Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Letter

pubs.acs.org/JPCL

Electric Field and Strain Effect on Graphene-MoS2 Hybrid Structure:


Ab Initio Calculations
Xingen Liu and Zhongyao Li*
College of Science, University of Shanghai for Science and Technology, Shanghai 200093, P. R. China

ABSTRACT: The electronic structures of graphene-MoS2 heterojunction under tension


and external electric field were examined on the basis of density-functional theory. The
tension of MoS2 changes the hybrid structure from semiconductor to metal. The transition
is from the sensitive dependence of the bandgap of MoS2 on the lattice constant. The
vertical electric field has little influence on the bandgap of MoS2, while it can also adjust
the charge transfer between monolayer MoS2 and graphene. In addition, the Schottky
barrier is linearly dependent on the electric field intensity with an effective vacuum spacing
of 1.3 Å. It is also discussed in detail that the bandgap of MoS2 dependence on the lattice
constant and the S−S spacing.

S ince the successful invention of isolated graphene in


experiments,1 a new materials fieldtwo-dimensional
(2D) materialhas received much attention. It is a popular
also result in a tunable photoresponsivity.32 Theoretically, the
electron mobility of the heterojunction is comparable to
graphene.33 Moreover, the bandgap can be changed by
material for manufacturing the miniaturization and integration adjusting the interfacial distance.33 It provides a new train of
of electronic devices, such as layered electrodes2,3 and thin film thought that the graphene-MoS2 heterojunction can be used to
electronic devices.4 Although the most famous 2D material, design devices with small bandgap and high carrier mobility. In
graphene, has many excellent properties and has been addition, it was reported that the bandgap of the graphene
integrated into many different applications, its gapless-semi- bilayer can be opened at weak electric gating,8 and the work
conductor band structure may be an obvious disadvantage in function and charge transfer of doped graphene can be well
some cases.5 Band gap is necessary in many application fields, adjusted by applying an external electric field.34 External electric
such as light-emitting diode (LED),6 solar battery7 and field may be also used to adjust the electronic properties of the
transistor technology.8 This disadvantage restricts the applica- 2D layered heterojunction. The electronic properties of
tion of graphene in many areas. The molybdenum disulfide9 is materials are directly related to various physical and chemical
another kind of 2D material. The atomic structure of properties and practical applications. In order to further study
monolayer MoS2 is two S-layers sandwiching a Mo-layer, and the potential applications of graphene−MoS2, it is necessary to
the atoms in layers are hexagonally packed.10 It has good know more about the possible electronic properties of the
chemical and thermal stability, large specific surface area, and heterojunction.
high surface activity.11−14 With unique physical and chemical In this work, we study the band structures of two kinds of
properties, it has potential application in catalysis, lubrication, graphene-MoS2 hybrid structures: (5×5)graphene−(4×4)MoS2
and electrochemical lithium storage.15−17 Single-layer MoS2 is a and (4×4)graphene−(3×3)MoS2. Graphene is stretched in the
direct bandgap semiconductor, which has very strong luminous
first model, while the monolayer MoS2 is stretched in the
intensity.18,19 Moreover, it can be used as a channel material to
second model. The stretching of MoS2 would change the
manufacture an ultralow standby power field effect transistor
hybrid structure from semiconductor to metal due to the charge
with high current switch ratio and high electron mobility.17
The layers of different 2D semiconductors can be stacked to transfer from graphene to MoS2. Morover, the external electric
form semiconductor heterojunction, and the novel physical field can also adjust the charge transfer between graphene and
phenomena in such heterojunctions have also become a focus MoS2. Although the bandgap of MoS2 is insensitive to the
of international nanoscience. Currently, the hot research of graphene-MoS2 interaction33 and the vertical electric field, it is
nanodevices is based on the heterojunction formed by graphene greatly dependent on the lattice constant and the S−S spacing.
and MoS2, which is a new type of 2D layered heterojunction; The electronic structures were calculated on the basis of
the device has a light weight, low power consumption, and density-functional theory (DFT) at the level of local density
flexibility.20−26 Experimentally,27 the growth of MoS2 on
graphene will increase the electron transfer rate and improve Received: June 10, 2015
the electrochemical performance. It has shown very good Accepted: August 4, 2015
performance for lithium ion batteries and aerogels.28−31 It can Published: August 4, 2015

© 2015 American Chemical Society 3269 DOI: 10.1021/acs.jpclett.5b01233


J. Phys. Chem. Lett. 2015, 6, 3269−3275
The Journal of Physical Chemistry Letters Letter

approximation (LDA).35 Although there may be some the interfacial spacing is about 3.29 Å. The external electric filed
limitations within the LDA method,36−38 most of the early is along the z-direction and perpendicular to the graphene
DFT results based on LDA about the interface formed between plane. The electric filed intensity is from −0.5 V/Å to +0.5 V/
graphene and substrate have been confirmed by experi- Å. The Γ-centered k-point grids of 9 × 9 × 1 and 21 × 21 × 1
ments.33,39−44 LDA is usually used in the calculations for were employed to sample the Brillouin zone for the supercell of
graphene, MoS 2 , and/or graphene−MoS 2 heterostruc- graphene-MoS2 heterojunction and for the unit cell of
ture.25,33,45−47 Moreover, it was reported that the DFT with monolayer MoS2 (Figure 1e,f), respectively.
van der Waals (vdW) correction is not more accurate than The band structure of (5×5)graphene−(4×4)MoS2 hetero-
conventional DFT methods for the graphene-based interface junction is shown in Figure 2a. It was reported that the
modeling study.25,33,48−55 Therefore, the conventional LDA is
used in all our calculations.
The projector-augmented wave (PAW) pseudopotentials56
method implemented in the VASP package57,58 were employed
to describe the effect of core electrons. The energy cutoff in
calculations was set to be 400 eV, and the total energy was
converged to better than 10−5eV. The equilibrium structures
were obtained through structural relaxation until Hellmann−
Feynman forces were less than 0.02 eV/Å. The atomic structure
was modeled by a periodic slab geometry, with a vacuum of at
least 20 Å between two neighboring slabs. The slab structures
are shown in Figure 1. There are two models considered in this

Figure 2. (a−c) The band structure of (5×5)graphene−(4×4)MoS2


heterojunction (M_1) with the external electric field of 0 V/Å (a),
−0.3 V/Å (b), and 0.3 V/Å (c), respectively. The Schottky barrier is
denoted as φB in panel a. (d−f) The band structure of (4×4)-
graphene−(3×3)MoS2 heterojunction (M_2) with the external
electric field of 0 V/Å (d), −0.3 V/Å (e), and 0.3 V/Å (f),
respectively. The Fermi level is set to zero. The Dirac point of
graphene (K), the valence band maximum (VBM), and the conduction
Figure 1. Structures of graphene−MoS2 heterojunction. (a) Side view band minimum (CBM) of MoS2 are marked with a black star, blue
of the (5×5)graphene−(4×4)MoS2 heterojunction (Model 1, square, and circle, respectively. The band gap of MoS2 is denoted as
abbreviated as M_1). The red rectangle is the unit cell of MoS2. (b) ΔE.
Side view of the (4×4)graphene−(3×3)MoS2 heterojunction (Model
2, abbreviated as M_2). (c) Top view of the (5×5)graphene−
(4×4)MoS2 heterojunction (M_1). (d) Top view of the (4×4)-
interlayer orientation would affect the structure and electronic
graphene−(3×3)MoS2 heterojunction (M_2). (e) Top view of properties of graphene−MoS2 heterostructure.61 However, for
(1×1)MoS2 as the red rhombus in (c). The lattice constant is the graphene−MoS2 heterostructure studied in this work, the
denoted as a. (f) Side view of (1×1)MoS2. The S−S layer spacing is [21̅1̅0] crystallographic directions of the two layers are parallel
denoted as ds. to one another.61 In this case, the weak graphene−MoS2
interaction has little influence on the band structure of the
heterostructure. It is also consistent with the reported results.33
work: (5×5)graphene−(4×4)MoS2 and (4×4)graphene− Therefore, the band structure of the heterostructure is simply a
(3×3)MoS2 hybrid structures. In the first model (Figure combination of the energy bands of individual graphene and
1a,c), the length of the superlattice vectors was set to be 12.772 MoS2. The weak graphene−MoS2 interaction is not enough to
Å. In this case, the lattice constant of MoS2 unit cell was set to change the morphology of their energy bands. The linear
be about 3.193 Å,59 and the stretching of the lattice constant of dispersion bands of graphene are in the large bandgap of MoS2.
graphene is about 3.8%. At equilibrium, the interfacial spacing, Moreover, the Dirac point of graphene is at the Fermi level (EF
between graphene and MoS2, is about 3.34 Å. In the second = 0), suggesting no charge transfer between graphene and
model (Figure 1b,d), the length of the superlattice vectors was MoS2. The Schottky barrier,22,62 which is the excitation energy
set to be 9.84 Å. In this case, the lattice constant of the of electrons from graphene to MoS2, is also marked as φB in
graphene unit cell is set to be about 2.46 Å,60 and the stretching Figure 2a. In the heterojunction with 3.8% strain of graphene,
of the lattice constant of MoS2 is about 2.7%. At equilibrium, the Schottky barrier φB is about 0.418 eV. At the Dirac point,
3270 DOI: 10.1021/acs.jpclett.5b01233
J. Phys. Chem. Lett. 2015, 6, 3269−3275
The Journal of Physical Chemistry Letters Letter

there is a tiny splitting, which is about 5 meV, due to the weak less than −0.4 V/Å, the Dirac point of graphene would be
graphene−MoS2 interaction.33 The band structure of the above the Fermi level for the (5×5)graphene−(4×4)MoS2
graphene−MoS2 heterojunction can be changed by the heterojunction; meanwhile, the CBM of MoS2 would be
stretching of MoS2. As a comparison, the band structure of below the Fermi level. Therefore, the heterojunction changes
(4×4)graphene−(3×3)MoS2 heterojunction is shown in Figure from semiconductor to metal under the electric field. Similarly,
2d. The bottom of the conduction band of MoS2 is below the the (4×4)graphene−(3×3)MoS2 heterojunction can also be
Fermi level, while the Dirac point of graphene is above the changed from semiconductor to metal when the electric filed
Fermi level. This suggests that electrons transfer from graphene intensity is less than 0.2 V/Å. In order to realize the
to MoS2. Therefore, semiconductor−metal transition can be semiconductor-metal transition, both the magnitude and the
realized by the stretching of MoS2 in the heterojunction. direction of electric field are different for the two hetero-
Moreover, the Schottky barrier and the charge transfer can be junctions. The difference is from the difference of Schottky
controlled by external electric field as shown in Figure 2. barrier induced by stretching.
Figure 3a shows the sketch map of a graphene−MoS2 As shown in Figure 3b,c, the Schottky barrier22 in the
heterojunction in an external electric field. The positive electric semiconductor heterojunctions can be fitted as
φB = φ0 + eEd (1)
where E is the electric field intensity, and d is the effective
vacuum spacing between graphene and MoS2. In the (5×5)-
graphene−(4×4)MoS2 heterojunction, φ0 = 0.418 eV and d =
1.26 Å; while φ0 = −0.311 eV and d = 1.29 Å in the
(4×4)graphene−(3×3)MoS2 heterojunction. Considering the
interfacial spacing between graphene and MoS2 D ∼ 3.3 Å, the
S atomic radius rs ∼ 1.04 Å and the C atomic radius rc ∼ 0.77 Å,
the effective vacuum spacing should be close to (D − rs − rc)
≃1.5 Å. Our fitting result, d∼ 1.3 Å, is in good agreement with
the estimated value (1.5 Å). Although the effective vacuum
spacing is insensitive to the stretching of graphene or MoS2, the
constant φ0 in eq 1 can be greatly changed by stretching. It
directly leads to the dependence of the Schottky barrier on
stretching.
The work function,34 which can be measured in experiment,
is also discussed. It is linear with the applied electric field.34
Note that positive electric field is from graphene to MoS2 in our
models. Both the work functions of graphene surface and MoS2
surface can be fitted as ϕW = ϕ0 + γeE, where ϕ0 is the work
function for the heterostructure without electric field, and γ is a
coefficient that is negative for graphene surface and positive for
MoS2 surface. In the (5×5)graphene−(4×4)MoS2 hetero-
Figure 3. (a) Sketch map of graphene-MoS2 heterojunction in external junction, ϕ0 is about 4.891 eV; the coefficient γ is about
electric field. The interfacial spacing and the effective vacuum spacing,
−3.943 and 5.375 Å for graphene and the MoS2 surface,
between MoS2 and graphene, are denoted as D (∼3.3 Å) and d (∼1.3
Å), respectively. (b) The energy of the Dirac point of graphene K, the respectively. As for the (4×4)graphene−(3×3)MoS2 hetero-
VBM, and the CBM of MoS2 in (5×5)graphene−(4×4)MoS2 junction, ϕ0 = 4.816 eV; γ = −4.805 Å for the graphene surface
heterojunction (M_1) under electric field. (c) The energy of K, and γ = 4.058 Å for the MoS2 surface.
VBM and CBM in (4×4)graphene−(3×3)MoS2 heterojunction The bandgap of MoS2, which is the energy difference
(M_2) under electric field. The Fermi level is set to be zero. The between the CBM and VBM of MoS2, is marked as ΔE in
Schottky barrier is denoted as φB. It can be fitted by φB = φ0 + eEd, Figure 3b,c. Although the electric field can be used to adjust the
where φ0 = 0.418 eV and d = 1.26 Å in panel b, while φ0 = −0.311 eV Schottky barrier, the work function and the charge transfer
and d = 1.29 Å in panel c. between graphene and MoS2, it has little influence on the
bandgap of MoS 2 . In the (5×5)graphene−(4×4)MoS 2
field intensity, which is along the z-direction, would enlarge the heterojunction (Figure 3b), the bandgap ΔE ≈ 1.42 eV;
Schottky barrier and prevent the charge transfer from graphene while it is about 0.8 eV in the (4×4)graphene−(3×3)MoS2
to MoS2. The interfacial spacing and the effective vacuum heterojunction (Figure 3c) due to the stretching of MoS2. Since
spacing, between graphene and MoS2, are marked as D (∼3.3 molybdenum disulfide has high mechanical flexibility,63
Å)33 and d (∼1.3 Å) in Figure 3a, respectively. The interfacial stretching may be an effective measure to adjust the band
spacing can be obtained by the structural relaxation, while the structure of MoS2 and the electronic properties of graphene−
effective vacuum spacing is from the linear fitting of the MoS2 heterojunction in practice.
Schottky barrier dependence on electric field, which will be In our graphene-MoS2 models, the bandgap of MoS2 is
discussed in the following. Under the electric field, the energy almost unchanged by graphene. The analysis of the bandgap of
of the Dirac point of graphene (K), the valence band maximum MoS2 was carried out in MoS2 monolayer. In order to further
(VBM), and the conduction band minimum (CBM) of MoS2 analyze the influence of strain on the bandgap of monolayer
are shown in Figure 3b for the (5×5)graphene−(4×4)MoS2 MoS2, we constructed the MoS2 unit cell as shown in Figures
heterojunction and in Figure 3c for the (4×4)graphene− 1(e) and 1(f). The lattice constant is denoted as a, and the S−S
(3×3)MoS2 heterojunction. When the electric field intensity is layer spacing is denoted as ds. Since the nature of the energy
3271 DOI: 10.1021/acs.jpclett.5b01233
J. Phys. Chem. Lett. 2015, 6, 3269−3275
The Journal of Physical Chemistry Letters Letter

bandgap would be changed from direct to indirect for an Therefore, the equilibrium S−S spacing should have the
applied tensile strain,64,65 both the direct and indirect bandgaps following dependence on the lattice constant:
are considered in the following. With the lattice constant of ⎡
3.193 Å, the equilibrium S−S spacing deq ds0 − dseq a − a0 ⎛ a − a 0 ⎞2 ⎤
s = 3.064 Å, the direct
bandgap at K point66,67 ΔEK = 1.594 eV and the ΓK indirect = c0 + c ⎢λ1′ − λ 2′⎜ ⎟⎥
ds0 ⎢⎣ a0 ⎝ a0 ⎠ ⎥⎦
bandgap65,67 ΔE = 1.426 eV. (4)
Figure 4a shows the indirect bandgaps with different lattice where c0 and c are constants. The fitted values show that c0 =
constants and S−S layer spacing. When the lattice constant a is 2.96 × 10−5 and c = 0.064. Since c0 ≪ 1, the formula can be
simplified as
⎡ ⎛ a − a 0 ⎞2 ⎤
ds0 − dseq a − a0
= c ⎢λ1′ − λ 2′⎜ ⎟⎥
ds0 ⎢⎣ a0 ⎝ a0 ⎠ ⎥⎦ (5)
Take eq 5 into eq 3, the ΓK indirect bandgap at equilibrium,
⎡ ⎛ a − a 0 ⎞2 ⎤
a − a0
ΔE eq = ⎢1 − 17.7 × + 55.73 × ⎜ ⎟ ⎥ΔE0
⎢⎣ a0 ⎝ a0 ⎠ ⎥⎦
(6)
It can also be equally expressed in the equilibrium S−S spacing:
⎡ d − dseq ⎤
ΔE eq = ⎢1 − 23.11 × s0 ⎥ΔE0
⎣ ds0 ⎦ (7)
Similarly, the linear relationship in Figure 4b suggests the direct
bandgap at the K point can be fitted by the fitting formula:
ΔE K0 − ΔE K a − a0 ⎛ a − a 0 ⎞2 d − ds
= λ 0 + λ1 − λ 2⎜ ⎟ + η s0
ΔE K0 a0 ⎝ a0 ⎠ ds0
(8)
Figure 4. (a) The ΓK indirect bandgap of MoS2 (ΔE). (b) The direct where ΔEK0 = 1.594 eV. The fitted values show that λ0 =
bandgap of MoS2 at the K point (ΔEK). The bandgaps at equilibrium
−0.003, λ1 = 5.68, λ2 = 17.88, and η = 1.62. Since |λ0| ≪ 1, the
S−S spacing (deqs ) are marked in green hexagon. (c) The p-component
of the VBM and the CBM of MoS2. (d) The d-component of the VBM direct bandgap at the K point ΔEK can be simplified as
and the CBM of MoS2. In the unit cell of MoS2 (Figure 1e,f), the ⎡ ⎛ a − a 0 ⎞2 ⎤
a − a0 d − ds ⎥
lattice constant is denoted as a, the constant a0 = 3.193 Å, and the S−S ΔE K = ⎢1 − λ1 + λ 2⎜ ⎟ − η s0 ΔE K0
spacing is denoted as ds. ⎢⎣ a0 ⎝ a0 ⎠ ds0 ⎥⎦
(9)
At equilibrium, considering the S−S spacing dependence on the
fixed, the indirect bandgap ΔE is linearly dependent on the S−S lattice constant, eq 5, it can be further simplified as
spacing ds. Based on this phenomenon, the indirect bandgap
can be fitted by the fitting formula: ⎡ ⎛ a − a 0 ⎞2 ⎤
⎢ a − a0
ΔE Keq = 1 − 6.92 × + 21.79 × ⎜ ⎟ ⎥Δ
ΔE0 − ΔE a − a0 ⎛ a − a 0 ⎞2 d − ds ⎢⎣ a0 ⎝ a0 ⎠ ⎥⎦
= λ 0′ + λ1′ − λ 2′⎜ ⎟ + η′ s0
ΔE0 a0 ⎝ a0 ⎠ ds0
(2)
E K0 (10)
where a0 = 3.193 Å, ds0 = 3.064 Å, and ΔE0 = 1.426 eV. The Comparing eq 10 with eq 6, the direct and indirect bandgaps at
fitted values show that λ′0 = −0.006, λ′1 = 11.97, λ′2 = 37.69, and equilibrium have a simply linear relationship:
η′ = 7.48. Since the constant |λ′0| ≪ 1, it can be neglected. ΔE Keq = 0.44 × ΔE eq + 0.97 eV (11)
Therefore, the ΓK indirect bandgap ΔE can be simplified as
⎡ ⎤ The above equations are in excellent agreement with our first-
a − a0 ⎛ a − a 0 ⎞2 d − ds ⎥ principles calculations. The lower bound of the bandgap for

ΔE = 1 − λ1′ + λ 2′⎜ ⎟ − η′ s0 ΔE0
⎢⎣ a0 ⎝ a0 ⎠ ds0 ⎥⎦ which the single-layer MoS2 remains as a direct bandgap
semiconductor is about 1.72 eV, which is about 1.68 eV in ref
(3) 65. It was also reported that single-layer MoS2 turns into metal
Although the quadratic component [(a − a0)/a0]2 ≪ (a − a0)/ at a biaxial tensile strain of 9% in refs 64 and 65. Note that the
a0 for a small tensile strain (<10%), its coefficient λ′2 = 37.69 ≫ lattice constant of unstrained MoS2 is set to be 3.16 Å in refs 64
1, its influence on the bandgap should not be neglected. The and 65, while it is 3.193 Å (a0) in this work. The indirect
fitting formula also suggests that the indirect bandgap is bandgap from eq 6 would be negative, suggesting single-layer
sensitively dependent on the lattice constant and the S−S layer MoS2 turns into metal when 7.4% < (a − a0)/a0 < 24.4%
spacing. (3.428 Å < a < 3.972 Å). Moreover, the direct bandgap from eq
At equilibrium, the indirect bandgap is also linearly 10 is positive, and it would be larger than the indirect bandgap
dependent on the S−S spacing as shown in Figure 4a. when −1.1% < (a − a0)/a0 < 32.9% (3.157 Å < a < 4.244 Å).
3272 DOI: 10.1021/acs.jpclett.5b01233
J. Phys. Chem. Lett. 2015, 6, 3269−3275
The Journal of Physical Chemistry Letters Letter

In experiment,68 the exciton absorption peak energy of bandgap at equilibrium. According to our fitting formulas, the
monolayer MoS2, which is associated with the direct bandgap bandgaps of MoS2 with strains can be obtained. Stretching may
transitions around the K point, linearly decreases with the be an effective measure to adjust the band structure of MoS2
increase of strain for the relatively small strains (<1%). This and the electronic properties of the graphene−MoS 2
may suggest the linear dependence of the direct bandgap on heterojunction in practice.


strain when the strain is relatively small (<1%). In fact, the
quadratic components in eq 10 and eq 6 can be neglected for AUTHOR INFORMATION
the relatively small strains (<1%). Therefore, the linear
Corresponding Author
dependence on strain can be obtained. According to our
equations, the bandgaps of MoS2 with different lattice constants *E-mail: lizyusst@gmail.com.
can be obtained. Some results deduced from the equations are Notes
listed in Table 1. They are consistent with the results reported The authors declare no competing financial interest.

Table 1. Bandgaps of MoS2 at Equilibrium with Different


Lattice Constantsa
■ ACKNOWLEDGMENTS
This work was supported by the National Natural Science
a deq
s d*s ΔEeq ΔEeq
K ΔE* ref Foundation of China under grant No. 11204178, the Hujiang
3.122 3.120 3.116 2.03 1.86 2.03/1.79 69 Foundation of China (B14004), and the Shanghai Young
3.130 3.110 3.120 1.96 1.83 1.86 70 College Teachers Training Program under Grant No. slg11030.
3.150 3.097 −− 1.78 1.75 1.74 71 The computational resources utilized in this research were
3.160 3.089 −− 1.70 1.71 1.70/1.78 66, 72 provided by the Shanghai Supercomputer Center. The authors
3.193 3.064 3.134 1.43 1.59 1.66 59, 73 also acknowledge the support from the University of Shanghai
a for Science and Technology (National Projects and Arts Base
The lattice constant a and the equilibrium S−S spacing deq s are in the
Training Program) under Grant No. 12XGQ01.


unit of Å. The equilibrium S−S spacing is deduced from eq 5. The S−
S spacing reported in the references is denoted as ds* in units of Å. The
equilibrium indirect bandgap ΔEeq and the equilibrium direct bandgap REFERENCES
ΔEeq
K are deduced from eq 6 and eq 10, respectively. The bandgaps (1) Novoselov, K.; Geim, A.; Morozov, S.; Jiang, D.; Zhang, Y.;
reported in the references are denoted as ΔE*. All the bandgaps are in Dubonos, S.; Grigorieva, I.; Firsov, A. Electric Field Effect in
units of eV. Atomically Thin Carbon Films. Science 2004, 306, 666−669.
(2) Zhang, L.; Zhao, S.; Tian, X.; Zhao, X. Layered Graphene Oxide
Nanostructures with Sandwiched Conducting Polymers as Super-
in refs 66 and 69−72, except for the results with a = 3.193 Å in capacitor Electrodes. Langmuir 2010, 26, 17624−17628.
refs 59 and 73 due to the relatively large difference in the (3) Gao, Z.; Yang, W.; Yan, Y.; Wang, J.; Ma, J.; Zhang, X.; Xing, B.;
equilibrium S−S spacing. Liu, L. Synthesis and Exfoliation of Layered α-Co(OH)2 Nanosheets
Although the bandgap of MoS2 is sensitively dependent on and Their Electrochemical Performance for Supercapacitors. Eur. J.
the lattice constant and the S−S spacing, the state components Inorg. Chem. 2013, 2013, 4832−4838.
of VBM and CBM are hardly changed by strain. As shown in (4) Lei, S.; Ge, L.; Najmaei, S.; George, A.; Kappera, R.; Lou, J.;
Figure 4c,d, the d-states of Mo are dominant in VBM; while the Chhowalla, M.; Yamaguchi, H.; Gupta, G.; Vajtai, R.; et al. Evolution
of the Electronic Band Structure and Efficient Photo-Detection in
p-component is comparable with the d-component in CBM. Atomic Layers of InSe. ACS Nano 2014, 8, 1263−1272.
With the increase of lattice constant, the d-component (p- (5) Kou, L.; Tang, C.; Zhang, Y.; Heine, T.; Chen, C.; Frauenheim,
component) would be slightly enlarged (decreased) in both T. Tuning Magnetism and Electronic Phase Transitions by Strain and
VBM and CBM. It may be related to the change in bandgap. Electric Field in Zigzag MoS2 Nanoribbons. J. Phys. Chem. Lett. 2012,
However, the slight change in state components makes it seem 3, 2934−2941.
impossible to well understand the large change in the bandgap (6) Liu, G.; Zhang, J.; Tan, C.; Tansu, N. Efficiency-Droop
of MoS2 under tension. Suppression by Using Large-Bandgap AlGaInN Thin Barrier Layers
In summary, we investigated the electronic structures of in InGaN Quantum-Well Light-Emitting Diodes. IEEE Photonics J.
graphene−MoS2 heterojunction under tension and external 2013, 5, 2201011.
electric field. The (5×5)graphene−(4×4)MoS2 heterojunction (7) Mariotti, D.; Mitra, S.; Svrcek, V. Surface-Engineered Silicon
Nanocrystals. Nanoscale 2013, 5, 1385−1398.
is a semiconductor with 3.8% strain of graphene, while (8) Zhang, Y.; Tang, T.; Girit, C.; Hao, Z.; Martin, M. C.; Zettl, A.;
electrons transfer from graphene to MoS2 in the (4×4)- Crommie, M. F.; Shen, Y.; Wang, F. Direct Observation of a Widely
graphene−(3×3)MoS2 heterojunction with 2.7% strain of Tunable Bandgap in Bilayer Graphene. Nature 2009, 459, 820−823.
MoS2. Semiconductor−metal transition can be realized by the (9) Ghatak, S.; Pal, A.; Ghosh, A. Nature of Electronic States in
stretching of MoS2 in the graphene-MoS2 heterojunction. The Atomically Thin MoS2 Field-Effect Transistors. ACS Nano 2011, 5,
charge transfer, the Schottky barrier, and the work function can 7707−7712.
also be controlled by external electric field. From our first- (10) Pan, H.; Zhang, Y. Tuning the Electronic and Magnetic
principles calculations, the Schottky barrier is linearly depend- Properties of MoS2 Nanoribbons by Strain Engineering. J. Phys. Chem.
ent on the electric field intensity with an effective vacuum C 2012, 116, 11752−11757.
spacing of 1.3 Å, and the work function is also linear with the (11) Wilson, J. A.; Yoffe, A. D. The Transition Metal
Dichalcogenides. Discussion and Interpretation of the Observed
applied electric field. Although the electric field can be used to Optical, Electrical and Structural Properties. Adv. Phys. 1969, 18,
adjust the Schottky barrier, the work function, and the charge 193−335.
transfer, it has little influence on the bandgap of MoS2. Both the (12) Mattheiss, L. F. Energy Bands for 2H-NbSe2 and 2H-MoS2.
direct and indirect bandgaps of MoS2 are sensitively dependent Phys. Rev. Lett. 1973, 30, 784−787.
on the lattice constant and the S−S spacing. Interestingly, the (13) Coehoorn, R.; Haas, C.; Dijkstra, J.; Flipse, C. J. F.; de Groot, R.
direct bandgap has a linear relationship with the indirect A.; Wold, A. Electronic Structure of MoSe2, MoS2, and WSe2. I. Band-

3273 DOI: 10.1021/acs.jpclett.5b01233


J. Phys. Chem. Lett. 2015, 6, 3269−3275
The Journal of Physical Chemistry Letters Letter

Structure Calculations and Photoelectron Spectroscopy. Phys. Rev. B: (33) Ma, Y.; Dai, Y.; Guo, M.; Niu, C.; Huang, B. Graphene
Condens. Matter Mater. Phys. 1987, 35, 6195−6202. Adhesion on MoS2 Monolayer: An ab Initio Study. Nanoscale 2011, 3,
(14) Li, T.; Galli, G. Electronic Properties of MoS2 Nanoparticles. J. 3883−3887.
Phys. Chem. C 2007, 111, 16192−16196. (34) Gholizadeh, R.; Yu, Y. Work Functions of Pristine and
(15) Mdleleni, M.; Hyeon, T.; Suslick, K. Sonochemical Synthesis of Heteroatom-Doped Graphenes under Different External Electric
Nanostructured Molybdenum Sulfide. J. Am. Chem. Soc. 1998, 120, Fields: An ab Initio DFT Study. J. Phys. Chem. C 2014, 118,
6189−6190. 28274−28282.
(16) Rapoport, L.; Bilik, Y.; Feldman, Y.; Homyonfer, M.; Cohen, S. (35) Ceperley, D. M.; Alder, B. J. Ground State of the Electron Gas
R.; Tenne, R. Hollow Nanoparticles of WS2 as Potential Solid-State by a Stochastic Method. Phys. Rev. Lett. 1980, 45, 566−569.
Lubricants. Nature 1997, 387, 791−793. (36) Ellis, J. K.; Lucero, M. J.; Scuseria, G. E. The Indirect to Direct
(17) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Band Gap Transition in Multilayered MoS2 as Predicted by Screened
Single-Layer MoS2 Transistors. Nat. Nanotechnol. 2011, 6, 147−150. Hybrid Density Functional Theory. Appl. Phys. Lett. 2011, 99, 261908.
(18) Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically (37) Feng, J.; Qian, X.; Huang, C.; Li, J. Strain-Engineered Artificial
Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, Atom as a Broad-Spectrum Solar Energy Funnel. Nat. Photonics 2012,
105, 136805. 6, 866−872.
(19) Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.; (38) Shi, H.; Pan, H.; Zhang, Y.; Yakobson, B. I. Quasiparticle Band
Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer MoS2. Structures and Optical Properties of Strained Monolayer MoS2 and
Nano Lett. 2010, 10, 1271−1275. WS2. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87, 155304.
(20) Bertolazzi, S.; Krasnozhon, D.; Kis, A. Nonvolatile Memory (39) Lee, B.; Han, S.; Kim, Y. First-Principles Study of Preferential
Cells Based on MoS2/Graphene Heterostructures. ACS Nano 2013, 7, Sites of Hydrogen Incorporated in Epitaxial Graphene on 6H-
3246−3252. SiC(0001). Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81,
(21) Choi, M. S.; Lee, G.; Yu, Y.; Lee, D.; Lee, S. H.; Kim, P.; Hone, 075432.
J.; Yoo, W. J. Controlled Charge Trapping by Molybdenum Disulphide (40) Kang, M. H.; Jung, S. C.; Park, J. W. Density Functional Study
and Graphene in Ultrathin Heterostructured Memory Devices. Nat. of the Au-Intercalated Graphene/Ni(111) Surface. Phys. Rev. B:
Commun. 2013, 4, 1624. Condens. Matter Mater. Phys. 2010, 82, 085409.
(22) Yu, L.; Lee, Y.; Ling, X.; Santos, E. J. G.; Shin, Y.; Lin, Y.; (41) Kamiya, K.; Umezawa, N.; Okada, S. Energetics and Electronic
Dubey, M.; Kaxiras, E.; Kong, J.; Wang, H.; et al. Graphene/MoS2 Structure of Graphene Adsorbed on HfO2(111): Density Functional
Hybrid Technology for Large-Scale Two-Dimensional Electronics. Theory Calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2011,
Nano Lett. 2014, 14, 3055−3063. 83, 153413.
(23) Roy, K.; Padmanabhan, M.; Goswami, S.; Sai, T.; Ramalingam, (42) Giovannetti, G.; Khomyakov, P. A.; Brocks, G.; Kelly, P. J.; van
G.; Raghavan, S.; Ghosh, A. Graphene-MoS2 Hybrid Structures for den Brink, J. Substrate-Induced Band Gap in Graphene on Hexagonal
Multifunctional Photoresponsive Memory Devices. Nat. Nanotechnol. Boron Nitride: Ab Initio Density Functional Calculations. Phys. Rev. B:
Condens. Matter Mater. Phys. 2007, 76, 073103.
2013, 8, 826−830.
(43) Nguyen, T. C.; Otani, M.; Okada, S. Semiconducting Electronic
(24) Xu, H.; Wu, J.; Feng, Q.; Mao, N.; Wang, C.; Zhang, J. High
Property of Graphene Adsorbed on (0001) Surfaces of SiO2. Phys. Rev.
Responsivity and Gate Tunable Graphene-MoS2 Hybrid Photo-
Lett. 2011, 106, 106801.
transistor. Small 2014, 10, 2300−2306.
(44) Okada, S. Semiconducting Electronic Structure of Graphene
(25) Sachs, B.; Britnell, L.; Wehling, T. O.; Eckmann, A.; Jalil, R.;
Adsorbed on Insulating Substrate: Fragility of the Graphene Linear
Belle, B. D.; Lichtenstein, A. I.; Katsnelson, M. I.; Novoselov, K. S.
Dispersion Band. Jpn. J. Appl. Phys. 2010, 49, 020204.
Doping Mechanisms in Graphene-MoS2 Hybrids. Appl. Phys. Lett.
(45) Li, X.; Yu, S.; Wu, S.; Wen, Y.; Zhou, S.; Zhu, Z. Structural and
2013, 103, 251607. Electronic Properties of Superlattice Composed of Graphene and
(26) Larentis, S.; Tolsma, J. R.; Fallahazad, B.; Dillen, D. C.; Kim, K.;
Monolayer MoS2. J. Phys. Chem. C 2013, 117, 15347−15353.
MacDonald, A. H.; Tutuc, E. Band Offset and Negative Compressi- (46) Sun, L.; Yan, J.; Zhan, D.; Liu, L.; Hu, H. L.; Li, H.; Tay, B. K.;
bility in Graphene-MoS2 Heterostructures. Nano Lett. 2014, 14, 2039− Kuo, J. L.; Huang, C. C.; Hewak, D. W.; et al. Spin-Orbit Splitting in
2045. Single-Layer MoS2 Revealed by Triply Resonant Raman Scattering.
(27) Chang, K.; Chen, W. In Situ Synthesis of MoS2/Graphene Phys. Rev. Lett. 2013, 111, 126801.
Nanosheet Composites with Extraordinarily High Electrochemical (47) Cao, C.; Wu, M.; Jiang, J.; Cheng, H. Transition Metal Adatom
Performance for Lithium Ion Batteries. Chem. Commun. 2011, 47, and Dimer Adsorbed on Graphene: Induced Magnetization and
4252−4254. Electronic Structures. Phys. Rev. B: Condens. Matter Mater. Phys. 2010,
(28) Jiang, H.; Ren, D.; Wang, H.; Hu, Y.; Guo, S.; Yuan, H.; Hu, P.; 81, 205424.
Zhang, L.; Li, C. 2D Monolayer MoS2-Carbon Interoverlapped (48) Vanin, M.; Mortensen, J. J.; Kelkkanen, A. K.; Garcia-Lastra, J.
Superstructure: Engineering Ideal Atomic Interface for Lithium Ion M.; Thygesen, K. S.; Jacobsen, K. W. Graphene on Metals: A van der
Storage. Adv. Mater. 2015, 27, 3687−3695. Waals Density Functional Study. Phys. Rev. B: Condens. Matter Mater.
(29) Worsley, M. A.; Shin, S. J.; Merrill, M. D.; Lenhardt, J.; Nelson, Phys. 2010, 81, 081408.
A. J.; Woo, L. Y.; Gash, A. E.; Baumann, T. F.; Orme, C. A. Ultra Low (49) Varykhalov, A.; Sanchez-Barriga, J.; Shikin, A. M.; Biswas, C.;
Density, Monolithic WS2, MoS2, and MoS2/Graphene Aerogels. ACS Vescovo, E.; Rybkin, A.; Marchenko, D.; Rader, O. Electronic and
Nano 2015, 9, 4698−4705. Magnetic Properties of Quasifreestanding Graphene on Ni. Phys. Rev.
(30) Li, H.; Yu, K.; Fu, H.; Guo, B.; Lei, X.; Zhu, Z. MoS2 /Graphene Lett. 2008, 101, 157601.
Hybrid Nanoflowers with Enhanced Electrochemical Performances as (50) Gruneis, A.; Vyalikh, D. V. Tunable Hybridization Between
Anode for Lithium-Ion Batteries. J. Phys. Chem. C 2015, 119, 7959− Electronic States of Graphene and a Metal Surface. Phys. Rev. B:
7968. Condens. Matter Mater. Phys. 2008, 77, 193401.
(31) Liu, Y.; Zhao, Y.; Jiao, L.; Chen, J. A Graphene-Like MoS2 (51) Gamo, Y.; Nagashima, A.; Wakabayashi, M.; Terai, M.; Oshima,
/Graphene Nanocomposite as a Highperformance Anode for Lithium C. Atomic Structure of Monolayer Graphite Formed on Ni(111). Surf.
Ion Batteries. J. Mater. Chem. A 2014, 2, 13109−13115. Sci. 1997, 374, 61−64.
(32) Zhang, W.; Chuu, C.; Huang, J.; Chen, C.; Tsai, M.; Chang, Y.; (52) Gowtham, S.; Scheicher, R. H.; Ahuja, R.; Pandey, R.; Karna, S.
Liang, C.; Chen, Y.; Chueh, Y.; He, J.; et al. Ultrahigh-Gain P. Physisorption of Nucleobases on Graphene: Density-Functional
Photodetectors Based on Atomically Thin Graphene-MoS2 Hetero- Calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 76,
structures. Sci. Rep. 2014, 4, 3826. 033401.

3274 DOI: 10.1021/acs.jpclett.5b01233


J. Phys. Chem. Lett. 2015, 6, 3269−3275
The Journal of Physical Chemistry Letters Letter

(53) Tran, F.; Laskowski, R.; Blaha, P.; Schwarz, K. Performance on


Molecules, Surfaces, and Solids of the Wu-Cohen GGA Exchange-
Correlation Energy Functional. Phys. Rev. B: Condens. Matter Mater.
Phys. 2007, 75, 115131.
(54) Sachs, B.; Wehling, T. O.; Katsnelson, M. I.; Lichtenstein, A. I.
Adhesion and Electronic Structure of Graphene on Hexagonal Boron
Nitride Substrates. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84,
195414.
(55) Hasegawa, M.; Nishidate, K. Semiempirical Approach to the
Energetics of Interlayer Binding in Graphite. Phys. Rev. B: Condens.
Matter Mater. Phys. 2004, 70, 205431.
(56) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the
Projector Augmented-Wave Method. Phys. Rev. B: Condens. Matter
Mater. Phys. 1999, 59, 1758−1775.
(57) Kresse, G.; Furthmuller, J. Efficiency of ab-Initio Total Energy
Calculations for Metals and Semiconductors Using a Plane-Wave Basis
Set. Comput. Mater. Sci. 1996, 6, 15−50.
(58) Kresse, G.; Furthmuller, J. Efficient Iterative Schemes for ab
Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys.
Rev. B: Condens. Matter Mater. Phys. 1996, 54, 11169−11186.
(59) Zhu, Z. Y.; Cheng, Y. C.; Schwingenschlogl, U. Giant Spin-
Orbit-Induced Spin Splitting in Two-Dimensional Transition-Metal
Dichalcogenide Semiconductors. Phys. Rev. B: Condens. Matter Mater.
Phys. 2011, 84, 153402.
(60) Baskin, Y.; Meyer, L. Lattice Constants of Graphite at Low
Temperatures. Phys. Rev. 1955, 100, 544.
(61) Ebnonnasir, A.; Narayanan, B.; Kodambaka, S.; Ciobanu, C. V.
Tunable MoS2 Bandgap in MoS2-Graphene Heterostructures. Appl.
Phys. Lett. 2014, 105, 031603.
(62) Kwak, J. Y.; Hwang, J.; Calderon, B.; Alsalman, H.; Munoz, N.;
Schutter, B.; Spencer, M. G. Electrical Characteristics of Multilayer
MoS2 FET’s with MoS2/Graphene Heterojunction Contacts. Nano
Lett. 2014, 14, 4511−4516.
(63) Castellanos-Gomez, A.; Barkelid, M.; Goossens, A. M.; Calado,
V. E.; van der Zant, H. S. J.; Steele, G. A. Laser-Thinning of MoS2: On
Demand Generation of a Single-Layer Semiconductor. Nano Lett.
2012, 12, 3187−3192.
(64) Scalise, E.; Houssa, M.; Pourtois, G.; Afanas’ev, V. V.; Stesmans,
A. Strain-Induced Semiconductor to Metal Transition in the Two-
Dimensional Honeycomb Structure of MoS2. Nano Res. 2012, 5, 43−
48.
(65) Li, T. Ideal Strength and Phonon Instability in Single-Layer
MoS2. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 85, 235407.
(66) Han, S. W.; Kwon, H.; Kim, S. K.; Ryu, S.; Yun, W. S.; Kim, D.
H.; Hwang, J. H.; Kang, J. S.; Baik, J.; Shin, H. J.; et al. Band-Gap
Transition Induced by Interlayer van der Waals Interaction in MoS2.
Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 045409.
(67) Conley, H. J.; Wang, B.; Ziegler, J. I.; Haglund, R. F.; Pantelides,
S. T.; Bolotin, K. I. Bandgap Engineering of Strained Monolayer and
Bilayer MoS2. Nano Lett. 2013, 13, 3626−3630.
(68) He, K.; Poole, C.; Mak, K. F.; Shan, J. Experimental
Demonstration of Continuous Electronic Structure Tuning via Strain
in Atomically Thin MoS2. Nano Lett. 2013, 13, 2931−2936.
(69) Kadantsev, E. S.; Hawrylak, P. Electronic Structure of a Single
MoS2 Monolayer. Solid State Commun. 2012, 152, 909−913.
(70) Ding, Y.; Wang, Y.; Ni, J.; Shi, L.; Shi, S.; Tang, W. First
Principles Study of Structural, Vibrational and Electronic Properties of
Graphene-Like MX2 (M = Mo, Nb, W, Ta; X = S, Se, Te) Monolayers.
Phys. B 2011, 406, 2254−2260.
(71) Qiu, D. Y.; da Jornada, F. H.; Louie, S. G. Optical Spectrum of
MoS2: Many-Body Effects and Diversity of Exciton States. Phys. Rev.
Lett. 2013, 111, 216805.
(72) Lebègue, S.; Eriksson, O. Electronic Structure of Two-
Dimensional Crystals from ab Initio Theory. Phys. Rev. B: Condens.
Matter Mater. Phys. 2009, 79, 115409.
(73) Xiao, D.; Liu, G.; Feng, W.; Xu, X.; Yao, W. Coupled Spin and
Valley Physics in Monolayers of MoS2 and Other Group-VI
Dichalcogenides. Phys. Rev. Lett. 2012, 108, 196802.

3275 DOI: 10.1021/acs.jpclett.5b01233


J. Phys. Chem. Lett. 2015, 6, 3269−3275

You might also like