Molecules 28 06122

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

molecules

Article
First Principles Study of the Photoelectric Properties of Alkaline
Earth Metal (Be/Mg/Ca/Sr/Ba)-Doped Monolayers of MoS2
Li-Zhi Liu 1 , Xian-Sheng Yu 1 , Shao-Xia Wang 2 , Li-Li Zhang 1, * , Xu-Cai Zhao 1, *, Bo-Cheng Lei 1 , Hong-Mei Yin 1
and Yi-Neng Huang 1,3

1 Xinjiang Laboratory of Phase Transitions and Microstructures in Condensed Matter Physics, College of
Physical Science and Technology, Yili Normal University, Yining 835000, China;
liulizhi9902@sina.com (L.-Z.L.); yuxs997@sina.com (X.-S.Y.); lbc0428@sina.com (B.-C.L.);
jisuan2000y@sohu.com (H.-M.Y.); ynhuang@nju.edu.cn (Y.-N.H.)
2 Physics and Electronic Engineering College, Kashi University, Kashi 844000, China; wwsx888666@sina.com
3 National Laboratory of Solid State Microstructures, School of Physics, Nanjing University,
Nanjing 210093, China
* Correspondence: suyi2046@sina.com (L.-L.Z.); zxc85619876@sina.com (X.-C.Z.)

Abstract: The energy band structure, density of states, and optical properties of monolayers of
MoS2 doped with alkaline earth metals (Be/Mg/Ca/Sr/Ba) are systematically studied based on
first principles. The results indicate that all the doped systems have a great potential to be formed
and structurally stable. In comparison to monolayer MoS2 , doping alkaline earth metals results in
lattice distortions in the doped system. Therefore, the recombination of photogenerated hole–electron
pairs is suppressed effectively. Simultaneously, the introduction of dopants reduces the band gap
of the systems while creating impurity levels. Hence, the likelihood of electron transfer from the
valence to the conduction band is enhanced, which means a reduction in the energy required for such
a transfer. Moreover, doping monolayer MoS2 with alkaline earth metals increases the static dielectric
constant and enhances its polarizability. Notably, the Sr–MoS2 system exhibits the highest value of
static permittivity, demonstrating the strongest polarization capability. The doped systems exhibit
Citation: Liu, L.-Z.; Yu, X.-S.; Wang,
a red-shifted absorption spectrum in the low-energy region. Consequently, the Be/Mg/Ca–MoS2
S.-X.; Zhang, L.-L.; Zhao, X.-C.; Lei,
systems demonstrate superior visible absorption properties and a favorable band gap, indicating
B.-C.; Yin, H.-M.; Huang, Y.-N. First
their potential as photo-catalysts for water splitting.
Principles Study of the Photoelectric
Properties of Alkaline Earth Metal
(Be/Mg/Ca/Sr/Ba)-Doped
Keywords: alkaline earth metals; MoS2 ; first principles; electronic structure; optical properties
Monolayers of MoS2 . Molecules 2023,
28, 6122. https://doi.org/10.3390/
molecules28166122
1. Introduction
Academic Editors: Qilin Hua and
Jianping Meng
Photocatalysts are frequently synthesized using materials with lower dimensions,
such as graphene [1,2], hexagonal boron nitride [2], transition metal sulfides [3,4], and
Received: 31 May 2023 g-C3 N4 [5], due to their high specific surface area, abundance of active catalytic sites, and
Revised: 30 July 2023 high electron–hole separation rate [6,7]. Moreover, monolayer molybdenum disulfide
Accepted: 9 August 2023 (MoS2 ) is a widely researched excellent representative of two-dimensional (2D) materi-
Published: 18 August 2023
als with graphene-like structures. MoS2 is highly valued in the fields of photocatalysis,
solar energy conversion, photovoltaics, and optoelectronics due to its optimal band gap,
exceptional carrier mobility, and superior current-switching performance [8–11]. MoS2 is a
layered material with an S–Mo–S sandwich structure, where a central layer of Mo atoms is
Copyright: © 2023 by the authors.
Licensee MDPI, Basel, Switzerland.
flanked by two layers of S atoms above and below [12]. Monolayer MoS2 possesses a direct
This article is an open access article
bandgap [13], enhancing its efficiency for water splitting. However, the efficient utilization
distributed under the terms and of sunlight by 2D materials is limited by quantum size effects [14]. Substitution doping is
conditions of the Creative Commons an effective step to enhance the electronic properties of 2D materials’ device applications,
Attribution (CC BY) license (https:// similar to their use in conventional semiconductors [15–19]. Monolayer MoS2 -doped Mn,
creativecommons.org/licenses/by/ Fe, Co, and Zn, for instance, as diluted magnetic semiconductors, could be applied in
4.0/). memory devices, store and read hard disks or MRAM [20,21]. P-type Au-doped MoS2 is

Molecules 2023, 28, 6122. https://doi.org/10.3390/molecules28166122 https://www.mdpi.com/journal/molecules


Molecules 2023, 28, 6122 2 of 15

applied in gas-sensing devices [22]. Re–MoS2 , Mn–MoS2 , Fe–MoS2 and Zn–MoS2 [16,23]
have been used as n-type semiconductors, and Nb–MoS2 has been excellently applied
in p-type semiconductors in optoelectronic devices, such as photodetectors and optical
modulators, etc. The compounds mentioned in the above paragraph have all been syn-
thesized experimentally. The experimental measurement analysis of the compounds has
demonstrated that doping transition metal elements can induce a transformation from the
intrinsic state to either n-type or p-type states in monolayer MoS2 , which allows the noble
compounds to be applied to different devices.
Moreover, the results of the first principles calculation and experimental test analyses
of Co2+ (or Ni2+ )-doped MoS2 all show that the introduction of Co2+ or Ni2+ results
in the enhancement of monolayer MoS2 photocatalysis [24,25]. The prediction of the
first principles computation of monolayer MoS2 ’s optical properties is consistent with
experimental results. The conclusion indicates that doping impurities enhance monolayer
MoS2 photocatalysis, and the first principles simulation method is a credible method.
The impurities mentioned in the above examples are transition metals. However,
transition metal doping cannot reduce the bandgap to a lower width because of its local
d orbital. The alkaline earth metals, including Be, Mg, Ca, Sr, Ba and others, constitute
the second group of elements in the periodic table. In contrast to transition metal doping,
alkaline earth metals do not involve localized d orbitals, so the doping of alkaline earth
metals can effectively reduce the optical threshold energy of semiconductors [26]. The
electronic structures of 2D materials with a graphene-like structure can be modified by
the incorporation of alkaline earth metals. Studies conducted by Ullah [27] et al. and
Tayyab M. [28] have demonstrated that doping graphene with Be and Mg can effectively
modulate its band gap. Kumar G. M. [29] employed a hydrothermal method to synthesize
Sn0.98 Mg0.02 S2 nanosheets and observed a red-shift in the absorption spectrum, indicating
an enhanced absorption of visible light. Ani A. et al. [30] proposed that Ca-doped SnSe2
monolayers exhibit superior photocatalytic capacity. Doping graphene, or two-dimensional
graphene-like materials, such as sulfur groups, with alkaline earth metals, has been demon-
strated to induce minor defects that impact the electronic structure. This ultimately leads
to an increase in carrier mobility and the improvement of absorption in visible light. It
is postulated that the incorporation of alkaline earth metals can effectively modulate the
photoelectric properties of monolayer MoS2 . However, a systematic investigation into the
impact of alkaline earth metals on the optoelectronic properties of monolayer MoS2 has
yet to be conducted. Therefore, it is worthwhile to delve deeply into the mechanisms and
regulatory laws governing its stable existence.
To sum up, monolayer MoS2 , a semiconductor material, has been widely studied
and applied, but the effective utilization of sunlight is limited by its electronic mobility,
which is restricted to the finite sizes of the synthesized monolayer MoS2 . It is found that
the doping of alkaline earth metals (Be/Mg/Ca/Sr/Ba) can change the electronic band
structure of the pristine materials and affect the transport of electrons. However, the effect
of the alkaline earth metal doping of MoS2 on the photoelectric performance of the system
has not been systematically theoretically studied. In our work, the energy band structure,
density of states, and optical properties of monolayer MoS2 doped with alkaline earth
metals were systematically studied based on first principles. The results show that the
Be/Mg/Ca-MoS2 systems had superior visible absorption performance and ideal band
gaps, and the potential to be used as photocatalyst candidates for water splitting.

2. Results
2.1. Analysis of Lattice Constants and Bond Population
The GGA-PBE calculation method was used to optimize the systems’ geometry. The
lattice constants optimized were a = b = 3.17 and c = 12.43; the errors relative to the
experimental values [31] and those of Hu [32] et al. were within 1%, which indicates
that the computational methods and parameters used in this paper are reliable. Table 1
summarizes the optimized crystal structures of monolayer MoS2 and X-MoS2 systems. The
Molecules 2023, 28, 6122 3 of 15

formation energy of the doped system is an indicator of the degree of formation difficulty.
A lower formation energy implies that the doped system more easily forms. The formation
energy is calculated as:

E f = Edoped − Eundoped + nµ Mo − nµ X (1)

Table 1. Formation energy, bond population and bond length of MoS2 systems pre- and post-doping.

Bond Pop- Bond Pop- Bond Bond


Formation ulation of ulation of Length of Length of 1 ∆γ/Å
Energy/eV Mo–S Mo–S Mo–S/Å Mo–S/Å
(Max) (Min) (Min) (Min)
MoS2 / 0.37 0.37 2.4083 2.4077 0.0006
Be-MoS2 10.92 0.52 0.34 2.4274 2.3369 0.0905
Mg-MoS2 7.27 0.47 0.36 2.4291 2.3960 0.0331
Ca-MoS2 12.17 0.50 0.34 2.4461 2.3901 0.0560
Sr-MoS2 13.45 0.49 0.35 2.4593 2.3841 0.0752
Ba-MoS2 14.60 0.47 0.34 2.4750 2.3775 0.0975
1 ∆γ is the difference between the maximum and minimum Mo–S bond lengths.

Edoped and Eundoped are the total energy of doped and undoped MoS2 , µ Mo and µ X are
the chemical potential of the Mo and X atoms, while n denotes the number of X atoms
doped into the supercell [10]. As shown in Table 1, the formation energy of the doped
systems decreased gradually with the decrease in the doped atoms’ radius (RBa > RSr > RCa
> RMg ), which suggest that the smaller the atomic radius of the doped system was, the more
easily the system formed. In Table 1, the absolute value of the formation energy of all the
doped systems is shown within a reasonable range. The compounds with an absolute value
of formation energy in that range can be synthesized via experiments [33–37]. Therefore,
it can be inferred that all the doped systems can be synthesized. The Mg–MoS2 system
exhibited the lowest formation energy among the five, indicating its superior suitability to
formation.
According to reference [38], the covalent properties of a chemical bond become
stronger as the value of the bond population increases. The results presented in Table 1
demonstrate that the maximum population of Mo–S bonds in each doped system surpassed
that of the monolayer MoS2 system, while the minimum Mo–S bonding population in all
doping systems was lower than that of the intrinsic system. However, the maximum value
varied much more than the minimum value, which demonstrates that the stability of the
doped systems was not destroyed by the introduction of impurity atoms.
The higher the ∆γ value, the more pronounced the distortion of the corresponding
system becomes [39]. In the table, we can see that the ∆γ values of all doped systems were
greater than that of a pure one, which implies that all the systems were distorted, and this
aligns with the conclusion that the band gaps of the doped monolayer MoS2 were reduced.
Moreover, Ba–MoS2 had the greatest degree distortion, which corresponds to its having the
smallest bap gap. The optimized Mo–S bond length of 2.41 coheres with the one Li et al.
reported [40]. It was observed that the ∆γ values of the doped systems, except Be–MoS2 ,
increased as the radius of the dopant atom increased. It was suggested that the degree
of lattice distortion also increased with the dopant atom’s radius. The lattice distortion
caused a separation between the positive and negative charge centers within the defect,
which resulted in the creation of an internal electric dipole moment and a local potential
difference. The introduction of alkaline earth metal elements was found to increase the rate
of separation of photogenerated electron–hole pairs and decrease the energy needed for
electron leap, as reported in reference [41].
In order to further verify its thermal stability, the AIMD (Ab Initio Molecular Dynam-
ics) [42] analysis of monolayer MoS2 molybdenum and X–MoS2 systems was performed,
and the temperature was set to room temperature (300 K), as shown in Figure 1. It can be
Table 1. Formation energy, bond population and bond length of MoS2 systems pre- and
post-doping.

Formation Bond Population of Bond Population of Bond Length of Bond Length of 1


Δγ/Å
Energy/eV Mo–S (Max) Mo–S (Min) Mo–S/Å (Min) Mo–S/Å (Min)
MoS
Molecules 2023,
2 28, 6122 / 0.37 0.37 2.4083 2.4077 0.0006
4 of 15
Be-MoS2 10.92 0.52 0.34 2.4274 2.3369 0.0905
Mg-MoS2 7.27 0.47 0.36 2.4291 2.3960 0.0331
Ca-MoS2 12.17 0.50 0.34 2.4461 2.3901 0.0560
seen that after 8000 steps of AIMD, with a time step of 1 fs, the structures of the six systems
Sr-MoS2 13.45 remained 0.49
stable—there was no0.35 2.4593
bond breaking, and 2.3841 was very
the energy fluctuation 0.0752
small.
Ba-MoS2 14.60 0.47show that the monolayer
The results 0.34 MoS2 and X-MoS2.4750
2 systems have2.3775
high stability0.0975
at room
temperature (300 K)between
1 Δγ is the difference [43,44].the maximum and minimum Mo–S bond lengths.

-345
-334.5 -333.5
-346 -335.0
-334.0
-347 -335.5
T o ta l E n e r g y (e V )

T o ta l E n e r g y (e V )
-334.5

T o ta l E n e r g y (e V )
-336.0
-348
-335.0
-336.5
-349
-337.0 -335.5
-350 -337.5 -336.0
-351 -338.0 -336.5
-338.5
-352 -337.0
-339.0
0 2000 4000 6000 8000 0 2000 4000 6000 8000 0 2000 4000 6000 8000
Time(fs) Time(fs) Time(fs)

(a) (b) (c)


-335.0 -332.0
-333.5 -332.5
-335.5 -333.0
-334.0
-333.5
-334.5 -334.0
T o ta l E n e r g y (e V )
-336.0
T o ta l E n e r g y (e V )

-334.5
T o ta l E n e r g y (e V )

-335.0 -335.0
-336.5
-335.5
-335.5
-336.0
-337.0 -336.5
-336.0
-337.0
-337.5 -336.5 -337.5
-337.0 -338.0
-338.0 -338.5
-337.5 -339.0
0 2000 4000 6000 8000 0 2000 4000 6000 8000 0 2000 4000 6000 8000
Time(fs) Time(fs) Time(fs)

(d) (e) (f)


Figure 1.
Figure 1. AIMD
AIMD (300
(300 K)
K) of
ofMoS
MoS2 systems
systems pre-
pre- and
and post-doping:
post-doping: (a)
(a) monolayer
monolayer MoS
MoS2;; (b)
(b) Be–MoS
Be–MoS2;;
2 2 2
(c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2. (yellow, blue and red colors represents S, Mo
(c) Mg–MoS2 ; (d) Ca–MoS2 ; (e) Sr–MoS2 ; (f) Ba–MoS2. (yellow, blue and red colors represents S, Mo
and X, respectively).
and X, respectively).
2.2. Band
2.2. Band Structures
Structures
The band
The band structures
structures ofof the
the monolayer
monolayer MoS and X–MoS
MoS22 and X–MoS22 systems
systems are
are presented
presented in in
Figure 2. The Fermi energy level was set at 0 eV, and the energy range
Figure 2. The Fermi energy level was set at 0 eV, and the energy range of −3~3 eV was of −3~3 eV was
selected for
selected for comparative
comparative analysis.
analysis. The band structure
The band structure of
of the
the monolayer
monolayer MoS
MoS22 system
system is is
illustrated ininFigure
illustrated Figure
2a; 2a; the conduction
the conduction band minimum
band minimum (CBM) and (CBM) andband
valence valence band
maximum
maximum
(VBM) (VBM)
are both are both
located located
at the high at the highpoint
symmetry symmetry
K. Thepoint K. The semiconductor
semiconductor exhibited a
exhibited
direct bandagapdirect band
of 1.75 eV. gap
This of 1.75was
value eV.only
This value
0.57% off w as only
from 0.57% off from
the experimental valuethe
of
experimental
1.74 eV [45], andvalue
wasofsimilar
1.74 eV [45],calculated
to the and was value
similar
of to the
1.78 eVcalculated value
found in the of 1.78[46].
literature eV
found in the literature [46].
band gap, indicating a minimal energy requirement for electron transition in this system.
Therefore, it can be inferred that Ba–MoS2 boasts superior optical absorption capabilities.
In the doped systems exposed to visible light, valence electrons were induced to
transition between valence and impurity levels as well as conduction bands, leading to
the generation of photonic electron–hole pairs. The relatively independent impurity
Molecules 2023, 28, 6122 levels effectively reduced the probability of the electron–hole pairs’ recombination,5 of 15
and
enhanced optical absorption capacity.

3 3 3

2 2 2

1 1 1
Eg=1.75eV Eg=1.73eV

Energy(eV)
Energy(eV)
Eg=1.68eV
Energy(eV)

0 Ef 0 Ef 0 Ef

-1 -1 -1

-2 -2 -2

-3 -3 -3
Γ M K Γ Γ M K Γ Γ M K Γ

(a) (b) (c)


3 3 3

2 2 2

1 1 1
Eg=1.62eV
Energy(eV)

Energy(eV)
Energy(eV)

Eg=1.48eV
0 Ef
Eg=1.25eV
0 Ef 0 Ef

-1 -1 -1

-2 -2 -2

-3 -3 -3
Γ M K Γ Γ M K Γ Γ M K Γ

(d) (e) (f)


Figure 2.
Figure 2. Band
Band structures
structures of
of MoS
MoS2 systems
systems pre-
pre- and
and post-doping:
post-doping: (a)
(a) monolayer
monolayer MoS
MoS2;; (b)
(b) Be–MoS
Be–MoS2;;
2 2 2
(c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2.
(c) Mg–MoS2 ; (d) Ca–MoS2 ; (e) Sr–MoS2 ; (f) Ba–MoS2 .

Figure 2b–f illustrates the band structure of MoS2 doped with alkaline earth metals
in a clear and concise manner, providing valuable insights into its electronic properties.
The CBM and VBM of the doped systems exhibited a downward shift in energy compared
to those of the monolayer MoS2 . The X–MoS2 systems had band gaps of 1.68, 1.73, 1.62,
1.48, and 1.25 eV, which were observed to be reduced as the atomic numbers of doped
alkali earth metal elements increased, except for Be. The decrease in the band gap indicated
an effective reduction in the energy required for electron transition, thereby enhancing
electron mobility. The doped systems displayed a series of impurity energy levels that
were relatively independent, acting as an intermediary for valence band electrons to
transition into the conduction band and enhancing the probability of electron transition
to the conduction band. The doped systems were found to possess an enhanced optical
absorption capacity. Notably, Ba–MoS2 exhibited the smallest direct band gap, indicating
a minimal energy requirement for electron transition in this system. Therefore, it can be
inferred that Ba–MoS2 boasts superior optical absorption capabilities.
In the doped systems exposed to visible light, valence electrons were induced to
transition between valence and impurity levels as well as conduction bands, leading to
the generation of photonic electron–hole pairs. The relatively independent impurity levels
effectively reduced the probability of the electron–hole pairs’ recombination, and enhanced
optical absorption capacity.
more significant. The introduction of impurities into the materials led to a reduction in
the band gaps, which is consistent with the results obtained from band structure analysis.
The valence bands of the Sr/Ba-doped MoS2 systems were determined by the Mo–4d, S–
3p, and S–3s states, while the Ba–5p state manifested a weak peak in the lower valence
Molecules 2023, 28, 6122
band of the Ba–MoS2 system, similarly to the results of other doped systems. The 6band of 15
gap of the Ba-doped system is the smallest. In Section 2.1, we can see that the Δγ value of
Ba–MoS2 was the smallest, which demonstrates that Ba–MoS2 manifested the greatest
degree of lattice distortion. This is why Ba–MoS2 had the smallest bandgap [10].
2.3. Density of States
The contributions of S or d-state electrons from doped elements near and above the
FermiThe density
level of state (DOS)
are particularly plotsthey
obvious; for monolayer MoS
determine the 2 and X–MoS
impurity systems
level,2which are pre-
causes the
sented in Figure
conduction band 3,minimum
encompassing
(CBM)energies ranging
and valence from
band −3 to 3 (VBM)
maximum eV. In Figure
in the 3a, the
doped
densities
systems toofshift
states in a significantly,
down monolayer MoS 2 are depicted,
reducing the band wherein
gap, andthe conduction
ultimately band
leading to isa
dominated by Mo–4d, Mo–5s, and S–3p states, and the valence band is dominated
decrease in photon excitation energy and a red shift in the predicted absorption spectrum by
Mo–4d, S–3s, and
of the doped systems.S–3p states. Both the CBM and VBM were mainly determined by the
Mo–4d and S–3p states, which is consistent with what has been found in the literature [10].

120 90 90
TDOS
TDOS 60 60 TDOS
80
30 30
40 0 0
60 Mo 5s 60 Mo 5s
0 Mo 4p Mo 4p
80
DOS/eV-1

30 Mo 4d
DOS/eV-1

30 Mo 4d
Mo 5s

DOS/eV-1
60
Mo 4p 0 0
40 Mo 4d 60 60
S 3s S 3s
20 40 S 3p 40 S 3p
0 20 20
80
S 3s 0 0
60 0.6 Be 2s Mg 3s
S 3p 0.6
Mg 2p
40
0.3 0.3
20
Molecules
0 2023, 28, x FOR PEER REVIEW
-3 -2 -1 0 1 2 3 0.0 0.0
-3 -2 -1 0 1 2
7 of 316
-3 -2 -1 0 1 2 3
Energy(eV) Energy(eV) Energy(eV)

(a) (b) (c)


90 90 90
TDOS TDOS TDOS
60 60 60
30 30 30
0 0 0
60 60 60
Mo 5s Mo 5s Mo 5s
Mo 4p 40 Mo 4p 40 Mo 4p
30 Mo 4d
Mo 4d Mo 4d
DOS/eV-1

20 20
DOS/eV-1

DOS/eV-1

0 0 0
40 S 3s 50 S 3s 40 S 3s
S 3p S 3p S 3p
20 25 20

0 0 0 Ba 6s
0.8 Ca 4s 2.0 Sr 5s 2
Ca 3p 1.6 Sr 4p Ba 5p
0.6
1.2 Sr 3d Ba 4d
0.4 1
0.8
0.2 0.4
0.0 0.0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
Energy(eV) Energy(eV) Energy(eV)

(d) (e) (f)


Figure
Figure 3. Density of
3. Density ofstates
statesofofMoS
MoSsystems
2 systems pre- and post-doping: (a) monolayer MoS2; (b) Be–
pre- and post-doping: (a) monolayer MoS2 ; (b) Be–MoS2 ;
2
MoS2; (c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2.
(c) Mg–MoS2 ; (d) Ca–MoS2 ; (e) Sr–MoS2 ; (f) Ba–MoS2 .

To investigate
Figure the microscopic
3b–f display the DOS plotschanges
of thein MoS
energy levels of each system, we created
2 systems doped with Be, Mg, Ca, Sr,
LUMO (Lowest Unoccupied
and Ba, respectively. Molecular
As illustrated in FigureOrbital)
3b–d, in and HOMO (Highestsystems,
the Be/Mg/Ca-doped Occupied the
Molecular
valence bands Orbital) diagrams
were mainly for the ofmonolayer
composed Mo–4d andMoS 2 and
S–3p. The X–MoS
CBM was 2 systems.
primarilyFigure
derived 4
shows the LUMO
from Mo–4d and HOMO,
and S–3p represented
states, with by the red and
minor contributions green
from areas,
Be–2s, respectively.
Mg–2p, Mg–3s, The
and
isosurface
Ca–4s states. wasThe
set Ca–4s
as 0.04state
e/Å3.had
Thealmost
larger no
theeffect
area, onthethe
greater its ability
conduction band.to gain (lose)
The Be–2s,
electrons. The LUMO orbitals possessing lower energy levels were more
Mg–3s, and Ca–4s states led to impurity levels near and above the Fermi level. The band conducive to the
obtaining
gaps of theofmaterials
electrons,werewhereas the HOMO
narrowed orbitals with
due to impurity higher
doping, whichelectron energies
is consistent were
with the
more
result likely to show
regarding band electron
structure.loss
The[47].
DOSInvalues
Figure 4a, Sr/Ba-doped
of the we can see that MoSthe LUMOreveal
2 systems and
HOMO of theofmonolayer
that the effects MoS2 states
Sr–3d and Ba–3d were on distributed
the conductionsymmetrically, indicating
band were more that The
significant. the
centers of the of
introduction positive and negative
impurities charges were
into the materials led toina the same position,
reduction andgaps,
in the band therewhich
was no is
internal
consistent electric field.
with the According
results obtained tofrom
Figure 4b–f,
band compared
structure to the
analysis. Themonolayer MoS2of
valence bands , the
the
LUMO and HOMO
Sr/Ba-doped maps ofwere
MoS2 systems the determined
doped systems by thewere redistributed,
Mo–4d, S–3p, and mostly clustered
S–3s states, while
the Ba–5p
around thestate
Mo andmanifested
S atomsa near
weakthe
peak in theatoms.
dopant lower Invalence
Figureband of the
4b, the Ba–MoS
area system,
of the2HOMO
of Mo is greater than that of the LUMO of S atoms near the Be atom, indicating that Mo
and S lost and gained electrons, respectively, and electrons were easily transferred from
the Mo atom to the S atom. In Figure 4c, we can see that the Mo atoms of the Mg–MoS2
had the greatest ability to gain electrons, while in the Ca–MoS2 system, Mo and S showed
TDOS TDOS TDOS
60 60 60
30 30 30
0 0 0
60 60 60
Mo 5s Mo 5s Mo 5s
Mo 4p 40 Mo 4p 40 Mo 4p
30 Mo 4d
Mo 4d Mo 4d

DOS/eV-1
20 20
DOS/eV-1

DOS/eV-1
Molecules
0 2023, 28, 6122 0 0 7 of 15
40 S 3s 50 S 3s 40 S 3s
S 3p S 3p S 3p
20 25 20

0 0 0 Ba 6s
0.8
0.6
Ca 4s
Ca 3p
similarly to the2.0 results of other dopedSrSrsystems.
1.6
5s
4p
The band2 gap of the Ba-dopedBasystem 5p is the
0.4 smallest. In Section
1.2
2.1, we can see Sr 3d the ∆γ value of1 Ba–MoS was the smallest,
that Ba 4d
which
0.8 2
0.2
0.0
demonstrates that 0.4
0.0
Ba–MoS2 manifested the greatest degree 0 of lattice distortion. This is why
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3
Energy(eV) Ba–MoS2 had the smallest bandgap
Energy(eV) [10]. Energy(eV)
The contributions of S or d-state electrons from doped elements near and above the
(d) Fermi level are particularly (e) obvious; they determine the impurity(f) level, which causes
the conduction band minimum (CBM) and valence band maximum (VBM) in the doped
Figure 3. Density of states of MoS2 systems pre- and post-doping: (a) monolayer MoS2; (b) Be–
systems to shift down significantly, reducing the band gap, and ultimately leading to a
MoS2; (c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2.
decrease in photon excitation energy and a red shift in the predicted absorption spectrum
of the doped systems.
To investigate the microscopic changes in energy levels of each system, we created
To investigate the microscopic changes in energy levels of each system, we created
LUMO (Lowest Unoccupied Molecular Orbital) and HOMO (Highest Occupied
LUMO (Lowest Unoccupied Molecular Orbital) and HOMO (Highest Occupied Molecular
Molecular Orbital) diagrams for the monolayer MoS2 and X–MoS2 systems. Figure 4
Orbital) diagrams for the monolayer MoS2 and X–MoS2 systems. Figure 4 shows the
shows the LUMO and HOMO, represented by the red and green areas, respectively. The
LUMO and HOMO, represented by the red and green areas, respectively. The isosurface
isosurface was set as 0.04 e/Å3. The larger the area, the greater its ability to gain (lose)
was set as 0.04 e/Å3 . The larger the area, the greater its ability to gain (lose) electrons.
electrons. The LUMO orbitals possessing lower energy levels were more conducive to the
The LUMO orbitals possessing lower energy levels were more conducive to the obtaining
obtaining of electrons, whereas the HOMO orbitals with higher electron energies were
of electrons, whereas the HOMO orbitals with higher electron energies were more likely
more likely to show electron loss [47]. In Figure 4a, we can see that the LUMO and
to show electron loss [47]. In Figure 4a, we can see that the LUMO and HOMO of the
HOMO of MoS
monolayer the monolayer MoS2 were distributed symmetrically, indicating that the
2 were distributed symmetrically, indicating that the centers of the positive
centers of the positive and
and negative charges were in the same negative charges were and
position, in the
theresamewas position,
no internaland there
electric was no
field.
internal electric field. According to Figure 4b–f, compared
According to Figure 4b–f, compared to the monolayer MoS2 , the LUMO and HOMO maps to the monolayer MoS 2, the

LUMO
of and systems
the doped HOMO were mapsredistributed,
of the doped systems
mostly were around
clustered redistributed,
the Mo mostly
and S atomsclustered
near
the dopant atoms. In Figure 4b, the area of the HOMO of Mo is greater than thatHOMO
around the Mo and S atoms near the dopant atoms. In Figure 4b, the area of the of the
of Mo isofgreater
LUMO S atoms than
near that
theofBethe LUMO
atom, of S atoms
indicating thatnear
Mo and the Be atom,
S lost andindicating that Mo
gained electrons,
and S lost and gained electrons, respectively, and electrons
respectively, and electrons were easily transferred from the Mo atom to the S atom. were easily transferred from
In
the Mo atom to the S atom. In Figure 4c, we can see that the
Figure 4c, we can see that the Mo atoms of the Mg–MoS2 had the greatest ability to gain2 Mo atoms of the Mg–MoS
had the greatest
electrons, while in ability to gain 2electrons,
the Ca–MoS system, Mo while
andinS the Ca–MoS
showed 2 system,
almost equalMo and Sgain
electron showedand
almost
loss. equal
The Mo electron
atoms ingain and loss.2 and
the Sr–MoS The Mo atoms
Ba–MoS 2
in the
systems Sr–MoS
showed 2 and
the Ba–MoS
greatest systems
2 electron

showed
loss. the greatest
The observed chargeelectron
transfer loss.
and The observed
asymmetric charge transfer
distribution and asymmetric
are in agreement with the
distribution are in agreement with
findings of the static dielectric constant below. the findings of the static dielectric constant below.

Molecules 2023, 28, x FOR PEER REVIEW 8 of 16

(a) (b) (c)

(d) (e) (f)


Figure 4. LUMO and HOMO of MoS2 systems before and after doping: (a) LUMO and HOMO of
Figure 4. LUMO and HOMO of MoS2 systems before and after doping: (a) LUMO and HOMO of
monolayer MoS2; (b) Be–MoS2; (c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2 (red and green
monolayer MoS2 ; (b) Be–MoS2 ; (c) Mg–MoS2 ; (d) Ca–MoS2 ; (e) Sr–MoS2 ; (f) Ba–MoS2 (red and green
filled areas indicate LUMO and HOMO, respectively).
filled areas indicate LUMO and HOMO, respectively).
2.4. Work Functions
To investigate charge transfer in monolayer MoS2 before and after doping, the work
functions of monolayer MoS2 and X–MoS2 systems were calculated. The work function, as
illustrated in Formula (2), is defined as the minimum energy required for an electron to
(d) (e) (f)
Figure 4. LUMO and HOMO of MoS2 systems before and after doping: (a) LUMO and HOMO of
monolayer MoS2; (b) Be–MoS2; (c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2 (red and green
Molecules 2023, 28, 6122 8 of 15
filled areas indicate LUMO and HOMO, respectively).

2.4. Work Functions


2.4. Work Functions charge transfer in monolayer MoS2 before and after doping, the work
To investigate
To investigate
functions of monolayercharge
MoS transfer in monolayer
2 and X–MoS 2 systems MoS 2 before
were and after
calculated. doping,
The work the work
function, as
functions ofinmonolayer
illustrated MoS
Formula (2), and X–MoS
is2 defined 2 systems
as the minimum were calculated.
energy The for
required workan function,
electron toas
illustratedfrom
transition in Formula (2),energy
the Fermi is defined
levelastothe
theminimum energy
vacuum layer [48].required for an electron to
transition from the Fermi energy level to the vacuum layer [48].
Ф=Е −Е (2)
Φ = Evac − EF (2)
where Φ represents work function, Е is vacuum level, and Е denotes Fermi level.
T he work
where function ofwork
Φ represents the monolayer
function, E MoS 2 was 4.79 eV, as shown in Figure 5a, which is
vac is vacuum level, and E F denotes Fermi level.
consistent
The work function of the monolayer MoS2 wasby
with the value of 4.89 eV calculated Wei
4.79 eV,[49]. The Be/Mg/Ca/Sr/Ba-doped
as shown in Figure 5a, which is
systems
consistent with the value of 4.89 eV calculated by Wei [49]. TheMoS
had lower work functions than that of the monolayer 2 (4.42, 4.35, 4.65, 4.54,
Be/Mg/Ca/Sr/Ba-doped
and 4.64 eV, respectively). It was shown that doping with alkaline
systems had lower work functions than that of the monolayer MoS2 (4.42, earth metals
4.35,effectively
4.65, 4.54,
reduced therespectively).
and 4.64 eV, energy required for electrons
It was shown that dopingto with
escape fromearth
alkaline the metals
interior to the
effectively
semiconductor
reduced the energysurface. Hence,
required forthe doped to
electrons systems
escapeexhibited a better to
from the interior capacity for electron
the semiconductor
transition.
surface. Hence, the doped systems exhibited a better capacity for electron transition.

10 10 10
Vacuum Level Vacuum Level Vacuum Level

0 F 0 F 0 F
Electrostatic Potential(eV)

Electrostatic Potential(eV)
Feimi Level Feimi Level
Electrostatic Potential(eV)

Feimi Level

-10 F=4.79eV -10 F=4.42eV -10 F=4.35eV

-20 -20 -20

-30 -30 -30

Molecules 2023, 28, x FOR PEER REVIEW 9 of 16


-40 -40 -40
Z-axis direction Z-axis direction Z-axis direction

(a) (b) (c)


10 10 10
Vacuum Level Vacuum Level Vacuum Level

0 F 0 F 0 F
Electrostatic Potential(eV)
Electrostatic Potential(eV)

Electrostatic Potential(eV)

Feimi Level Feimi Level Feimi Level


F=4.54eV F=4.64eV
-10 F=4.65eV -10 -10

-20 -20 -20

-30 -30 -30

-40 -40 -40


Z-axis direction Z-axis direction Z-axis direction

(d) (e) (f)


Figure 5. Work
Figure 5. Workfunctions
functionsofofMoS
MoSsystems
2 systems before and after doping: (a) monolayer MoS2; (b) Be–
before and after doping: (a) monolayer MoS2 ; (b) Be–MoS2 ;
2
MoS2; (c) Mg–MoS2; (d) Ca–MoS2; (e) Sr–MoS2; (f) Ba–MoS2.
(c) Mg–MoS2 ; (d) Ca–MoS2 ; (e) Sr–MoS2 ; (f) Ba–MoS2 .

2.5. Optical
2.5. Optical Properties
Properties
Figure 6a
Figure 6a shows
shows the
the real part of the dielectric function (ε11) of monolayer MoS22 and and
X–MoS22 systems.
X–MoS systems. The
Thereal
realpart
part reflects
reflects the
the polarization
polarization strength
strength ofof the
the semiconductor
semiconductor
material in
material in the
the presence
presence ofof an
an external
external electric
electric field
field with
with aa change
change in in incident
incident photon
photon
energy. A higher value of ε indicates an enhancement in the capacity of
energy. A higher value of ε11 indicates an enhancement in the capacity of the systems to the systems to
bind charges,
bind charges,thereby
therebyincreasing
increasingtheir polarizability
their [50].[50].
polarizability When the incident
When photonphoton
the incident energy
was 0 eV,
energy wasthe systems
0 eV, exhibited
the systems a static dielectric
exhibited constantconstant
a static dielectric (ε0 ). From
(ε0).Figure
From 6a, we can
Figure see
6a, we
can see that the ε0 values of the monolayer MoS2 and the doped systems were 5.88, 10.31,
8.89, 7.05, 18.20, and 14.94, respectively. In comparison to the monolayer MoS2, all the
doped systems showed an increase in values of ε0, indicating an enhancement in the
polarizability of the X–MoS2 systems. Therefore, in doped systems, the migration and
0
0
0 1 2 3 4 0 2 4 6 8 10
Energy(eV) Energy(eV)

(a) (b)
Molecules 2023, 28, 6122 9 of 15
Figure 6. Dielectric function diagram of MoS2 systems before and after doping: (a) real part and (b)
imaginary part.

Thetheabsorption
that ε0 values ofspectra and refractive
the monolayer MoS2 and indices of monolayer
the doped systems were MoS 2 and
5.88, X–MoS
10.31, 8.89, 27.05,
systems are presented in Figure 7. As depicted in Figure 7a, in comparison
18.20, and 14.94, respectively. In comparison to the monolayer MoS2 , all the doped systems to the
monolayer
showed an MoS 2, the in
increase doped
valuessystems displayed
of ε0 , indicating notable redshifts,
an enhancement in the and their light
polarizability of the
responses
X–MoSwere2 extended
systems. to the
Therefore, low-energy
in doped region.
systems, Therefore,
the migrationthe doped
and systems
separation showed
rates of joint
a better light absorption
electron–hole capacity.
pairs were The Sr–MoS
accelerated, leading2 and
to anBa–MoS 2 systems
improvement displayed
in the the same
optical performance
absorption peaks The
of the system. in the low-energy
Sr–MoS 2 systemregion (0–1.6
exhibited eV), demonstrating
the highest the two
values, indicating systems
that it possessed
hadthe
strong capacity to use visible light.
strongest polarization, excitation strength, and charge binding capability.

10
pure
4 Be-MoS2
MoS2
8 Mg-MoS2
Be-MoS2
A b so rp tio n /1 0 4 cm -1

Ca-MoS2
Mg-MoS2 3 Sr-MoS2

Refractive
6 Ca-MoS2 Ba-MoS2
Sr-MoS2
Ba-MoS2 2
4

2 1

0 0
0 1 2 3 4 0 2 4 6 8 10
Energy(eV) Energy(eV)

(a) (b)
Figure 7. (a)
Figure Light absorption
6. Dielectric functiondiagram,
diagram (b) refractive
of MoS index of MoS2 systems before and after
2 systems before and after doping: (a) real part and
doping.
(b) imaginary part.

Based on the 6b,


In Figure refractive index curves
the imaginary depicted
parts of in Figure
the dielectric 7b, it (ε
function can be the
2 ) of inferred that the
monolayer MoS2
static
andrefractive index ofare
doped systems theshown,
monolayer MoS
varying 2 was
with 2.452 when
incident photontheenergy.
photonThe incident
larger energy
the values
wasof0 εeV. This
2 , the value the
greater is consistent
probability with the theoretical
of photon valueand
absorption, of 2.475 [51]. After
the stronger the doping,
absorption
the energy
static refractive index values
of the materials. of the of
The value X–MoS systems were
ε2 is 2dependent all number
on the greater than that of the
of electrons in the
pure system.
excited However,
state, with thethe
which increases increase in incidence
likelihood of the nextphoton energy,
transition the refractive
occurring [50]. It can
be observed
indices that thetended
of all systems ε2 of the
to monolayer
overlap andMoS 2 displayed
approach a dielectric
unity, absorption
which indicates that peak
the at
approximately
monolayer 2.90 eV. This
MoS2 remained peak can
transparent be before
both primarilyandattributed to the interband transition
after doping.
between the S–3p and Mo–4d states of electrons. The dielectric peaks of ε2 in the X–MoS2
2.6. systems were shifted towards the low-energy region compared to the monolayer MoS2 .
Band Alignment
InTo
theunderstand
low-energythe region,
effectsthe of
intensity
alkalineof earth
the imaginary peak was
metal doping on significantly higher in
the photocatalytic
the doped systems than in the intrinsic system. Because of
capacity of the monolayer MoS2, the band alignments of both the monolayer MoS2 anda higher number of electrons
X–
present in the impurities formed by the doped system near the Fermi level, an enhancement
in the light absorption of the doped systems in the low-energy region was observed. In
the X–MoS2 systems, ε2 exhibited a significant response at a photon energy of 0 eV due
to the partial occupation of energy levels of dopant atoms and the impurity energy levels
straddling the Fermi energy levels. The presence of dopant elements facilitated electron
transfer between multiple degenerate impurity levels, thereby enhancing the low-energy
photon absorption of MoS2 .
The absorption spectra and refractive indices of monolayer MoS2 and X–MoS2 systems
are presented in Figure 7. As depicted in Figure 7a, in comparison to the monolayer MoS2 ,
the doped systems displayed notable redshifts, and their light responses were extended
to the low-energy region. Therefore, the doped systems showed a better light absorption
capacity. The Sr–MoS2 and Ba–MoS2 systems displayed the same absorption peaks in the
low-energy region (0–1.6 eV), demonstrating the two systems had strong capacity to use
visible light.
Molecules 2023, 2023,
Molecules 28, x FOR PEER REVIEW
28, 6122 10 of 16
10 of 15

20
pure 7 pure
Be-MoS2 Be-MoS2
16 Mg-MoS2 6 Mg-MoS2

Dielectic function (Im)


Dielective function(Re)
Ca-MoS2 Ca-MoS2
Sr-MoS2 5 Sr-MoS2
12
Ba-MoS2 Ba-MoS2
4

8 3

2
4
1
0
0
0 1 2 3 4 0 2 4 6 8 10
Energy(eV) Energy(eV)

(a) (b)
Figure 6. Dielectric
Figure function
7. (a) Light diagram
absorption of MoS(b)
diagram, 2 systems before
refractive and
index after2doping:
of MoS systems(a) real and
before partafter
and (b)
doping.
imaginary part.
Based on the refractive index curves depicted in Figure 7b, it can be inferred that the
static absorption
The spectra
refractive index and
of the refractiveMoS
monolayer indices
2 was of monolayer
2.452 when theMoS 2 and
photon X–MoS
incident energy
2

systems
was 0are eV.presented
This value in Figure 7. with
is consistent As depicted in Figure
the theoretical value7a, in comparison
of 2.475 to the the
[51]. After doping,
static refractive
monolayer index
MoS2, the valuessystems
doped of the X–MoS 2 systems
displayed were redshifts,
notable all greater and
than that
theiroflight
the pure
system.
responses However,
were with
extended to the
the increase in incidence
low-energy photon energy,
region. Therefore, the refractive
the doped systems indices
showedof all
systems
a better light tended to overlap
absorption capacity.andTheapproach
Sr–MoSunity, which indicates
2 and Ba–MoS 2 systems that the monolayer
displayed the same MoS2
remained
absorption transparent
peaks both beforeregion
in the low-energy and after doping.
(0–1.6 eV), demonstrating the two systems
had strong capacity to use visible light.
2.6. Band Alignment
10 To understand the effects of alkaline earth metal doping on the photocatalytic capacity
pure
of the monolayer MoS2 , the band alignments of 4 both the monolayer MoS 2 and
Be-MoS 2
X–MoS2
MoS2
systems
8 were calculated, and are illustrated in Figure 8. We used the concept
Mg-MoS of
2 semicon-
Be-MoS2
A b so rp tio n /1 0 4 cm -1

ductor electronegativity [52] to determine the redox and oxidation potentials for2 each water
Ca-MoS
Mg-MoS2 3 Sr-MoS2
system. The position of the band edge was calculated using the Milligan electronegativity
Refractive

6 Ca-MoS2 Ba-MoS2
equation,Sr-MoS
which is illustrated in [53–55].
2

Ba-MoS2 2
4
EVBM = χ − Eelec + 0.5Eg (3)

2 1
ECBM = EVBM − Eg (4)
0 In this equation, χ represents the geometric 0 mean of the Milligan electronegativity
0
of all atoms,1 Eelec 2
represents
Energy(eV)
3
the 4
free electron 0
energy 2 4.5 eV
of 4 on the
Energy(eV)
6 standard
8 10hydrogen

electrode scale, and Eg represents the semiconductor band gap [56]. This method did not
(a) (b)
provide exact values, but enabled a general evaluation of each system’s position on the
Figurenormal
7. (a) hydrogen electrode
Light absorption scale.(b)
diagram, Therefractive
band edges indexofofmonolayer
MoS2 systems MoSbefore
2 were suitable
and after for
doping.
a semiconductor photocatalyst, and this is consistent with experimental results [57]. Yet
the oxidation and reduction potential of Sr/Ba–MoS2 systems were too low to generate a
Based on the
sufficient refractive
driving force index
for O2curves
and Hdepicted
2 productionin Figure 7b, it can
compared be inferred
to standard that the
water-splitting
staticredox
refractive indexIt of
potential. was thesuggested
monolayer thatMoS was 2.452 when
the2Sr/Ba–MoS 2 the
systems photon
needed incident
to elevate energy
their CBM
was and
0 eV.VBMThispositions
value is consistent
relative to thewithwater
the theoretical value of The
splitting potential. 2.475 [51]. After
Be–MoS 2 , doping,
Mg–MoS 2 , and
the static
Ca–MoS refractive index
2 systems values
all had band of gaps
the X–MoS 2 systems
that straddled thewere all potential
redox greater than that ofHowever,
of water. the
purethe system.
CBM of However,
Ca–MoS2with the increase
was found in closein incidence
proximity photon
to the H2 /H + reduction
energy, the refractive
potential, ren-
indices of all
dering its systems
reductiontended
capacity toweaker
overlapthanandthat
approach unity, which indicates
of the Be/Mg/Ca–MoS 2 systems. that the the
Hence,
Be/Mg/Ca–MoS
monolayer MoS2 remained
2 systems possessed
transparent suitable
both before band
andedge
after positions,
doping. making them promising
candidates for high-performance photocatalysts. In comparison, Sr/Ba–MoS2 systems
2.6. Band
wereAlignment
not suitable for water splitting. However, the Sr/Ba–MoS2 system had higher work
functions and smaller
To understand band gaps,
the effects which demonstrates
of alkaline that Sr/Ba–MoS
earth metal doping 2 could potentially
on the photocatalytic
eliminate
capacity of thethe Schotk barrier
monolayer MoS2, height
the bandof alignments
MoS2 -basedoftransistors and reduceMoS
both the monolayer the 2contact
and X–resis-
Molecules 2023, 28, 6122 11 of 15

Molecules 2023, 28, x FOR PEER REVIEW 11 of 16

tance. Schottky contact is one of the bottlenecks currently limiting the application of TMDs
(Transition
MoS2 systemsMetal Dichalcogenides)
were calculated, andinare
semiconductors
illustrated in[58–60]. The
Figure 8. Wesignificant
used thereduction in
concept of
the conduction band of the Sr/Ba–MoS2 system means that metal–MoS2 doped with Sr/Ba
semiconductor electronegativity [52] to determine the redox and oxidation potentials for
has the capacity to reduce the Schottky barrier, which thus establishes a new method for
each water system. The position of the band edge was calculated using the Milligan
the modulation of Schottky barrier height [59]. In the future, we will study Sr/Ba–MoS2
electronegativity equation, which is illustrated in [53–55].
contact.

EAVS
-4.00
CB edge positions

-4.50 H2/H+

-5.00
MoS2 Be-MoS2 Mg-MoS2 Ca-MoS2 Sr-MoS Ba-MoS
2 2

-5.50

H2O/O2
-6.00
Molecules 2023, 28, x FOR PEER REVIEW (a) (b) (c) (d) 12 of 16
(e)
(f)
-6.50
Figure
Schottky 8. Band edge alignments
barrier, of MoS systems before and after doping (redox potentials of
Figure 8. Band edgewhich thus establishes
alignments of MoS22 systemsa new method
before and for
afterthe modulation
doping of Schottky
(redox potentials of
H /H + and H O/O at pH=0 were used as standards). (a) monolayer MoS ; (b) Be–MoS ; (c) Mg–
Hbarrier
/H andheight
22 + 2 [59].
H2O/O In thewere
2 at2 pH=0 future,
used we will study(a)Sr/Ba–MoS
as standards). contact.
monolayer2 MoS 2;2 (b) Be–MoS2;2 (c) Mg–
MoS22; ;(d)
MoS (d)Ca–MoS
Ca–MoS2;2 ;(e)
(e)Sr–MoS
Sr–MoS2;2(f)
; (f)Ba–MoS
Ba–MoS2.2 .
3. Experimental
3. Experimental Section Section
ItItisisworth
worth notingnoting thatthat the𝐸crystal
the crystal = χstructure
structure− 𝐸of MoS + 0.5𝐸
ofcan MoS 2 can adopt various phases,
adopt various phases, including (3)
2
including 1T, 2H, and 3R phases, depending
1T, 2H, and 3R phases, depending upon its arrangement. In this paper, upon its arrangement. In this
we paper,
have optedwe havefor
opted
the stable for2H-MoS
the sTable[61] 2H-MoS
with 2 [61] with space group P63/mc [62] as the basis to construct a
space 𝐸
group =𝐸
P63/mc −𝐸
[62] as the basis to construct a monolayer
2 (4)
monolayer
MoS MoS2 supercell of 4 × 4 × 1. This supercell contained a total of 48 atoms,
2 supercell of 4 × 4 × 1. This supercell contained a total of 48 atoms, including 16 Mo
including
atoms In andthis16 32Mo
equation, atoms
S atoms. and
In 32 S atoms.
χ represents
Figure the
9, In Figure
geometric
the monolayer 9,MoS
mean theofmonolayer
2 the Milligan
is shown MoS is shown
electronegativity
with 2an alkaline with an
of
earth
alkaline
all atoms, 𝐸
earth metal atom
represents X
the (X = Be/Mg/Ca/Sr/Ba)
free electron energy
metal atom X (X = Be/Mg/Ca/Sr/Ba) replacing one Mo atom. The supercell system’s replacing
of 4.5 eV one
on Mo
the atom.
standard The supercell
hydrogen
system’s scale,
electrode
geometry geometry
was and 𝐸 represents
was
optimized optimized
both pre- both
the and pre- and post-doping,
semiconductor
post-doping, bandwith gap with properties
[56]. This
properties method
calculatedcalculated
did not
using
using single-point
provide
single-point exactenergyvalues,energy
after after optimization.
but optimization.
enabled a general The calculations
The evaluation
calculations ofinvolved involved
each system’s the following
position
the following on the
valence
valencehydrogen
normal
electronic electronic
arrangements: arrangements:
electrode Moscale. 1Mo
4d5 5sThe 3s4d 55s
2 3p
, S band 4 ,1,Be
edges S 2s
3sof
22,3p 4, Be2 2s26, Mg 3s
monolayer
Mg 3s 2p , MoSCa 4s 22p
2 3p
2 were
66,,SrCa5s4s 23p
2 4p
suitable 6, 10
6for
3d aSr,
5s 4p
2
Ba3d6
semiconductor
and 2
6s ,5p
10 and6 4d 10
Ba. 6s 5p 4d . and this is consistent with experimental results [57]. Yet the
2
photocatalyst, 6 10

oxidation and reduction potential of Sr/Ba–MoS2 systems were too low to generate a
sufficient driving force for O2 and H2 production compared to standard water-splitting
redox potential. It was suggested that the Sr/Ba–MoS2 systems needed to elevate their
CBM and VBM positions relative to the water splitting potential. The Be–MoS2, Mg–MoS2,
and Ca–MoS2 systems all had band gaps that straddled the redox potential of water.
However, the CBM of Ca–MoS2 was found in close proximity to the H2/H+ reduction
potential, rendering its reduction capacity weaker than that of the Be/Mg/Ca–MoS2
systems. Hence, the Be/Mg/Ca–MoS2 systems possessed suitable band edge positions,
making them promising candidates for high-performance photocatalysts. In comparison,
Figure9.9.MoS
Figure
Sr/Ba–MoS MoS supercellwere
22 supercell
2 systems model
model notboth
both pre-and
pre-
suitable andforpost-doping:
post-doping: (a)top
(a)
water splitting. topview
view ofmonolayer
of
However,monolayer MoS22;;(b)
MoS
the Sr/Ba–MoS (b)top
top2
view
view ofof X-MoS
X-MoS
system had higher 2 2;; (c)
(c) side
side view
view of
of monolayer
monolayer
work functions and smaller MoS
MoS 2 X-MoS
2 and X-MoS 2..
band2 gaps, which demonstrates that
Sr/Ba–MoS2 could potentially eliminate the Schotk barrier height of MoS2-based
This
Thisstudy study employed
employed the the
CASTEPCASTEP software package [63,64] and utilized the projected
transistors and reduce the contact resistance.software
Schottkypackage contact is [63,64]
one ofand utilized
the bottlenecks the
augmented
projected wave (PAW)
augmented wavemethod
(PAW) within
method a density
within a functional
density theory framework
functional theory to con-
framework
currently limiting the application of TMDs (Transition Metal Dichalcogenides) in
duct computational
to conduct computational simulations. The PAW Themethod was employed to depict the impact
semiconductors [58–60]. Thesimulations.
significant reduction PAW in method was employed
the conduction band oftothe depict
Sr/Ba– the
impact
MoS of the ionic reality on the valence electrons, while the exchange–correlation
2 system means that metal–MoS2 doped with Sr/Ba has the capacity to reduce the
interaction was characterized by employing the Perdew–Burke–Ernzerhof (PBE) [65] test
within the generalized gradient approximation (GGA). In this study, we utilized the
semi-empirical Tkatchenko–Scheffler (TS) dispersion correction method to fine-tune van
Molecules 2023, 28, 6122 12 of 15

of the ionic reality on the valence electrons, while the exchange–correlation interaction
was characterized by employing the Perdew–Burke–Ernzerhof (PBE) [65] test within the
generalized gradient approximation (GGA). In this study, we utilized the semi-empirical
Tkatchenko–Scheffler (TS) dispersion correction method to fine-tune van der Waals forces.
To mitigate interlayer coupling effects, we introduced a vacuum layer with a thickness
of 20 Å in the c-direction. We accorded convergence tests and employed the BFGS algo-
rithm for both geometry optimization and electronic calculation; the cutoff energy for the
plane–wave basis was set as 450 eV [65], according to the Monkhorst–Pack scheme [31],
and the k grid was selected as 5 × 5 × 1 (detailed information of convergence tests is
reported in Figure S1 of Supplementary Materials). The self-consistent precision SCF was
set as 2.0 × 10−6 eV/atom, and the maximum stress was set as 0.1 GPa. The maximum
displacement was less than 2 × 10−3 Å and the convergence accuracy of the interatomic
force field was 0.05 eV/Å.

4. Conclusions
The electronic structures and optical properties of the MoS2 systems were studied.
The authors employed first principles based on density function theory to investigate the
effects of doping alkaline earth metals into monolayer MoS2 . Following analyses of the
results regarding AIMD and bond population, it was found that all impurity systems
exhibited thermodynamic stability. The formation energy’s numerical value also suggests
the potential for formation in all impurity systems, with Mg–MoS2 exhibiting the highest
propensity for formation. The ∆γ values indicate that all impurity systems will show
different degrees of lattice distortion, resulting in a reduced band gap width compared
to pure systems. Ba–MoS2 showed the greatest degree of distortion, so its band gap
width was also the narrowest. The introduction of impurities leads to the emergence of
impurity levels in close proximity to the Fermi level, which are clearly visible in the band
structures of all impurity systems, thereby facilitating electron transitions and promoting
the generation of photo-generated electron–hole pairs. In the analysis of the work function,
it was observed that the impurity systems exhibited lower values compared to pure systems,
indicating that the introduction of impurities enhances electron transfer from the interior
of the system to its surface, facilitating their participation in photocatalytic reactions. The
static dielectric function values of impure systems are higher than those of pure systems,
indicating that the introduction of impurities generates a polarizing electric field and
impedes the recombination of photo-generated electron–hole pairs. According to the
analysis of the absorption spectra, it is evident that all impurity systems exhibit a red
shift in comparison to the pure system, indicating an expanded range of response to
solar light upon the introduction of impurities. This expansion has proven advantageous
for enhancing photocatalytic capabilities. From the band alignment analysis, we can
conclude that Be/Mg/Ca–MoS2 systems possess suitable band edge positions, making
them promising candidates for use as high-performance photocatalysts. In comparison,
the Sr/Ba–MoS2 systems were not suitable for water splitting. However, the Sr/Ba–MoS2
system showed higher work functions and smaller band gaps, which demonstrates that
Sr/Ba–MoS2 could potentially reduce the height of the Schotk barrier in MoS2 -based
transistors, and reduce the contact resistance. Sr/Ba–MoS2 contact is a new approach to the
modulation of Schottky barrier height. In the future, we will study Sr/Ba–MoS2 contact.

Supplementary Materials: The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/molecules28166122/s1, Figure S1: Convergence tests of MoS2 .
Author Contributions: Conceptualization, S.-X.W., L.-L.Z., X.-C.Z., B.-C.L. and Y.-N.H.; Methodology,
L.-L.Z. and X.-C.Z.; Software, X.-C.Z.; Validation, Y.-N.H.; Formal analysis, B.-C.L. and H.-M.Y.;
Investigation, L.-Z.L., S.-X.W., X.-C.Z. and B.-C.L.; Resources, X.-C.Z.; Data curation, L.-Z.L., X.-
S.Y. and B.-C.L.; Writing—original draft, L.-Z.L., X.-S.Y. and S.-X.W.; Writing—review & editing,
L.-Z.L. and L.-L.Z.; Visualization, X.-S.Y.; Supervision, L.-L.Z., X.-C.Z., H.-M.Y. and Y.-N.H.; Project
Molecules 2023, 28, 6122 13 of 15

administration, L.-L.Z.; Funding acquisition, L.-L.Z. and H.-M.Y. All authors have read and agreed to
the published version of the manuscript.
Funding: This work was supported by Xinjiang Uygur Autonomous Region Key laboratory opening
project (No. 2020D04011), and Science and Technology Project of Yili Kazak Autonomous Prefecture
(No. YZ2022B021).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data preesented in this syudy are available on request from the
corresponding authors.
Conflicts of Interest: The authors declare no conflict of interest.
Sample Availability: Not applicable.

References
1. Xia, Y.; Cheng, B.; Fan, J.J.; Yu, J.G.; Liu, G. Near-infrared absorbing 2D/3D ZnIn2 S4 /N-doped graphene photocatalyst for highly
efficient CO2 capture and photocatalytic reduction. Sci. China Mater. 2020, 63, 552–565. [CrossRef]
2. Li, X.; Zhang, J.; Zhang, S.J.; Xu, S.S.; Wu, X.G.; Chang, J.C.; He, Z.L. Hexagonal boron nitride composite photocatalysts for
hydrogen production. J. Alloys Compd. 2021, 864, 158153. [CrossRef]
3. Li, M.R.; Zheng, K.T.; Zhang, J.J.; Li, X.M.; Xu, C.J. Design and construction of 2D/2D sheet-on-sheet transition metal sul-
fide/phosphide heterostructure for efficient oxygen evolution reaction. Appl. Surf. Sci. 2021, 565, 150510. [CrossRef]
4. Ao, K.L.; Shao, Y.F.; Chan, I.N.; Shi, X.Q.; Kawazoe, Y.; Yang, M.; Ng, K.W.; Pan, H. Design of novel pentagonal 2d transitional-
metal sulphide monolayers for hydrogen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 16201–16209. [CrossRef]
5. Dongxu, W.; Juan, C.; Xin, G.; Yanhui, A.; Peifang, W. Maximizing the utilization of photo-generated electrons and holes of
g-C3 N4 photocatalyst for harmful algae inactivation. Chem. Eng. J. 2022, 431, 134105.
6. Liu, D.Q.; Barbar, A.; Najam, T.; Javed, M.S.; Shen, J.; Tsiakaras, P.; Cai, X.K. Single noble metal atoms doped 2D materials for
catalysis. Appl. Catal. B Environ. 2021, 297, 120389. [CrossRef]
7. Chang, C.; Chen, W.; Chen, Y.; Chen, Y.; Chen, Y.; Ding, F.; Fan, C.; Fan, H.; Fan, Z.; Gong, C.; et al. Recent progress on
two-dimensional materials. Acta Phys. Chem. Sin. 2021, 37, 2108017. [CrossRef]
8. Liu, C.; Kong, C.; Zhang, F.J.; Kai, C.M.; Cai, W.Q.; Sun, X.Y.; Oh, W.C. Research progress of defective MoS2 for photocatalytic
hydrogen evolution. J. Korean Ceram. Soc. 2021, 58, 135–147. [CrossRef]
9. Xia, Z.H.; Tao, Y.Q.; Pan, Z.G.; Shen, X.D. Enhanced photocatalytic performance and stability of 1T MoS2 transformed from 2H
MoS2 via li intercalation. Results Phys. 2019, 12, 2218–2224. [CrossRef]
10. Zhao, Y.; Tu, J.; Zhang, Y.; Hu, X.; Xu, Y.; He, L. Improved photocatalytic property of monolayer MoS2 by B and F co-doping: First
principles study. J. Catal. 2020, 382, 280–285. [CrossRef]
11. Le, O.K.; Chihaia, V.; Pham-Ho, M.; Son, D.N. Electronic and optical properties of monolayer MoS2 under the influence of
polyethyleneimine adsorption and pressure. RSC Adv. 2020, 10, 4201–4210. [CrossRef] [PubMed]
12. Du, H.; Zhang, Q.; Zhao, B.; Marken, F.; Gao, Q.; Lu, H.; Guan, L.; Wang, H.; Shao, G.; Xu, H.; et al. Novel hierarchical structure
of MoS2 /TiO2 /Ti3 C2 Tx composites for dramatically enhanced electromagnetic absorbing properties. J. Adv. Ceram. 2021, 10,
1042–1051. [CrossRef]
13. Kuc, A.; Zibouche, N.; Heine, T. Influence of quantum confinement on the electronic structure of the transition metal sulfide
TMS2 . Phys. Rev. B 2011, 83, 245213. [CrossRef]
14. Yoffe, A.D. Low-dimensional systems: Quantum size effects and electronic properties of semiconductor microcrystallites
(zero-dimensional systems) and some quasi-two-dimensional systems. Adv. Phys. 2010, 51, 799–890. [CrossRef]
15. Zhang, Z.; Chen, K.; Zhao, Q.; Huang, M.; Ouyang, X. Electrocatalytic and photocatalytic performance of noble metal doped
monolayer MoS2 in the hydrogen evolution reaction: A first principles study. Nano Mater. Sci. 2021, 3, 89–94. [CrossRef]
16. Loh, L.; Zhang, Z.; Bosman, M.; Eda, G. Substitutional doping in 2D transition metal dichalcogenides. Nano Res. 2020, 14,
1668–1681. [CrossRef]
17. Yeonwoong, J.; Yu, Z.; Cha, J.J. Intercalation in two-dimensional transition metal chalcogenides. Inorg. Chem. Front. 2016, 3,
452–463.
18. Zhang, K.; Bersch, B.M.; Joshi, J.; Addou, R.; Cormier, C.R.; Zhang, C.; Ke, X.; Briggs, N.C.; Ke, W.; Subramanian, S. Tuning the
electronic and photonic properties of monolayer MoS2 via in situ rhenium substitutional doping. Adv. Funct. Mater. 2018, 28,
1706950. [CrossRef]
19. Xu, Z.; Luo, B.; Chen, M.; Xie, W.; Xiao, X. Enhanced photogalvanic effects in the two-dimensional WTe2 monolayer by vacancy-
and substitution-doping. Appl. Surf. Sci. 2021, 548, 148751. [CrossRef]
20. Cheng, Y.C.; Zhu, Z.Y.; Mi, W.B.; Guo, Z.B.; Schwingenschlögl, U. Prediction of two-dimensional diluted magnetic semiconductors:
Doped monolayer MoS2 systems. Phys. Rev. B 2013, 87, 1214–1222. [CrossRef]
Molecules 2023, 28, 6122 14 of 15

21. Zhou, J.; Lin, J.; Sims, H.; Jiang, C.; Cong, C.; Brehm, J.A.; Zhang, Z.; Niu, L.; Chen, Y.; Zhou, Y. Synthesis of co-doped MoS2
monolayers with enhanced valley splitting. Adv. Mater. 2020, 32, 1906536. [CrossRef] [PubMed]
22. Burman, D.; Raha, H.; Manna, B.; Pramanik, P.; Guha, P. Substitutional doping of MoS2 for superior gas-sensing applications: A
proof of concept. ACS Sens. 2021, 6, 3398–3408. [CrossRef] [PubMed]
23. Suh, J.; Park, T.E.; Lin, D.Y.; Fu, D.; Park, J.; Jung, H.J.; Chen, Y.; Ko, C.; Jang, C.; Sun, Y. Doping against the native propensity of
MoS2 : Degenerate hole doping by cation substitution. Nano Lett. 2014, 14, 6976–6982. [CrossRef] [PubMed]
24. Dong, T.; Zhang, X.; Wang, P.; Chen, H.S.; Yang, P. Formation of Ni-doped MoS2 nanosheets on N-doped carbon nanotubes
towards superior hydrogen evolution. Electrochim. Acta 2020, 338, 135885. [CrossRef]
25. Zhang, K.; Jia, Y.; Wang, Z.; Wang, L.; Hu, X. Awakening solar hydrogen evolution of MoS2 in alkalescent electrolyte via doping
of Co. ChemSusChem 2019, 12, 3336–3342.
26. Sanakousar, F.M.; Vidyasagar, C.C.; Jiménez-Pérez, V.M.; Prakash, K. Recent progress on visible-light-driven metal and non-metal
doped ZnO nanostructures for photocatalytic degradation of organic pollutants. Mat. Sci. Semicon. Proc. 2022, 140, 106390.
[CrossRef]
27. Ullah, S.; Hussain, A.; Waqaradil, S.; Muhammad Saqlain, A.; Idrees, A.; Ortwin, L.; Altaf, K. Band-gap tuning of graphene by be
doping and Be, B co-doping: A DFT study. RSC Adv. 2015, 5, 55762–55773. [CrossRef]
28. Tayyab, M.; Hussain, A.; Asif, Q.U.A.; Adil, W. Band-gap tuning of graphene by Mg doping and adsorption of Br and Be on
impurity: A DFT study. Comput. Condens. Matter 2020, 23, e00469. [CrossRef]
29. Kumar, G.M.; Cho, H.D.; Ilanchezhiyan, P.; Siva, C.; Ganesh, V.; Yuldashev, S.; Kumar, A.M.; Kang, T.W. Evidencing enhanced
charge-transfer with superior photocatalytic degradation and photoelectrochemical water splitting in Mg modified few-layered
SnS2 . J. Colloid Interface Sci. 2019, 540, 476–485. [CrossRef]
30. Inamdar, A.N.; Som, N.N.; Pratap, A.; Jha, P.K. Hydrogen evolution and oxygen evolution reactions of pristine and alkali metal
doped SnSe2 monolayer. Int. J. Hydrogen Energy 2020, 45, 18657–18665. [CrossRef]
31. Moss, B.S.A.J. Anisotropic mean-square displacements (msd) in single-crystals of 2H-and 3R-MoS2 . Acta Crystallogr. B 1983, 39,
404–407.
32. Hu, J.S.; Duan, W.; He, H.; Lv, H.; Ma, X. A promising strategy to tune the schottky barrier of a MoS2(1-x) Se2x /graphene
heterostructure by asymmetric Se doping. J. Mater. Chem. C 2019, 7, 7798–7805. [CrossRef]
33. Zhu, J.; Vasilopoulou, M.; Davazoglou, D. Intrinsic defects and h doping in WO3 . Sci. Rep. 2017, 7, 40882. [CrossRef] [PubMed]
34. Yang, F.; Lin, S.; Yang, L.; Liao, J.; Chen, Y.; Wang, C.Z. First-principles investigation of metal-doped cubic BaTiO3 . Mater. Res.
Bull. 2017, 96, 372–378. [CrossRef]
35. Wang, S.; Fan, X.; An, Y. Room-temperature ferromagnetism in alkaline-earth-metal doped alp: First-principle calculations.
Comput. Mater. Sci. 2018, 142, 338–345. [CrossRef]
36. Ruan, L.W.; Zhu, Y.J.; Qiu, L.G.; Lu, Y.X. Mechanical properties of doped g-C3 N4 -a first-principle study. Vacuum 2014, 106, 79–85.
[CrossRef]
37. Wang, Y.R.; Li, S.; Yi, J.B. Transition metal-doped tin monoxide monolayer: A first-principles study. J. Phys. Chem. C 2018, 122,
4651–4661. [CrossRef]
38. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868.
[CrossRef]
39. Mo, Y.; Peyerimhoff, S.D. Theoretical analysis of electronic delocalization. J. Chem. Phys. 1998, 109, 1687–1697. [CrossRef]
40. Li, G.; Chen, M.Q.; Zhao, S.X.; Li, P.W.; Hu, J.; Sang, S.B.; Hou, J.J. Effect of Se doping on the electronic band structure and optical
absorption properties of single layer MoS2 . Acta Phys.-Chim. Sin. 2016, 32, 2905–2912. [CrossRef]
41. Pan, F.C.; Lin, X.L.; Cao, Z.J.; Li, X.F. Electronic structures and optical properties of Fe, Co, and Ni doped GaSb. Acta Phys. Sin.
2019, 68, 184202. [CrossRef]
42. Ma, J.; Shang, S.L.; Kim, H.; Liu, Z.K. An ab initio molecular dynamics exploration of associates in Ba-Bi liquid with strong
ordering trends. Acta Mater. 2020, 190, 81–92. [CrossRef]
43. Guo, X.; Gu, J.; Lin, S.; Zhang, S.; Huang, S. Tackling the activity and selectivity challenges of electrocatalysts towards nitrogen
reduction reaction via atomically dispersed Bi-atom catalysts. J. Am. Chem. Soc. 2020, 124, 5709–5712. [CrossRef]
44. Lv, L.; Shen, Y.; Liu, J.; Meng, X.; Gao, X.; Zhou, M.; Zhang, Y.; Gong, D.; Zheng, Y.; Zhou, Z. Computational screening of high
activity and selectivity Tm/g-G3 N4 single-atom catalysts for electrocatalytic reduction of nitrates to ammonia. J. Phys. Chem. Lett.
2021, 12, 11143–11150. [CrossRef]
45. Kam, K.K.; Parkinson, B.A. Detailed photocurrent spectroscopy of the semiconducting group vib transition metal dichalcogenides.
J. Phys. Chem. 1982, 86, 463–467. [CrossRef]
46. Lebgue, S.; Eriksson, O. Electronic structure of two-dimensional crystals from ab initio theory. Phys. Rev. B 2009, 79, 115409.
[CrossRef]
47. Pereira, F.; Xiao, K.; Latino, D.A.; Wu, C.; Zhang, Q.; Aires-De-Sousa, J. Machine learning methods to predict density functional
theory B3LYP energies of HOMO and LUMO orbitals. J. Chem. Inf. Model. 2017, 57, 11–21. [CrossRef]
48. Obeid, M.M.; Stampfl, C.; Bafekry, A.; Guan, Z.; Jappor, H.R.; Nguyen, C.V.; Naseri, M.; Hoat, D.M.; Hieu, N.N.; Krauklis, A.E.
First-principles investigation of nonmetal doped single-layer BiOBr as a potential photocatalyst with a low recombination rate.
Phys. Chem. Chem. Phys. 2020, 22, 15354–15364. [CrossRef] [PubMed]
Molecules 2023, 28, 6122 15 of 15

49. Wei, Y.; Ma, X.G.; Zhu, L.; He, H.; Huang, C.Y. Interfacial cohesive interaction and band modulation of two-dimensional
MoS2 /graphene heterostructure. Acta Phys. Sin. 2017, 66, 87101.
50. Zhang, L.L.; Xia, T.; Liu, G.A.; Lei, B.C.; Zhao, X.C.; Wang, S.X.; Huang, Y.N. Electronic and optical properties of N-Pr co-doped
anatase TiO2 from first-principles. Acta Phys. Sin.-Cn. Ed. 2019, 68, 17401. [CrossRef]
51. Lang, Q.Z.; Huang, Y.B.; Wei, J.M.; Wang, Y.; Ding, Z. The magnetic, optical and electronic properties of Mn-X(X = O, Se, Te, Po)
co-doped MoS2 monolayers via first principle calculation. Mater. Res. Express 2020, 7, 116301. [CrossRef]
52. Liping, Z.; Mietek, J. Toward designing semiconductor-semiconductor heterojunctions for photocatalytic applications. Appl. Surf.
Sci. 2018, 430, 2–17.
53. Ma, L.; Li, M.N.; Zhang, L.L. Electronic and optical properties of GaN/MoSe2 and its vacancy heterojunctions studied by
first-principles. J. Appl. Phys. 2023, 133, 45307. [CrossRef]
54. Pang, Y.; Feng, Q.; Kou, Z.; Xu, G.; Gao, F.; Wang, B.; Pan, Z.; Lv, J.; Zhang, Y.; Wu, Y. A surface precleaning strategy intensifies the
interface coupling of the Bi2 O3 /TiO2 heterostructure for enhanced photoelectrochemical detection properties. Mat. Chem. Front.
2020, 4, 638–644. [CrossRef]
55. Zhang, Z.; Guo, Y.; Robertson, J. Chemical bonding and band alignment at X2 O3 /GaN (X = Al, Sc) interfaces. Appl. Phys. Lett.
2019, 114, 161601. [CrossRef]
56. Fawadkhan, I.M.; Nguyen, C.; Iftikharahmad, B. A first-principles study of electronic structure and photocatalytic performance of
GaN-MX2 (M = Mo, W. X = S, Se) van der Waals heterostructures. RSC Adv. 2020, 10, 24683–24690.
57. Jia, G.; Liu, Y.; Gong, J.Y.; Lei, D.; Wang, D.L.; Huang, Z.X. Excitonic quantum confinement modified optical conductivity of
monolayer and few-layered MoS2 . J. Mater. Chem. C 2016, 37, 4. [CrossRef]
58. Wang, X.; Kim, S.Y.; Wallace, R.M. Interface chemistry and band alignment study of Ni and Ag contacts on MoS2 . ACS Appl.
Mater. Interfaces 2021, 13, 15802–15810. [CrossRef]
59. McDonell, S.; Azcatl, A.; Addou, R.; Gong, C.; Battaglia, C.; Chuang, S.; Cho, K.; Javey, A.; Wallace, R.M. Hole contacts on
transition metal dichalcogenides: Interface chemistry and band alignments. ACS Nano 2014, 8, 6265–6272. [CrossRef]
60. Gong, C.; Huang, C.; Miller, J.; Cheng, L.; Hao, Y.; Cobden, D.; Kim, J.; Ruoff, R.S.; Wallace, R.M.; Cho, K. Metal contacts on
physical vapor deposited monolayer MoS2 . ACS Nano 2013, 7, 11350–11357. [CrossRef]
61. Li, S.S.; Sun, J.R.; Guan, J.Q. Strategies to improve electrocatalytic and photocatalytic performance of two-dimensional materials
for hydrogen evolution reaction. Chin. J. Catal. 2021, 42, 511–556. [CrossRef]
62. Lin, L.; Guo, Y.; He, C.; Tao, H.; Huang, J.; Yu, W.; Chen, R.; Lou, M.; Yan, L. First-principles study of magnetism of 3d transition
metals and nitrogen co-doped monolayer MoS2 . Chin. Phys. B 2020, 29, 517–524. [CrossRef]
63. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas,
illustrations and the castep code. J. Phys. Condens. Mat. 2002, 14, 2717–2744. [CrossRef]
64. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using castep:
Zeitschrift für kristallographie-crystalline materials. Z. Krist. 2005, 220, 567–570.
65. Sun, J.; Leng, J.; Zhang, G. Theoretically exploring the possible configurations, the electronic and transport properties of MoS2
-OH bilayer. Phys. Lett. A 2020, 384, 126575. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like