roh, C.-W., Choi, hydrogen production. Energy Environ. Sci. 9,
J., Yoon, D., Park, M., Jung, H., and Lee, H. 2789–2793. 56. Ravula, S., Essner, J.B., Robertson, J.D., Baker,
(2018). Ultrathin IrO2 nanoneedles for G.A., Zhang, C., and Lin, J. (2017). Ionic liquid-
electrochemical water oxidation. Adv. Funct. 45. Ouyang, C., Wang, X., and Wang, S. (2015). assisted synthesis of nanoscale
Mater. 28, 1704796. Phosphorus-doped CoS2 nanosheet arrays as (MoS2)x(SnO2)1-x on reduced graphene oxide
ultra-efficient electrocatalysts for the for the electrocatalytic hydrogen evolution
33. Li, D.N., Wang, A.J., Feng, J.J., Wei, J., and hydrogen evolution reaction. Chem. reaction. ACS Appl. Mater. Interfaces 9, 8065–
Zhang, Q.L. (2018). Dentritic platinum- Commun. (Camb.) 51, 14160–14163. 8074.
palladium/palladium core-shell nanocrystals/
reduced graphene oxide: one-pot synthesis 46. Wang, S., Zhang, D., Li, B., Zhang, C., Du, Z., 57. Chia, X., Eng, A.Y.S., Ambrosi, A., Tan, S.M.,
and excellent electrocatalytic performances. Yin, H., Bi, X., and Yang, S. (2018). Ultrastable and Pumera, M. (2015). Electrochemistry of
J. Colloid Interface Sci. 514, 93–101. in-plane 1T-2H MoS2 heterostructures for nanostructured layered transition-metal

Matter 2, 526–553, March 4, 2020 549


dichalcogenides. Chem. Rev. 115, 11941– evolution reaction: a density functional study. metal atom doping of the basal plane of MoS2
11966. Phys. Chem. Chem. Phys. 16, 13156–13164. monolayer nanosheets for electrochemical
hydrogen evolution. Chem. Sci. 9, 4769–4776.
58. Kong, D., Wang, H., Cha, J.J., Pasta, M., Koski, 71. Chen, S., Chen, X., Wang, G., Liu, L., He, Q., Li,
K.J., Yao, J., and Cui, Y. (2013). Synthesis of X.-B., and Cui, N. (2018). Reaction mechanism 84. Xiong, Q., Wang, Y., Liu, P.-F., Zheng, L.-R.,
MoS2 and MoSe2 films with vertically aligned with thermodynamic structural screening for Wang, G., Yang, H.-G., Wong, P.-K., Zhang,
layers. Nano Lett. 13, 1341–1347. electrochemical hydrogen evolution on H., and Zhao, H. (2018). Cobalt covalent
monolayer 1T0 phase MoS2. Chem. Mater. 30, doping in MoS2 to induce bifunctionality of
59. Merki, D., and Hu, X. (2011). Recent 5404–5411. overall water splitting. Adv. Mater. 30,
developments of molybdenum and tungsten 1801450.
sulfides as hydrogen evolution catalysts. 72. Lukowski, M.A., Daniel, A.S., Meng, F.,
Energy Environ. Sci. 4, 3878–3888. Forticaux, A., Li, L., and Jin, S. (2013). 85. Xiong, Q., Zhang, X., Wang, H., Liu, G., Wang,
Enhanced hydrogen evolution catalysis from G., Zhang, H., and Zhao, H. (2018). One-step
60. Yang, Y., Fei, H., Ruan, G., Li, Y., and Tour, chemically exfoliated metallic MoS2 synthesis of cobalt-doped MoS2 nanosheets
J.M. (2015). Vertically aligned WS2 nanosheets nanosheets. J. Am. Chem. Soc. 135, 10274– as bifunctional electrocatalysts for overall
for water splitting. Adv. Funct. Mater. 25, 10277. water splitting under both acidic and alkaline
6199–6204. conditions. Chem. Commun. (Camb.) 54,
73. Li, J., Zhao, Z., Ma, Y., and Qu, Y. (2017). 3859–3862.
61. Ouyang, Y., Ling, C., Chen, Q., Wang, Z., Shi, Graphene and their hybrid electrocatalysts for
L., and Wang, J. (2016). Activating inert basal water splitting. Chemcatchem 9, 1554. 86. Luo, Z., Xiao, M., Ge, J., Liu, C., Xing, W.,
planes of MoS2 for hydrogen evolution Ouyang, Y., Wang, J., Zhang, H., Jiang, Z.,
reaction through the formation of different 74. Wang, H., Ouyang, L., Zou, G., Sun, C., Hu, J., Tang, D., and Cao, X. (2018). Chemically
intrinsic defects. Chem. Mater. 28, 4390–4396. Xiao, X., and Gao, L. (2018). Optimizing MoS2 activating MoS2 via spontaneous atomic
edges by alloying isovalent W for robust palladium interfacial doping towards efficient
62. Hinnemann, B., Moses, P.G., Bonde, J., hydrogen evolution activity. ACS Catal. 8, hydrogen evolution. Nat. Commun. 9, 2120.
Jørgensen, K.P., Nielsen, J.H., Horch, S., 9529–9536.
Chorkendorff, I., and Nørskov, J.K. (2005). 87. Yang, S.-Z., Gong, Y., Manchanda, P., Zhang,
Biomimetic hydrogen evolution: MoS2 75. Chi, J.-Q., Gao, W.-K., Lin, J.-H., Dong, B., Y.-Y., Ye, G., Chen, S., Song, L., Pantelides,
nanoparticles as catalyst for hydrogen Yan, K.-L., Qin, J.-F., Liu, B., Chai, Y.-M., and S.T., Ajayan, P.M., Chisholm, M.F., and Zhou,
evolution. J. Am. Chem. Soc. 127, 5308–5309. Liu, C.-G. (2019). N, P dual-doped hollow W. (2018). Rhenium-doped and stabilized
carbon spheres supported MoS2 hybrid MoS2 atomic layers with basal-plane catalytic
63. Thomas, F.J., Kristina, P.J., Jacob, B., Jane, electrocatalyst for enhanced hydrogen activity. Adv. Mater. 30, 1803477.
H.N., Sebastian, H., and Ib, C. (2007). evolution reaction. Catal. Today 330, 259–267.
Identification of active edge sites for 88. Huang, Y., Nielsen, R.J., and Goddard, W.A.,
electrochemical H2 evolution from MoS2 76. Wang, Y., Liu, S., Hao, X., Zhou, J., Song, D., III (2018). Reaction mechanism for the
nanocatalysts. Science 317, 100. Wang, D., Hou, L., and Gao, F. (2017). hydrogen evolution reaction on the basal
Fluorine- and nitrogen-codoped MoS2with a plane sulfur vacancy site of MoS2 using grand
64. Wang, G., Zhang, J., Yang, S., Wang, F., catalytically active basal plane. ACS Appl. canonical potential kinetics. J. Am. Chem.
Zhuang, X., Mullen, K., and Feng, X. (2018). Mater. Interfaces 9, 27715–27719. Soc. 140, 16773–16782.
Vertically aligned MoS2 nanosheets patterned
on electrochemically exfoliated graphene for 77. Liu, P., Zhu, J., Zhang, J., Xi, P., Tao, K., Gao, 89. Li, L., Qin, Z., Ries, L., Hong, S., Michel, T.,
high-performance lithium and sodium D., and Xue, D. (2017). P dopants triggered Yang, J., Salameh, C., Bechelany, M., Miele,
storage. Adv. Energy Mater. 8, 1702254. new basal plane active sites and enlarged P., Kaplan, D., et al. (2019). Role of sulfur
interlayer spacing in MoS2 nanosheets toward vacancies and undercoordinated Mo regions
65. Liu, Z., Zhao, L., Liu, Y., Gao, Z., Yuan, S., Li, X., electrocatalytic hydrogen evolution. ACS in MoS2 nanosheets toward the evolution of
Li, N., and Miao, S. (2019). Vertical nanosheet Energy Lett. 2, 745–752. hydrogen. ACS Nano 13, 6824–6834.
array of 1T phase MoS2 for efficient and stable
hydrogen evolution. Appl. Catal. B Environ. 78. Wei, Y., Zhang, X., Zhao, Z., Chen, H.-S., 90. Tang, B., Yu, Z.G., Seng, H.L., Zhang, N., Liu,
246, 296–302. Matras-Postolek, K., Wang, B., and Yang, P. X., Zhang, Y.-W., Yang, W., and Gong, H.
(2019). Controllable synthesis of P-doped (2018). Simultaneous edge and electronic
66. Zhu, P., Chen, Y., Zhou, Y., Yang, Z., Wu, D., MoS2 nanopetals decorated N-doped hollow control of MoS2 nanosheets through Fe
Xiong, X., and Ouyang, F. (2018). A metallic carbon spheres towards enhanced hydrogen doping for an efficient oxygen evolution
MoS2 nanosheet array on graphene- evolution. Electrochim. Acta 297, 553–563. reaction. Nanoscale 10, 20113–20119.
protected Ni foam as a highly efficient
electrocatalytic hydrogen evolution cathode. 79. Zhu, T., Ding, J., Shao, Q., Qian, Y., and 91. Tang, W., Wang, X., Zhong, Y., Xie, D., Zhang,
J. Mater. Chem. A 6, 16458–16464. Huang, X. (2019). P,Se-Codoped MoS2 X., Xia, X., Wu, J., Gu, C., and Tu, J. (2018).
nanosheets as accelerated electrocatalysts for Hierarchical MoS2/carbon composite
67. Voiry, D., Yamaguchi, H., Li, J., Silva, R., Alves, hydrogen evolution. Chemcatchem 11, 689. microspheres as advanced anodes for
D.C.B., Fujita, T., Chen, M., Asefa, T., Shenoy, lithium/sodium-ion batteries. Chemistry 24,
V., Eda, G., and Chhowalla, M. (2013). 80. Zhang, H., Yu, L., Chen, T., Zhou, W., and Lou, 11220–11226.
Enhanced catalytic activity in strained X.W. (2018). Surface modulation of
chemically exfoliated WS2 nanosheets for hierarchical MoS2 nanosheets by Ni single 92. Larcher, D., and Tarascon, J.M. (2015).
hydrogen evolution. Nat. Mater. 12, 850–855. atoms for enhanced electrocatalytic Towards greener and more sustainable
hydrogen evolution. Adv. Funct. Mater. 28, batteries for electrical energy storage. Nat.
68. Ma, Y., Liu, B., Zhang, A., Chen, L., Fathi, M., 1807086. Chem. 7, 19–29.
Shen, C., Abbas, A.N., Ge, M., Mecklenburg,
M., and Zhou, C. (2015). Reversible 81. Kim, M., Anjum, M.A.R., Lee, M., Lee, B.J., and 93. Ko, M., Chae, S., and Cho, J. (2015).
semiconducting-to-metallic phase transition Lee, J.S. (2019). Activating MoS2 basal plane Challenges in accommodating volume
in chemical vapor deposition grown mono with Ni2P nanoparticles for Pt-like hydrogen change of Si anodes for Li-ion batteries.
layer WSe2 and applications for devices. ACS evolution reaction in acidic media. Adv. Funct. ChemElectroChem 2, 1645–1651.
Nano 9, 7383–7391. Mater. 29, 1809151.
94. Wen, L., Li, F., and Cheng, H.-M. (2016).
69. Yin, Y., Zhang, Y., Gao, T., Yao, T., Zhang, X., 82. Wang, Q., Zhao, Z.L., Dong, S., He, D., Carbon nanotubes and graphene for flexible
Han, J., Wang, X., Zhang, Z., Xu, P., Zhang, P., Lawrence, M.J., Han, S., Cai, C., Xiang, S., electrochemical energy storage: from
et al. (2017). Synergistic phase and disorder Rodriguez, P., Xiang, B., et al. (2018). Design materials to devices. Adv. Mater. 28, 4306–
engineering in 1T-MoSe2 nanosheets for of active nickel single-atom decorated MoS2 4337.
enhanced hydrogen-evolution reaction. Adv. as a pH-universal catalyst for hydrogen
Mater. 29, 1700311. evolution reaction. Nano Energy 53, 458–467. 95. Du, J., Pei, S., Ma, L., and Cheng, H.-M. (2014).
25th anniversary article: carbon nanotube-
70. Tsai, C., Chan, K., Abild-Pedersen, F., and 83. Lau, T.H.M., Lu, X., Kulhavý, J., Wu, S., Foord, and graphene-based transparent conductive
Norskov, J.K. (2014). Active edge sites in J.S., Tsang, S.C.E., Lu, L., Wu, T.S., Soo, Y.L., films for optoelectronic devices. Adv. Mater.
MoSe2 and WSe2 catalysts for the hydrogen Kato, R., and Suenaga, K. (2018). Transition 26, 1958–1991.

550 Matter 2, 526–553, March 4, 2020


96. Yang, M., Ko, S., Im, J.S., and Choi, B.G. heterostructure for superior lithium storage. graphene sheets for high-performance
(2015). Free-standing molybdenum disulfide/ Carbon 133, 162–169. sodium-ion batteries. Adv. Energy Mater. 8,
graphene composite paper as a binder- and 1703300.
carbon-free anode for lithium-ion batteries. 111. Chen, M., Wang, J., Yan, X., Ren, J., Dai, Y.,
J. Power Source 288, 76–81. Wang, Q., Wang, Y., and Cheng, X. (2017). 123. Cao, J., Jafta, C.J., Gong, J., Rang, Q., Lin, X.,
Flower-like molybdenum disulfide Felix, R., Wilks, R.G., Baer, M., Yuan, J.,
97. Li, L., Wu, Z., Yuan, S., and Zhang, X.-B. (2014). nanosheets grown on carbon nanosheets to Ballauff, M., and Lu, Y. (2016). Synthesis of
Advances and challenges for flexible energy form nanocomposites: novel structure and dispersible mesoporous nitrogen-doped
storage and conversion devices and systems. excellent electrochemical performance. hollow carbon nanoplates with uniform
Energy Environ. Sci. 7, 2101–2122. J. Alloys Compd. 722, 250–258. hexagonal morphologies for supercapacitors.
ACS Appl. Mater. Interfaces 8, 29628–29636.
98. González, A., Goikolea, E., Barrena, J.A., and 112. Quan, T., Goubard-Bretesche, N., Hark, E.,
Mysyk, R. (2016). Review on supercapacitors: Kochovski, Z., Mei, S., Pinna, N., Ballauff, M., 124. Mao, H., Shi, L., Song, S., Xiao, C., Liang, J.,
technologies and materials. Renew. Sustain. and Lu, Y. (2019). Highly dispersible Dong, B., and Ding, S. (2017). N-Doped
Energy Rev. 58, 1189–1206. hexagonal carbon-MoS2-carbon nanoplates hollow carbon nanosheet supported SnO2
with hollow sandwich structures for nanoparticles. Inorg. Chem. Front. 4, 1742–
99. Kang, B., and Ceder, G. (2009). Battery supercapacitors. Chemistry 25, 4757. 1747.
materials for ultrafast charging and
discharging. Nature 458, 190–193. 113. Wang, S., Zhu, J., Shao, Y., Li, W., Wu, Y., 125. Zhu, J., Sun, W., Yang, D., Zhang, Y., Hoon,
Zhang, L., and Hao, X. (2017). Three- H.H., Zhang, H., and Yan, Q. (2015).
100. Simon, P., and Gogotsi, Y. (2008). Materials for dimensional MoS2@CNT/RGO network Multifunctional architectures constructing of
electrochemical capacitors. Nat. Mater. 7, composites for high-performance flexible PANI nanoneedle arrays on MoS2 thin
845–854. supercapacitors. Chemistry 23, 3438. nanosheets for high-energy supercapacitors.
Small 11, 4123–4129.
101. Wu, S., Du, Y., and Sun, S. (2017). Transition 114. Gigot, A., Fontana, M., Serrapede, M.,
metal dichalcogenide based nanomaterials Castellino, M., Bianco, S., Armandi, M., 126. Feng, F., Wu, J., Wu, C., and Xie, Y. (2015).
for rechargeable batteries. Chem. Eng. J. 307, Bonelli, B., Pirri, C.F., Tresso, E., and Rivolo, P. Regulating the electrical behaviors of 2D
189–207. (2016). Mixed 1T-2H phase MoS2/reduced inorganic nanomaterials for energy
graphene oxide as active electrode for applications. Small 11, 654–666.
102. Yu, Z., Tetard, L., Zhai, L., and Thomas, J. enhanced supercapacitive performance. ACS
(2015). Supercapacitor electrode materials: Appl. Mater. Interfaces 8, 32842–32852. 127. Wang, Y., Ma, Z., Chen, Y., Zou, M., Yousaf,
nanostructures from 0 to 3 dimensions. M., Yang, Y., Yang, L., Cao, A., and Han, R.P.S.
Energy Environ. Sci. 8, 702–730. 115. Kumar, N.A., Dar, M.A., Gul, R., and Baek, (2016). Controlled synthesis of core-shell
J.-B. (2015). Graphene and molybdenum carbon@MoS2 nanotube sponges as high-
103. da Silveira Firmiano, E.G., Rabelo, A.C., disulfide hybrids: synthesis and applications. performance battery electrodes. Adv. Mater.
Dalmaschio, C.J., Pinheiro, A.N., Pereira, E.C., Mater. Today 18, 286–298. 28, 10175–10181.
Schreiner, W.H., and Leite, E.R. (2014).
Supercapacitor electrodes obtained by 116. Kamila, S., Mohanty, B., Samantara, A.K., 128. Fang, Y., Lv, Y., Gong, F., Elzatahry, A.A.,
directly bonding 2D MoS2 on reduced Mishra, B.K., Jena, B.K., Guha, P., Ghosh, A., Zheng, G., and Zhao, D. (2016). Synthesis of
graphene oxide. Adv. Energy Mater. 4, Satyam, P.V., and Jena, B. (2017). Highly active 2D-mesoporous-carbon/MoS2
1301380. 2D layered MoS2-rGO hybrids for energy heterostructures with well-defined interfaces
conversion and storage applications. Sci. Rep. for high-performance lithium-ion batteries.
104. Yang, Y., Fei, H., Ruan, G., Xiang, C., and Tour, 7, 8378. Adv. Mater. 28, 9385–9390.
J.M. (2014). Edge-oriented MoS2 nanoporous
films as flexible electrodes for hydrogen 117. Yang, W., He, L., Tian, X., Yan, M., Yuan, H., 129. Yuan, W., Zhang, Y., Cheng, L., Wu, H., Zheng,
evolution reactions and supercapacitor Liao, X., Meng, J., Hao, Z., and Mai, L. (2017). L., and Zhao, D. (2016). The applications of
devices. Adv. Mater. 26, 8163–8168. Carbon-MEMS-based alternating stacked carbon nanotubes and graphene in advanced
MoS2@rGO-CNT micro-supercapacitor with rechargeable lithium batteries. J. Mater.
105. Xu, X., Zhao, R., Ai, W., Chen, B., Du, H., Wu, high capacitance and energy density. Small Chem. A 4, 8932–8951.
L., Zhang, H., Huang, W., and Yu, T. (2018). 13, 1700639.
Controllable design of MoS2 nanosheets 130. Wang, X., Weng, Q., Yang, Y., Bando, Y., and
anchored on nitrogen-doped graphene: 118. Yang, B., Hao, C., Wen, F., Wang, B., Mu, C., Gotberg, D. (2016). Hybrid two-dimensional
toward fast sodium storage by tunable Xiang, J., Li, L., Xu, B., Zhao, Z., Liu, Z., and materials in rechargeable battery applications
pseudocapacitance. Adv. Mater. 30, 1800658. Tian, Y. (2017). Flexible black-phosphorus and their microscopic mechanisms. Chem.
nanoflake/carbon nanotube composite paper Soc. Rev. 45, 4042–4073.
106. Pan, Y., Zhang, J., and Lu, H. (2017). Uniform for high-performance all-solid-state
yolk–shell MoS2@Carbon microsphere supercapacitors. ACS Appl. Mater. Interfaces 131. Zhang, Z., Xiao, F., Xiao, J., Wang, S., Qian, L.,
anodes for high-performance lithium-ion 9, 44478–44484. and Liu, Y. (2014). Facile synthesis of 3D
batteries. Chemistry 23, 9937–9945. MnO2-graphene and carbon nanotube-
119. Sun, T., Liu, X., Li, Z., Ma, L., Wang, J., and graphene composite networks for high-
107. Wu, M., Xia, S., Ding, J., Zhao, B., Jiao, Y., Du, Yang, S. (2017). Graphene-wrapped performance, flexible, all-solid-state
A., and Zhang, H. (2018). Growth of MoS2 CNT@MoS2 hierarchical structure: synthesis, asymmetric supercapacitors. Adv. Energy
nanoflowers with expanded interlayer characterization and electrochemical Mater. 4, 1400064.
distance onto N-doped graphene for application in supercapacitors. New J. Chem.
reversible lithium storage. ChemElectroChem 41, 7142–7150. 132. Zhang, Z., Chi, K., Xiao, F., and Wang, S.
5, 2263–2270. (2015). Advanced solid-state asymmetric
120. Wang, X., Ding, J., Yao, S., Wu, X., Feng, Q., supercapacitors based on 3D graphene/
108. Geng, X., Jiao, Y., Han, Y., Mukhopadhyay, A., Wang, Z., and Geng, B. (2014). High MnO2 and graphene/polypyrrole hybrid
Yang, L., and Zhu, H. (2017). Freestanding supercapacitor and adsorption behaviors of architectures. J. Mater. Chem. A 3, 12828–
metallic 1T MoS2 with dual ion diffusion paths flower-like MoS2 nanostructures. J. Mater. 12835.
as high rate anode for sodium-ion batteries. Chem. A 2, 15958–15963.
Adv. Funct. Mater. 27, 1702998. 133. Guo, X., Zheng, S., Zhang, G., Xiao, X., Li, X.,
121. Ren, L., Zhang, G., Yan, Z., Kang, L., Xu, H., Shi, Xu, Y., Xue, H., and Pang, H. (2017).
109. Jiang, Y., Wu, Y., Chen, Y., Qi, Z., Shi, J., Gu, L., F., Lei, Z., and Liu, Z.-H. (2015). Three- Nanostructured graphene-based materials
and Yu, Y. (2018). Design nitrogen (N) and dimensional tubular MoS2/PANI hybrid for flexible energy storage. Energy Storage
sulfur (S) Co-doped 3D graphene network electrode for high rate performance Mater. 9, 150–169.
architectures for high-performance sodium supercapacitor. ACS Appl. Mater. Interfaces
storage. Small 14, 1703471. 7, 28294–28302. 134. Lv, W., Li, Z., Deng, Y., Yang, Q.-H., and Kang,
F. (2016). Graphene-based materials for
110. Zheng, X., Wang, S., Xiong, C., and Hu, G. 122. Li, P., Jeong, J.Y., Jin, B., Zhang, K., and Park, electrochemical energy storage devices:
(2018). In situ growth of 1T-MoS2 on liquid- J.H. (2018). Vertically oriented MoS2 with opportunities and challenges. Energy Storage
exfoliated graphene: a unique graphene-like spatially controlled geometry on nitrogenous Mater. 2, 107–138.

Matter 2, 526–553, March 4, 2020 551


135. Yao, X., and Zhao, Y. (2017). Three- supercapacitors and hydrogen evolution MoS2 monolayer. J. Phys. Chem. C 118, 1515–
dimensional porous graphene networks and reaction. New J. Chem. 43, 3601–3608. 1522.
hybrids for lithium-ion batteries and
supercapacitors. Chem 2, 171–200. 149. Fu, G., and Lee, J.-M. (2019). Ternary metal 161. Zhang, J., Wang, T., Liu, P., Liu, Y., Ma, J., and
sulfides for electrocatalytic energy Gao, D. (2016). Enhanced catalytic activities of
136. El-Kady, M.F., Shao, Y., and Kaner, R.B. (2016). conversion. J. Mater. Chem. A 7, 9386–9405. metal-phase-assisted 1T@2H-MoSe2
Graphene for batteries, supercapacitors and nanosheets for hydrogen evolution.
beyond. Nat. Rev. Mater. 1, 16033. 150. Fu, G., Wang, J., Chen, Y., Liu, Y., Tang, Y., Electrochim. Acta 217, 181–186.
Goodenough, J.B., and Lee, J.-M. (2018).
137. Fu, G., Tang, Y., and Lee, J.-M. (2018). Recent Exploring indium-based ternary thiospinel as 162. Ye, L., Chen, S., Li, W., Pi, M., Wu, T., and
advances in carbon-based bifunctional conceivable high-potential air-cathode for Zhang, D. (2015). Tuning the electrical
oxygen electrocatalysts for Zn-Air batteries. rechargeable Zn-air batteries. Adv. Energy transport properties of multilayered
ChemElectroChem 5, 1424–1434. Mater. 8, 1802263. molybdenum disulfide nanosheets by
intercalating phosphorus. J. Phys. Chem. C
138. Yun, Q., Lu, Q., Zhang, X., Tan, C., and Zhang, 151. Yan, Y., Li, K., Chen, X., Yang, Y., and Lee, 119, 9560–9567.
H. (2018). Three-dimensional architectures J.-M. (2017). Heterojunction-assisted
constructed from transition-metal Co3S4@Co3O4 core-shell octahedrons for 163. Ekspong, J., Sandstrom, R., Rajukumar, L.P.,
dichalcogenide nanomaterials for supercapacitors and both oxygen and carbon Terrones, M., Wagberg, T., and Gracia-
electrochemical energy storage and dioxide reduction reactions. Small 13, Espino, E. (2018). Stable sulfur-intercalated
conversion. Angew. Chem. Int. Ed. 57, 1701724. 1T’MoS2 on graphitic nanoribbons as
626–646.
hydrogen evolution electrocatalyst. Adv.
152. Huang, F., Yan, A., Sui, Y., Wei, F., Qi, J., Funct. Mater. 28, 1802744.
139. Jayakumar, A., Antony, R.P., Wang, R., and Meng, Q., and He, Y. (2017). One-step
Lee, J.-M. (2017). MOF-derived hollow cage hydrothermal synthesis of Ni3S4@MoS2
NixCo3-xO4 and their synergy with graphene 164. Chang, K., Hai, X., Pang, H., Zhang, H., Shi, L.,
nanosheet on carbon fiber paper as a Liu, G., Liu, H., Zhao, G., Li, M., and Ye, J.
for outstanding supercapacitors. Small 13, binder-free anode for supercapacitor.
1603102. (2016). Targeted synthesis of 2H- and 1T-
J. Mater. Sci. Mater. Electron. 28, 12747– phase MoS2 monolayers for catalytic
12754. hydrogen evolution. Adv. Mater. 28, 10033–
140. Xu, X., Wu, L., Sun, Y., Wang, T., Chen, X.,
Wang, Y., Zhong, W., and Du, Y. (2018). High- 10041.
153. Lai, F., Zhou, G., Li, F., Miao, Y.E., Liu, T.,
rate, flexible all-solid-state super-capacitor Huang, Y., He, Z., Yong, D., Bai, W., Pan, B.,
based on porous aerogel hybrids of MoS2/ 165. Chia, X., Ambrosi, A., Sedmidubsk, D., Sofer,
and Tjiu, W.W. (2018). Highly dual- Z., and Pumera, M. (2014). Precise tuning of
reduced graphene oxide. J. Electroanal. heteroatom-doped ultrathin carbon
Chem. 811, 96–104. the charge transfer kinetics and catalytic
nanosheets with expanded interlayer distance properties of MoS2 materials via
for efficient energy storage. ACS Sustain. electrochemical methods. Chemistry 20,
141. Anjali, J., Jose, V.K., and Lee, J.-M. (2019).
Chem. Eng. 6, 3143–3153. 17426–17432.
Carbon-based hydrogels: synthesis and their
recent energy applications. J. Mater. Chem. A 154. Xiao, H., Wang, S., Zhang, S., Wang, Y., Xu, Q.,
7, 15491–15518. Hu, W., Zhou, Y., Wang, Z., An, C., and Zhang,
166. Yang, H., Kim, S.W., Chhowalla, M., and Lee,
Y.H. (2017). Erratum: Structural and quantum-
142. Tian, J., Zhang, H., and Li, Z. (2018). Synthesis J. (2017). Interlayer expanded molybdenum
state phase transitions in van der Waals
of double-layer nitrogen-doped microporous disulfide nanosheets assembly for
layered materials (vol 13, pg 931, 2017). Nat.
electrochemical supercapacitor with
hollow carbon@MoS2/MoO2 nanospheres for Phys. 13, 1232.
supercapacitors. ACS Appl. Mater. Interfaces enhanced performance. Mater. Chem. Phys.
10, 29511–29520. 192, 100–107. 167. Tang, Q. (2018). Tuning the phase stability of
155. Liu, Y., Wang, X., Song, X., Dong, Y., Mo-based TMD monolayers through coupled
143. Zhang, T., Kong, L.-B., Liu, M.-C., Dai, Y.-H., vacancy defects and lattice strain. J. Mater.
Yan, K., Hu, B., Luo, Y.-C., and Kang, L. (2016). Yang, L., Wang, L., Jia, D., Zhao, Z., and
Qiu, J. (2016). Interlayer expanded MoS2 Chem. C 6, 9561.
Design and preparation of MoO2/MoS2 as
negative electrode materials for enabled by edge effect of graphene
168. Ke, Q., Zhang, X., Zang, W., Elshahawy, A.M.,
supercapacitors. Mater. Des. 112, 88–96. nanoribbons for high performance lithium
Hu, Y., He, Q., Pennycook, S.J., Cai, Y., and
and sodium ion batteries. Carbon 109,
Wang, J. (2019). Strong charge transfer at 2H-
144. Zhang, H.-J., Wu, T.-H., Wang, K.-X., Wu, 461–471.
1T phase boundary of MoS2 for superb high-
X.-Y., Chen, X.-T., Jiang, Y.-M., Wei, X., and performance energy storage. Small 15,
Chen, J.-S. (2013). Uniform hierarchical MoO2/ 156. Ji, Y., Wei, Q., and Su, Y. (2018). Superior
capacitive performance enabled by edge- 1900131.
carbon spheres with high cycling performance
for lithium ion batteries. J. Mater. Chem. A 1, oriented and interlayer-expanded MoS2
nanosheets anchored on reduced graphene 169. Jin, H., Li, J., Wei, Y., Dai, Y., and Guo, H.
12038–12043. (2018). Unraveling the mechanism of
oxide sheets. Ind. Eng. Chem. Res. 57, 4571–
4576. photoinduced charge-transfer process in
145. Sari, F.N.I., and Ting, J.-M. (2018). MoS2/
bilayer heterojunction. ACS Appl. Mater.
MoOx-nanostructure-decorated activated
157. Chao, J., Yang, L., Liu, J., Hu, R., and Zhu, M. Interfaces 10, 25401–25408.
carbon cloth for enhanced supercapacitor
performance. ChemSusChem 11, 897–906. (2018). Sandwiched MoS2/polyaniline
nanosheets array vertically aligned on 170. Vikraman, D., Hussain, S., Prasanna, K.,
146. Vattikuti, S.V.P., Nagajyothi, P.C., Anil Kumar reduced graphene oxide for high Karuppasamy, K., Jung, J., and Kim, H.-S.
Reddy, P., Kotesh Kumar, M., Shim, J., and performance supercapacitors. Electrochim. (2019). Facile method to synthesis hybrid
Byon, C. (2018). Tiny MoO3 nanocrystals self- Acta 270, 387–394. phase 1T@2H MoSe2 nanostructures for
assembled on folded molybdenum disulfide rechargeable lithium ion batteries.
nanosheets via a hydrothermal method for 158. Wang, D., Xiao, Y., Luo, X., Wu, Z., Wang, J. Electroanal. Chem. 833, 333–339.
supercapacitor. Mater. Res. Lett. 6, 432–441. Y.-J., and Fang, B. (2017). Swollen
ammoniated MoS2 with 1T/2H hybrid phases 171. Enyashin, A.N., Yadgarov, L., Houben, L.,
147. Mu, J., Guan, Y., Wang, L., Li, H., Liu, Y., Che, for high-rate electrochemical energy storage. Popov, I., Weidenbach, M., Tenne, R., Bar-
H., Liu, A., Guo, Z., Zhang, X., and Zhang, Z. ACS Sustain. Chem. Eng. 5, 2509–2515. Sadan, M., and Seifert, G. (2011). New route
(2019). Flexible heat-treated PAN nanofiber/ for stabilization of 1T-WS2 and MoS2 phases.
MoO2/MoS2 composites as high performance 159. Acerce, M., Voiry, D., and Chhowalla, M. J. Phys. Chem. C 115, 24586–24591.
supercapacitor electrodes. J. Mater. Sci. (2015). Metallic 1T phase MoS2 nanosheets as
Mater. Electron. 30, 8210–8219. supercapacitor electrode materials. Nat. 172. Savjani, N., Lewis, E.A., Bissett, M.A., Brent,
Nanotechnol. 10, 313–318. J.R., Dryfe, R.A.W., Haigh, S.J., and O’Brien, P.
148. Song, X.-Z., Sun, F.-F., Meng, Y.-L., Wang, (2016). Synthesis of lateral size-controlled
Z.-W., Su, Q.-F., and Tan, Z. (2019). Hollow 160. Kan, M., Wang, J.Y., Li, X.W., Zhang, S.H., Li, monolayer 1H-MoS2@Oleylamine as
core–shell NiCo2S4@MoS2 dodecahedrons Y.W., Kawazoe, Y., Sun, Q., and Jena, P. supercapacitor electrodes. Chem. Mater. 28,
with enhanced performance for (2014). Structures and phase transition of a 657–664.

552 Matter 2, 526–553, March 4, 2020


173. Parveen, N., Ansari, S.A., Cho, M.H., and 174. Gopalsamy, K., Balamurugan, J., Thanh, T.D., 175. Nayak, A.P., Moran, S.T., Akinwande, D., Jin,
Ansari, M.O. (2015). Simultaneous sulfur Kim, N.H., and Lee, J.H. (2017). Fabrication of C., Yuan, Z., Cao, B., Li, T., Liu, J., Wu, J., and
doping and exfoliation of graphene from nitrogen and sulfur co-doped graphene Lin, J.F. (2015). Pressure-modulated
graphite using an electrochemical method for nanoribbons with porous architecture for conductivity, carrier density, and mobility of
supercapacitor electrode materials. J. Mater. high-performance supercapacitors. Chem. multilayered tungsten disulfide. ACS Nano 9,
Chem. A 4, 233–240. Eng. J. 312, 180–190. 9117–9123.

Matter 2, 526–553, March 4, 2020 553

You might also like