You are on page 1of 9

This is an open access article published under a Creative Commons Non-Commercial No

Derivative Works (CC-BY-NC-ND) Attribution License, which permits copying and


redistribution of the article, and creation of adaptations, all for non-commercial purposes.

http://pubs.acs.org/journal/acsodf Article

Computational Study of Janus Transition Metal Dichalcogenide


Monolayers for Acetone Gas Sensing
Chen-Hao Yeh*
Cite This: ACS Omega 2020, 5, 31398−31406 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Recently, Janus two-dimensional (2D) transition metal


Downloaded via INDIAN INST OF TECH KHARAGPUR on July 1, 2022 at 09:06:36 (UTC).

dichalcogenides (TMDs) have been widely investigated and have


provided exciting prospects in many fields such as photoelectric
materials, photocatalysis, and gas sensors. In this study, we performed
density functional theory (DFT) calculations to study the sensitivity of
four volatile organic compounds (VOCs), including acetone,
methanol, ethanol, and formyl aldehyde, over pristine 2D TMDs
and 2D Janus TMD monolayers. We found that MoS2, Janus MoSSe,
and Janus MoSTe demonstrated greater sensitivity toward acetone
than other VOCs. Furthermore, the band gap values of the Janus
MoSSe and Janus MoSTe monolayers dramatically changed after
acetone adsorption on their sulfur layers, which was quite larger than
the band gap change after acetone adsorption on the MoS2 monolayer.
This result also leads to the extremely large conductivity change of Janus MoSSe and Janus MoSTe after sensing acetone. Hence,
Janus MoSSe and Janus MoSTe monolayers show much higher sensitivity toward acetone in comparison with the pristine MoS2
monolayer. Finally, our finding indicates that Janus MoSSe and Janus MoSTe monolayers can be proposed as ultrahigh-sensitivity
2D TMD materials for acetone sensors.

1. INTRODUCTION chemical vapor deposition (CVD) method by Li and co-


Since the discovery of graphene in 2004 by Andre Geim and workers16 and Lou and co-workers17 even though Janus TMDs
Konstantin Novoselov,1 two-dimensional (2D) materials have do not exist in nature. The 2D Janus TMD (denoted as MXY)
been gradually attracting the attention of scientists due to their possesses a symmetry-broken structure with a C3v point group
many unique properties that differ from those of 3D bulk compared with the single-layer MX2 TMD structure, which has a
materials.2−9 To date, there are many kinds of 2D materials, such higher symmetry of the D3h point group.18,19 Theoretically,
as carbon materials (graphene and its derivatives), h-BN, group-VI chalcogenide MXY monolayers, including MoSSe,
transition metal dichalcogenides (TMDs, e.g. MoS2), and WSSe, WSeTe, and WSTe monolayers, have been demonstrated
MXene (such as Ti3C2F2). Among these 2D materials, transition to be stable by calculating the phonon dispersion and the
metal dichalcogenides are one of the most widely studied 2D molecular dynamics simulations.20,21 Using many-body Green’s
materials due to their specific electrical, optical, and other function perturbation theory, Li et al. found that the Janus
physical properties,10−14 and MoS2 is the most well understood MoSSe monolayer showed strong excitonic effects, which can be
TMD material. applied in optoelectronic materials.22
Recently, a new class of 2D materials, named Janus 2D With the development of industries, emissions of toxic gas and
materials, has been gradually attracting considerable interest air pollutants have rapidly increased globally, including carbon
because Janus 2D materials have distinct properties different monoxide (CO), nitrogen oxides (NOx), ozone (O3),
from those of traditional 2D materials. Janus 2D materials are particulates (PM2.5), asbestos, and volatile organic compounds
those materials that have asymmetric structures possessing two (VOCs).23,24 However, in recent years, volatile organic
different facets on two sides of 2D materials. The first reported compound (VOC) emissions have also increased significantly.
Janus 2D material named graphone, which was simulated by
Zhou et al.,15 is a graphene-based material obtained by
hydrogenating only one side of graphene. Graphone has a Received: October 9, 2020
small band gap of 0.46 eV, and it is a ferromagnetic Accepted: November 16, 2020
semiconductor. In addition, there are also many experimental Published: November 25, 2020
and theoretical studies on Janus TMD materials. In the
experimental studies, a single-layer Janus TMD, MoSSe, was
successfully fabricated in the laboratory in 2017 by a modified

© 2020 American Chemical Society https://dx.doi.org/10.1021/acsomega.0c04938


31398 ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 1. Top and side views of the optimized pristine and Janus TMD monolayers: (a) MoS2, (b) MoSe2, (c) MoTe2, (d) MoSSe, (e) MoSTe, and (f)
MoSeTe. Purple, yellow, green, and tawny spheres represent Mo, S, Se, and Te atoms, respectively.

Figure 2. Top and side views of the optimized adsorption structures of (a) methanol, (b) formyl aldehyde, (c) ethanol, and (d) acetone on the pristine
MoS2 monolayer. Purple and yellow spheres represent Mo and S atoms, respectively, while brown, red, and light pink balls are C, O, and H atoms,
respectively.

Pollutants are released primarily from chemical processes (such first-principle calculations.29−31 By means of either the defected
as paints) and have serious adverse effects on human health, structures or the strain deformation of these Janus TMD
such as cancer, central nervous system disorders, and skin monolayers, it has been demonstrated that the Janus TMD
problems. In the literature, 2D materials, including graphene monolayers possess high sensitivity to NO2 and NH3. However,
oxides, phosphorene, and TMDs, have recently been widely
despite MoS2 being reported as a good VOC sensor,32 the
investigated for the applications of toxic gas sensing.25,26
sensitivity of Janus TMD monolayers with respect to VOCs is
Besides, to reduce high operating temperatures, which would
increase the power consumption, recent research of gas sensing still unclear. Thus, in this work, we studied the adsorption and
has been devoted to the fabrication of room-temperature sensors sensitivity of four different VOC molecules, including acetone,
toward VOCs and toxic gases. Among them, TMDs have been methanol, ethanol, and formyl aldehyde, using molybdenum-
reported to show a good VOC sensing ability at room based pristine TMDs and Janus TMDs. Except for the
temperature. This operating temperature could be used for a adsorption energy, we calculate and present a discussion of
low-temperature VOC sensor, being lower than that of the electronic analyses, including density derived electrostatic
traditional VOC sensors. For TMDs, it has been reported that and chemical (DDEC) atomic charges and electron density
the nonfunctionalized MoS2 can detect toluene and hexane27 difference (EDD), to interpret the interaction between the
while the Au-decorated MoS2 shows sensitivity to ethanol and
VOCs and the TMDs. Moreover, to determine the VOC sensing
acetone.28 These studies reveal that the functionalization and
modification of TMDs can efficiently tune their sensing capability, we calculated the band structure and compared the
properties toward gas molecules. conductivity change difference of VOCs on the Janus TMDs
Very recently, Janus MoSSe and WSSe monolayers have been with that on the pristine TMDs via the calculated band gap
studied in the sensing of NO2 and NH3 gaseous molecules using values.
31399 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

2. RESULTS AND DISCUSSION Table 1. Calculated Adsorption Energy, Eads, the Band Gap
Figure 1 shows the (4 × 4) supercell of all TMDs with a 2H Value Change, ΔEg (in the Fourth Decimal Place), and the
phase. The optimized unit cell lattice constants of the MoS2, Conductivity Change, Δσ, after Gas Molecule Adsorption on
Pristine TMD Monolayers
MoSe2, and MoTe2 monolayers are 3.18, 3.31, and 3.55 Å,
respectively. These values are in good agreement with the TMD species Eads (eV) ΔEg (eV) Δσ (%)
experimental values.33,34 Besides, the optimized lattice constants CH3OH −0.26 ∼0 ∼0
of the Janus MoSSe, MoSTe, and MoSeTe monolayers are 3.25, MoS2 C2H5OH −0.25 −0.0003 0.586
3.36, and 3.43 Å, respectively. By considering the spin−orbit CH2O −0.25 −0.0003 0.586
coupling (SOC) interaction, the calculated fundamental band C3H6O −0.35 −0.0155 35.230
gaps of the MoS2, MoSe2, and MoTe2 monolayers are 1.5854, CH3OH −0.25 −0.0074 15.499
1.3241, and 0.9324 eV, respectively, while the calculated MoSe2 C2H5OH −0.29 −0.0008 1.570
fundamental band gaps of Janus MoSSe, MoSTe, and MoSeTe CH2O −0.28 −0.0022 4.377
monolayers are 1.4553, 0.9556, and 1.1367 eV, respectively. The C3H6O −0.33 −0.0002 0.390
calculated electronic band structure diagrams are depicted in CH3OH −0.25 −0.0001 0.195
Figure S1. Except for the Janus MoSTe monolayer, all of the MoTe2 C2H5OH −0.33 −0.0003 0.586
TMDs show a direct band gap. The band gap properties and CH2O −0.25 −0.0005 0.978
values of these TMDs calculated using the PBE + SOC + D3 C3H6O −0.32 −0.0001 0.195
functional possess an error of around 0.10−0.30 eV with respect
to the experimental or calculated values.16,35−39 The calculations gap value change, the MoS2 monolayer demonstrates high
via PBE + SOC + D3 functional show reliable electronic sensitivity with respect to the acetone molecule.
properties compared to the experimental observation. Subsequently, as can be seen in Figures S3 and S4, we
2.1 Adsorption of VOC Molecules on the TMDs. We first calculate the adsorption energies of methanol, ethanol, formyl
calculate the adsorption of methanol, ethanol, formyl aldehyde, aldehyde, and acetone on both MoSe2 and MoTe2. The
and acetone on the MoS2 monolayer. The calculated adsorption adsorption energies of methanol, ethanol, formyl aldehyde,
energies of methanol, ethanol, and formyl aldehyde on the MoS2 and acetone on the MoSe2 monolayer are −0.25, −0.29, −0.28,
monolayer are similar, being −0.26, −0.25, and −0.25 eV, and −0.33 eV, respectively. Moreover, the adsorption energies
respectively. However, the adsorption energy of acetone on the of methanol, ethanol, formyl aldehyde, and acetone on the
MoS2 monolayer is −0.35 eV, larger than those of methanol, MoTe2 monolayer are −0.25, −0.33, −0.25, and −0.32 eV,
ethanol, and formyl aldehyde. Figure 2 shows the optimized respectively. We found that the adsorption energies of these
adsorption structures of these VOCs on the MoS2 monolayer. VOCs on both MoSe2 and MoTe2 are similar to those on the
We found that after adsorption, both methanol and ethanol can MoS2 monolayer, while only the adsorption energy of ethanol is
point to the sulfur atom via their hydroxyl groups, while formyl slightly larger on the MoSe2 and MoTe2 monolayers. However,
aldehyde and acetone can be almost parallelly (having a small as shown in the band structure plots in Figures S5 and S6, the
tilting angle) adsorbed on the MoS2 monolayer. In addition, calculated band gap value changes after the adsorption of these
during adsorption, weak C−H...S interactions between these VOCs are almost the same with respect to the ones before
VOCs and the MoS2 monolayer are formed. The adsorption of adsorption except for methanol on the MoSe2 monolayer, as
these VOC molecules on the MoS2 monolayer is contributed by listed in Table 1. Even though the calculated band gap value and
the C−H...S interaction and the dispersion force so that the conductivity changes of methanol on the MoSe2 monolayer are
adsorption energies of these VOC molecules are around −0.25 larger than those of other VOCs, they are still smaller than those
to −0.36 eV. of the acetone counterpart on the MoS2 monolayer. Therefore,
Besides, for semiconducting sensors, the band gap changes in comparison with the sensing ability of these four VOCs on
and their conductivity change after sensing the target molecule MoS2, MoSe2, and MoTe2 monolayers, MoS2 demonstrates the
are also criteria to identify their sensitivity. We found that the best sensing properties toward the acetone molecule.
2.2. Adsorption of VOC Molecules on the Janus TMDs.
calculated band gap value change after methanol, ethanol, and
We studied the Janus MoSSe, MoSTe, and MoSeTe monolayers
formyl aldehyde were adsorbed on the MoS2 monolayer was
to discuss their sensing properties toward these VOCs, as can be
smaller than 0.001 eV, while only acetone possessed a larger
seen in Table 2. Because the Janus MoXY nanosheet has two
band gap value change of −0.0155 eV. The calculated electronic asymmetric sides, we then calculate the adsorption of these
band structure diagrams are shown in Figure S2. The negative VOCs on both sides of these Janus MoXY nanosheets. On the
value of the band gap change implies an increase of electric Janus MoSSe monolayer, the adsorption energies of methanol,
conductivity and thus a larger band gap change can lead to a ethanol, formyl aldehyde, and acetone on the S-layer of MoSSe
further increase in conductivity. This conductivity change are −0.25, −0.28, −0.24, and −0.31 eV, respectively. On the
reflects the sensitivity of the semiconducting sensor. Hence, other hand, the adsorption energies of methanol, ethanol, formyl
the calculated conductivity change can increase by 35.23% after aldehyde, and acetone on the Se-layer of MoSSe are −0.24,
acetone adsorption on the MoS2 monolayer. On the contrary, −0.23, −0.24, and −0.34 eV, respectively. Figure 3 shows the
the other three VOC molecules can only make a negligible optimized adsorption structures of these VOCs on the Janus
conductivity change after being captured by the MoS 2 MoSSe monolayer. The adsorption structure and adsorption
monolayer, as listed in Table 1. Our results show that the energy of these VOCs on Janus MoSSe are also similar to those
MoS2 monolayer has the largest band gap value change after the on the MoS2 monolayer. However, as can be seen from Table 2
adsorption of acetone. Since acetone also possesses the largest and Figure S7, despite the adsorption energy of acetone on the
adsorption energy on the MoS2 monolayer, in the summa- S-layer of Janus MoSSe being slightly smaller than that on the
rization of these two criteria of the adsorption energy and band MoS2 monolayer, the band gap change after acetone adsorption
31400 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

Table 2. Calculated Adsorption Energy, Eads, the Band Gap changes, as listed in Table 2. These results reveal that the S-layer
Value Change, ΔEg (in the Fourth Decimal Place), and the of Janus MoSSe possesses a significantly larger electric
Conductivity Change, Δσ, after Gas Molecule Adsorption on conductivity change after sensing the acetone molecule in
Janus TMD Monolayers comparison with the other three VOC molecules. Besides, the
TMD species Eads (eV) ΔEg (eV) Δσ (%)
band gap change is demonstrated to be almost negligible after
the adsorption of methanol, ethanol, formyl aldehyde, and
CH3OH −0.25 −0.0007 1.372
MoSSe−S C2H5OH −0.28 −0.0001 0.195
acetone on the Se-layer of Janus MoSSe, being similar to that on
CH2O −0.24 −0.0005 0.978
MoSe2, as listed in Table 2. Thus, the Se-layer of the Janus
C3H6O −0.31 −0.1028 640.125 MoSSe monolayer shows pool sensitivity to the VOCs.
CH3OH −0.24 −0.0009 1.768 On the Janus MoSTe monolayer, the adsorption energies of
MoSSe−Se C2H5OH −0.23 ∼0 ∼0 methanol, ethanol, formyl aldehyde, and acetone on the S-layer
CH2O −0.24 −0.0002 0.390 of the MoSTe are −0.24, −0.32, −0.23, and −0.31 eV,
C3H6O −0.34 −0.0004 0.762 respectively. We found that the adsorption energy of ethanol
CH3OH −0.24 −0.0054 11.087 increases from −0.25 (on MoS2) to −0.32 eV (on MoSTe), but
MoSTe−S C2H5OH −0.32 −0.0184 43.085 the adsorption energy of acetone is similar to that on other
CH2O −0.23 −0.0062 12.831 TMDs. The optimized adsorption geometries of these VOCs on
C3H6O −0.31 +0.1226 −90.811 both S- and Te-layers of the Janus MoSTe monolayer are shown
CH3OH −0.25 −0.0004 0.782 in Figure 4, which are also similar to those on the MoS2
MoSTe−Te C2H5OH −0.24 +0.0002 −0.389 monolayer. In addition, as can be seen from the band structure
CH2O −0.24 −0.0020 3.971 diagrams in Figure S8, the calculated band gap changes after the
C3H6O −0.34 −0.0014 2.763 adsorption of methanol, ethanol, formyl aldehyde, and acetone
CH3OH −0.26 −0.0006 1.175
on the S-layer of MoSTe are −0.005, −0.018, −0.006, and
MoSeTe−Se C2H5OH −0.33 −0.0017 3.366
+0.123 eV, respectively. We found that the band gap change is
CH2O −0.24 −0.0003 0.586
positive for acetone on the S-layer of MoSTe, which indicates
C3H6O −0.33 −0.0005 0.978
CH3OH −0.24 −0.0005 0.978
that the band gap value becomes larger and then the electrical
MoSeTe−Te C2H5OH −0.29 −0.0017 3.366
resistance can also become larger after acetone adsorption.
CH2O −0.26 −0.0008 1.570 Therefore, the calculated conductivity change can reduce by
C3H6O −0.32 −0.0002 0.390 90.81% after acetone sensing by the MoSTe monolayer. This
also demonstrates that the S-layer of MoSTe can enhance the
on the S-side of Janus MoSSe is −0.1028 eV, which is almost five sensitivity of acetone with respect to the MoS2 monolayer. On
times that of MoS2. Thus, the calculated conductivity change can the other hand, the adsorption energies of methanol, ethanol,
increase to a great extent by 640.12% after acetone adsorption formyl aldehyde, and acetone on the Te-layer of MoSTe are
on the S-layer of the Janus MoSSe monolayer. On the other −0.25, −0.24, −0.24, and −0.34 eV, respectively. Likewise, the
hand, the band gap change after the adsorption of methanol, band gap change is also demonstrated to be almost negligible
ethanol, and formyl aldehyde on the S-layer of Janus MoSSe is after the adsorption of methanol, ethanol, formyl aldehyde, and
smaller than 0.001 eV, also leading to negligible conductivity acetone on the Te-layer of Janus MoSTe. This reveals that the

Figure 3. Top and side views of the optimized adsorption structures of (a) methanol, (b) formyl aldehyde, (c) ethanol, and (d) acetone on the S-layer
of the Janus MoSSe monolayer and (e) methanol, (f) formyl aldehyde, (g) ethanol, and (h) acetone on the Se-layer of the Janus MoSSe monolayer.
Purple, yellow, and green spheres represent Mo, S, and Se atoms, respectively, while brown, red, and light pink balls are C, O, and H atoms, respectively.

31401 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 4. Top and side views of the optimized adsorption structures of (a) methanol, (b) formyl aldehyde, (c) ethanol, and (d) acetone on the S-layer
of the Janus MoSTe monolayer and (e) methanol, (f) formyl aldehyde, (g) ethanol, and (h) acetone on the Te-layer of the Janus MoSTe monolayer.
Purple, yellow, and tawny spheres represent Mo, S, and Te atoms, respectively, while brown, red, and light pink balls are C, O, and H atoms,
respectively.

Te-layer of the Janus MoSTe monolayer also shows significant Table 3. Calculated DDEC Atomic Charges for Acetone
pool sensitivity to the VOCs. Molecule after Adsorption on Both Pristine and Janus TMD
Finally, we calculated the sensing properties of these four Monolayersa
VOCs on the Janus MoTeSe monolayer, and their adsorption
TMD species DDEC (|e|)
energies are listed in Table 2, which are also similar to their
adsorption energies on the TMDs mentioned above. Likewise, MoS2 C3H6O 0.036
the adsorption structures of these VOCs on the Janus MoTeSe MoSe2 C3H6O 0.015
monolayer are similar to those on the TMDs mentioned above, MoTe2 C3H6O −0.016
as shown in Figure S9. However, the band gap changes of these MoSSe−S C3H6O 0.028
four VOCs on either Se-layer or Te-layer are almost unchanged MoSSe−Se C3H6O 0.026
after the adsorption of these four VOCs, as can be seen in the MoSTe−S C3H6O −0.032
band structure plots in Figure S10. This phenomenon indicates MoSTe−Te C3H6O 0.006
that the molybdenum-based TMDs containing either Se or Te MoSeTe−Se C3H6O 0.006
elements cannot cause an obvious conductivity change after MoSeTe−Te C3H6O 0.003
a
sensing the target VOC molecules. However, the molybdenum- The positive and negative values indicate the losing and gaining of
based TMDs and their Janus derivatives, which include sulfur electrons of the acetone molecule.
atoms, demonstrate an apparent conductivity change after
sensing the target VOC molecules. These results reveal that the plot of acetone adsorption on the MoS2 monolayer. The EDD
sulfur atom of the TMDs might possess good sensitivity to plot illustrates that there are electron accumulation and
acetone. Among them, our results show that Janus MoSSe can depletion around the acetone’s oxygen atom and H atom of
have the highest sensing ability to acetone due to the largest the C−H bond, respectively. It can also be noticed from the
conductivity change after acetone adsorption on Janus MoSSe. EDD plot that the electron density around the sulfur atoms
2.3. Electronic Property Analysis. Based on the results of displays electron accumulation between the sulfur atoms and the
the band gap change before and after VOC adsorption, we found acetone’s C−H bond and the depletion between the sulfur
that MoS2, Janus MoSSe, and Janus MoSTe possess larger band atoms and the acetone’s oxygen atom. These results reveal that
gap changes in comparison with other Mo-based TMDs. To the oxygen atom of acetone could obtain electrons from MoS2,
understand the electronic property between acetone and MoS2, while its C−H bond would transfer electrons to the MoS2
Janus MoSSe, and Janus MoSTe monolayers, we calculated the monolayer, constructing a two-way electron transfer flow.
net atomic charges by DDEC6 and the electron density According to the calculated net atomic charges, acetone loses
difference (EDD). As can be seen in Table 3, the calculated electrons after its adsorption, meaning that the electron transfer
DDEC atomic charges of acetone are 0.036 |e|, 0.015 |e|, and of the C−H bond of acetone to the MoS2 monolayer is stronger
−0.016 |e| on the MoS2, MoSe2, and MoTe2 monolayers, than that from the MoS2 to the oxygen of acetone.
respectively. These results indicate that the acetone molecule On Janus TMDs, the calculated DDEC atomic charges of
donates electrons to the MoS2 and MoSe2 monolayers while it acetone are 0.028 |e| and 0.026 |e| on the S-layer and Se-layer of
gains electrons from the MoTe2 monolayer. Besides, the the Janus MoSSe monolayer, respectively. Moreover, the
electron transfer magnitude implies that the larger band gap calculated DDEC atomic charge of acetone is −0.032 |e| on
change of acetone on MoS2 might result from a higher electron the S-layer of the Janus MoSTe monolayer, whereas the atomic
transfer magnitude. In addition, Figure 5a shows the 3D EDD charges of acetone are smaller than 0.010 |e| on the Te-layer of
31402 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

monolayer, as can be seen in Figure 5b,c, respectively. The EDD


in Figure 5b shows that a few electrons aggregate around the
carbon atom of the CO group in acetone after the adsorption
of acetone on the S-layer of Janus MoSSe. In comparison with
the EDD plot of acetone on MoS2, acetone gains more electrons
from Janus MoSSe but donates similar electrons to the MoSSe
monolayer. This result reflects why the net electron transfer of
acetone to the MoSSe monolayer is smaller than that of the
MoS2 monolayer.
On the other hand, Figure 5c shows the EDD plot of acetone
on the S-layer of the Janus MoSTe monolayer, where a larger
region of electron density accumulation around the acetone
molecule than that on either MoS2 or Janus MoSSe monolayer is
displayed. This result is also reflected in the net atomic charge
calculation that acetone obtains more electrons from the Janus
MoSTe monolayer. Thus, the electronic property analysis
between acetone and these TMDs reveals that the electron
transfer direction and strength can be altered by using the Janus
TMD structure with respect to the MoS2 monolayer. It also
implies that when acetone obtains more electrons from the
sulfur atoms of the Janus TMDs, the band gap value change
becomes larger.
2.4. Discussion. According to our results, we found that the
energy levels of the valence band maximum (VBM) or
conduction band minimum (CBM) of these TMDs change
only when acetone is adsorbed on the S-layer of MoS2, MoSSe,
and MoSTe. Thus, based on the calculated band structure, it is
demonstrated that only acetone possesses a relatively strong
electronic interaction with the sulfur atoms of MoS2, MoSSe,
and MoSTe. Moreover, the S-layer of the Janus MoSSe and
MoSTe monolayers showed better sensitivity with respect to
pristine MoS2 after acetone adsorption. As many previous
studies have pointed out,40,41 one of the reasons for this is that
the polarization induced by symmetry-breaking in Janus
materials modulates the electron transfer. Polarization of the
Janus materials causes dipole−dipole interactions between the
polar materials or gaseous species.42−44 Hence, the internal
dipole interaction in the Janus materials can couple with the
electronic interaction between acetone and the Janus materials
and then alter the magnitude or direction of the electron
transfer. In addition, based on Ohm’s law, the electric
conductivity is the slope of the I−V curve, and the electron
transport phenomenon before and after gas adsorption has
trends similar to the results of the calculated conductivity
change.45,46 The smaller band gap value can lead to larger
conductivity as well as a more tilted slope and vice versa.
Therefore, the calculated conductivity change based on the band
gap value can also predict the trends of the I−V curve.

3. CONCLUSIONS
Figure 5. Calculated electron density difference (EDD) plots of
acetone adsorption on (a) MoS2, (b) S-layer of Janus MoSSe, and (c) S- In this work, we have employed density functional theory
layer of Janus MoSTe monolayers. Green and red represent the regions (DFT) calculations to study the sensitivity of acetone, methanol,
of electron depletion and accumulation, respectively. The isosurface ethanol, and formyl aldehyde over pristine 2D transition metal
level is 0.0003 |e|/Bohr.3 dichalcogenides (TMDs) and 2D Janus TMD monolayers. On
the MoS2, MoSe2, and MoTe2 monolayers, we found that the
Janus MoSTe and both sides of Janus MoSeTe monolayers, as adsorption energy of acetone is larger than those of its
listed in Table 3. These results also demonstrate that the counterparts methanol, ethanol, and formyl aldehyde. In
numbers of electron transfer are larger on the S-layer of either addition, the adsorption of acetone on the MoS2 monolayer
Janus MoSSe or MoSTe monolayer than on other sides of can lead to a larger band gap change, resulting in a larger
various Janus TMDs. Furthermore, according to the EDD conductivity change with an increase of 35.23%. Based on the
analysis, similar to the electronic interaction between acetone EDD calculation, we notice that the acetone forms a two-way
and MoS2, there also exists a two-way electron transfer between electron transfer direction with the MoS2 monolayer, in which
acetone and the S-layer of either Janus MoSSe or MoSTe one is the electron transfer from MoS2 to the oxygen of the
31403 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

acetone and the other is the electron donation from acetone’s adsorbate VOC molecule, and ETMD is the energy of pristine
C−H bond to MoS2 via the C−H...S interaction. The calculated MoX2 or Janus MoXY monolayers. Therefore, negative
DDEC atomic charges also reveal that there is a stronger adsorption energy indicates a thermodynamically favored
electron exchange interaction between acetone and the MoS2 exothermic adsorption process. To analyze the electrical
monolayer. properties, the density derived electrostatic and chemical
Furthermore, the adsorption energies of four VOCs on the 2D (DDEC6) approach was used to examine electron transfer
Janus TMD monolayers are similar to those on the pristine between the molecule and the surface.61−63 Besides, it is well
TMDs. However, we found that after the adsorption of acetone known that band gap is an important factor for determining the
on the S-layer of both Janus MoSSe and MoSTe monolayers, it electrical conductivity of the semiconductor sensor.64−66 The
displays an extremely large band gap value change in comparison relationship between band gap and electronic conductivity is
with the adsorption of acetone on MoS2. After sensing the understood by the following equation

ji −Eg zyz
σ ∝ expjjj zz
acetone molecule, the band gap value changes of Janus MoSSe
and MoSTe are −0.103 and 0.123 eV, respectively. These results
cause Janus MoSSe and MoSTe to become conductive and k 2κT { (2)
resistive sensors since the electrical conductance and electrical
resistance increase by 640.125 and 90.811% on Janus MoSSe where σ is the electrical conductivity and κ is the Boltzmann
and MoSTe, respectively. On the other hand, the band gap constant. According to this equation, the smaller band gap (Eg)
changes as well as the conductivity change are almost negligible leads to larger electric conductivity at a given temperature (T).
after the adsorption of four VOCs on the Se-layer and Te-layer To eliminate the coefficient in the equation for calculating
of MoSSe and MoSTe monolayers. Likewise, the sensitivity electronic conductivity, we calculated the conductivity change
based on the conductivity change of the Janus MoSeTe toward by using the following formula
these four VOCs is also quite poor. Consequently, our results σads − σTMD
indicate that the sulfur layer of Janus MoSSe and MoSTe, Δσ = × 100%
namely, the modification of the MoS2 monolayer to Janus σTMD (3)
materials, can efficiently enhance sensitivity with respect to the
where σads and σTMD are the electrical conductivities of after and
acetone molecule. This finding might provide guidance in the
before CB adsorption on the various nanotubes considered in
designing of new 2D volatile organic compound sensors.
the present work, respectively.
4. COMPUTATIONAL DETAILS
All periodic DFT calculations in this study were performed with
the generalized gradient approximation (GGA) of Perdew−
■ ASSOCIATED CONTENT
* Supporting Information

Burke−Ernzerhof (PBE) exchange−correlation functional47 The Supporting Information is available free of charge at
employing Vienna ab initio simulation program (VASP).48−51 https://pubs.acs.org/doi/10.1021/acsomega.0c04938.
The projector augmented wave (PAW) method52,53 was applied Calculated electronic band structure of before and after
to describe the electron core interactions. The Kohn−Sham the VOCs’ adsorption on pristine and Janus TMD
orbitals are expanded in a plane-wave basis set with a kinetic monolayers and top and side views of the optimized
energy cutoff of 520 eV. All 2D transition metal dichalcogenide adsorption structure of VOCs on MoSe2, MoTe2, and
monolayers were studied using the 2H phase of either pristine Janus MoTeSe monolayers (PDF)
MoX2 (X = S, Se, and Te) or Janus MoXY (X ≠ Y = S, Se, and
Te). The (4 × 4) supercell is considered for all TMD
monolayers with 15 Å vacuum along the c-axis to avoid
interlayer interaction, as shown in Figure 1. As dipole interaction
■ AUTHOR INFORMATION
Corresponding Author
was important for the Janus TMDs, dipole corrections Chen-Hao Yeh − Department of Materials Science and
perpendicular to all monolayers were carried out in the Engineering, Feng Chia University, Seatwen, Taichung 40724,
calculations. The Brillouin zone was sampled using (4 × 4 × Taiwan; orcid.org/0000-0001-5665-6496;
1) γ centered Monkhorst−Pack k-point mesh54 for VOCs’ Email: chenhyeh@fcu.edu.tw
adsorption on different TMDs. The convergence threshold was
set to be 1 × 10−5 eV for the total electronic energy in the self- Complete contact information is available at:
consistent loop. The atomic positions were relaxed using the https://pubs.acs.org/10.1021/acsomega.0c04938
quasi-Newton algorithm until the x-, y-, and z-components of
the unconstrained atomic force were smaller than 1 × 10−2 eV/ Notes
The author declares no competing financial interest.


Å. The dispersion energy correction was considered using the
DFT-D3 method by Grimme et al.55,56 In addition, many
previous studies have reported that the spin−orbit coupling ACKNOWLEDGMENTS
(SOC) interaction plays an important role in the 2D TMD The authors gratefully acknowledge the financial support for this
monolayers.39,57−59 Hence, we have included the spin−orbit work from the grant 109-2113-M-035-003-MY3 of the Ministry
coupling (SOC) interaction in the calculations.60 of Science and Technology (MOST), Taiwan. The National
The following equation defines the adsorption energies (Eads) Center of High-Performance Computing (NCHC) contributed
of VOCs on the TMD monolayers to this project by allowing access to their computer facilities and
Eads = E VOC/TMD − E VOC − E TMD donating computer time. The authors greatly acknowledge Prof.
(1)
Jyh-Chiang Jiang (Department of Chemical Engineering,
where EVOC/TMD is the total energy of VOCs adsorbed on National Taiwan University of Science and Technology,
pristine MoX2 or Janus MoXY, EVOC is the energy of a free Taiwan) for support and helpful discussions.
31404 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega


http://pubs.acs.org/journal/acsodf Article

REFERENCES two-dimensional topological phases in Janus materials by substitutional


doping in transition metal dichalcogenide monolayers. npj 2D Mater.
(1) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Appl. 2019, 3, 35.
Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. Electric field effect in (22) Li, F.; Wei, W.; Huang, B.; Dai, Y. Excited-State Properties of
atomically thin carbon films. Science 2004, 306, 666−669. Janus Transition-Metal Dichalcogenides. J. Phys. Chem. C 2019, 124,
(2) Chen, Z.; Wang, J.; Pan, D.; Wang, Y.; Noetzel, R.; Li, H.; Xie, P.; 1667−1673.
Pei, W.; Umar, A.; Jiang, L.; Li, N.; Rooij, N. F.; Zhou, G. Mimicking a
(23) Gao, F.; Goodman, D. W. CO oxidation over ruthenium:
Dogʼs Nose: Scrolling Graphene Nanosheets. ACS Nano 2018, 12,
identification of the catalytically active phases at near-atmospheric
2521−2530.
pressures. Phys. Chem. Chem. Phys. 2012, 14, 6688−6697.
(3) Fiori, G.; Bonaccorso, F.; Iannaccone, G.; Palacios, T.; Neumaier,
(24) Ishizu, Y. General equation for the estimation of indoor
D.; Seabaugh, A.; Banerjee, S. K.; Colombo, L. Electronics based on
pollution. Environ. Sci. Technol. 1980, 14, 1254−1257.
two-dimensional materials. Nat. Nanotechnol. 2014, 9, 768−779.
(25) Donarelli, M.; Ottaviano, L. 2D Materials for Gas Sensing
(4) Kim, Y. H.; Kim, S. J.; Kim, Y. J.; Shim, Y. S.; Kim, S. Y.; Hong, B.
Applications: A Review on Graphene Oxide, MoS(2), WS(2) and
H.; Jang, H. W. Self-Activated Transparent All-Graphene Gas Sensor
Phosphorene. Sensors 2018, 18, 3638.
with Endurance to Humidity and Mechanical Bending. ACS Nano
(26) Zeng, Y.; Lin, S.; Gu, D.; Li, X. Two-Dimensional Nanomaterials
2015, 9, 10453−10460.
for Gas Sensing Applications: The Role of Theoretical Calculations.
(5) Kim, Y. H.; Park, J. S.; Choi, Y.-R.; Park, S. Y.; Lee, S. Y.; Sohn, W.;
Nanomaterials 2018, 8, 851.
Shim, Y.-S.; Lee, J.-H.; Park, C. R.; Choi, Y. S.; Hong, B. H.; Lee, J. H.;
(27) Kim, J. S.; Yoo, H. W.; Choi, H. O.; Jung, H. T. Tunable volatile
Lee, W. H.; Lee, D.; Jang, H. W. Chemically fluorinated graphene oxide
for room temperature ammonia detection at ppb levels. J. Mater. Chem. organic compounds sensor by using thiolated ligand conjugation on
A 2017, 5, 19116−19125. MoS2. Nano Lett. 2014, 14, 5941−5947.
(6) Mas-Ballesté, R.; Gomez-Navarro, C.; Gomez-Herrero, J.; (28) Cho, S. Y.; Koh, H. J.; Yoo, H. W.; Kim, J. S.; Jung, H. T. Tunable
Zamora, F. 2D materials: to graphene and beyond. Nanoscale 2011, Volatile-Organic-Compound Sensor by Using Au Nanoparticle
3, 20−30. Incorporation on MoS2. ACS. Sens. 2017, 2, 183−189.
(7) Novoselov, K. S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A. (29) Chaurasiya, R.; Dixit, A. Defect engineered MoSSe Janus
H. 2D materials and van der Waals heterostructures. Science 2016, 353, monolayer as a promising two dimensional material for NO2 and NO
No. aac9439. gas sensing. Appl. Surf. Sci. 2019, 490, 204−219.
(8) Sun, Z.; Liao, T.; Dou, Y.; Hwang, S. M.; Park, M. S.; Jiang, L.; (30) Chaurasiya, R.; Dixit, A. Ultrahigh sensitivity with excellent
Kim, J. H.; Dou, S. X. Generalized self-assembly of scalable two- recovery time for NH3 and NO2 in pristine and defect mediated Janus
dimensional transition metal oxide nanosheets. Nat. Commun. 2014, 5, WSSe monolayers. Phys. Chem. Chem. Phys. 2020, 22, 13903−13922.
No. 3813. (31) Jin, C.; Tang, X.; Tan, X.; Smith, S. C.; Dai, Y.; Kou, L.; Janus, A.
(9) Torrisi, F.; Coleman, J. N. Electrifying inks with 2D materials. Nat. MoSSe monolayer: a superior and strain-sensitive gas sensing material.
Nanotechnol. 2014, 9, 738−739. J. Mater. Chem. A 2019, 7, 1099−1106.
(10) Huang, J. K.; Pu, J.; Hsu, C. L.; Chiu, M. H.; Juang, Z. Y.; Chang, (32) Chen, W. Y.; Yen, C. C.; Xue, S.; Wang, H.; Stanciu, L. A. Surface
Y. H.; Chang, W. H.; Iwasa, Y.; Takenobu, T.; Li, L. J. Large-area Functionalization of Layered Molybdenum Disulfide for the Selective
synthesis of highly crystalline WSe(2) monolayers and device Detection of Volatile Organic Compounds at Room Temperature. ACS
applications. ACS Nano 2014, 8, 923−930. Appl. Mater. Interfaces 2019, 11, 34135−34143.
(11) Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A. (33) Roy, A.; Movva, H. C.; Satpati, B.; Kim, K.; Dey, R.; Rai, A.;
2D transition metal dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033. Pramanik, T.; Guchhait, S.; Tutuc, E.; Banerjee, S. K. Structural and
(12) Tan, C.; Lai, Z.; Zhang, H. Ultrathin Two-Dimensional Electrical Properties of MoTe2 and MoSe2 Grown by Molecular Beam
Multinary Layered Metal Chalcogenide Nanomaterials. Adv. Mater. Epitaxy. ACS Appl. Mater. Interfaces 2016, 8, 7396−7402.
2017, 29, No. 1701392. (34) Wakabayashi, N.; Smith, H. G.; Nicklow, R. M. Lattice dynamics
(13) Wang, J.; Wei, Y.; Li, H.; Huang, X.; Zhang, H. Crystal phase of hexagonal MoS2studied by neutron scattering. Phys. Rev. B 1975, 12,
control in two-dimensional materials. Sci. China: Chem. 2018, 61, 659−663.
1227−1242. (35) Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically
(14) Zhang, X.; Lai, Z.; Ma, Q.; Zhang, H. Novel structured transition thin MoS(2): a new direct-gap semiconductor. Phys. Rev. Lett. 2010,
metal dichalcogenide nanosheets. Chem. Soc. Rev. 2018, 47, 3301− 105, No. 136805.
3338. (36) Lu, X.; Utama, M. I.; Lin, J.; Gong, X.; Zhang, J.; Zhao, Y.;
(15) Zhou, J.; Wang, Q.; Sun, Q.; Chen, X. S.; Kawazoe, Y.; Jena, P. Pantelides, S. T.; Wang, J.; Dong, Z.; Liu, Z.; Zhou, W.; Xiong, Q.
Ferromagnetism in semihydrogenated graphene sheet. Nano. Lett. Large-area synthesis of monolayer and few-layer MoSe2 films on SiO2
2009, 9, 3867−3870. substrates. Nano. Lett. 2014, 14, 2419−2425.
(16) Lu, A. Y.; Zhu, H.; Xiao, J.; Chuu, C. P.; Han, Y.; Chiu, M. H.; (37) Ruppert, C.; Aslan, O. B.; Heinz, T. F. Optical properties and
Cheng, C. C.; Yang, C. W.; Wei, K. H.; Yang, Y.; Wang, Y.; Sokaras, D.; band gap of single- and few-layer MoTe2 crystals. Nano. Lett. 2014, 14,
Nordlund, D.; Yang, P.; Muller, D. A.; Chou, M. Y.; Zhang, X.; Li, L. J. 6231−6236.
Janus monolayers of transition metal dichalcogenides. Nat. Nano- (38) Jin, H.; Wang, T.; Gong, Z. R.; Long, C.; Dai, Y. Prediction of an
technol. 2017, 12, 744−749. extremely long exciton lifetime in a Janus-MoSTe monolayer. Nanoscale
(17) Zhang, J.; Jia, S.; Kholmanov, I.; Dong, L.; Er, D.; Chen, W.; Guo, 2018, 10, 19310−19315.
H.; Jin, Z.; Shenoy, V. B.; Shi, L.; Lou, J. Janus Monolayer Transition- (39) Yang, X.; Singh, D.; Xu, Z.; Wang, Z.; Ahuja, R. An emerging
Metal Dichalcogenides. ACS Nano 2017, 11, 8192−8198. Janus MoSeTe material for potential applications in optoelectronic
(18) Dong, L.; Lou, J.; Shenoy, V. B. Large In-Plane and Vertical devices. J. Mater. Chem. C 2019, 7, 12312−12320.
Piezoelectricity in Janus Transition Metal Dichalchogenides. ACS Nano (40) Tang, X.; Du, A.; Kou, L. Gas sensing and capturing based on
2017, 11, 8242−8248. two-dimensional layered materials: Overview from theoretical
(19) Li, R.; Cheng, Y.; Huang, W. Recent Progress of Janus 2D perspective. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2018, 8,
Transition Metal Chalcogenides: From Theory to Experiments. Small No. e1361.
2018, 14, No. e1802091. (41) Tang, X.; Shang, J.; Gu, Y.; Du, A.; Kou, L. Reversible gas capture
(20) Cheng, Y. C.; Zhu, Z. Y.; Tahir, M.; Schwingenschlögl, U. Spin- using a ferroelectric switch and 2D molecule multiferroics on the
orbit−induced spin splittings in polar transition metal dichalcogenide In2Se3 monolayer. J. Mater. Chem. A 2020, 8, 7331−7338.
monolayers. Europhys. Lett. 2013, 102, 57001. (42) Wei, Y.; Xu, X.; Wang, S.; Li, W.; Jiang, Y. Second harmonic
(21) Maghirang, A. B.; Huang, Z.-Q.; Villaos, R. A. B.; Hsu, C.-H.; generation in Janus MoSSe a monolayer and stacked bulk with vertical
Feng, L.-Y.; Florido, E.; Lin, H.; Bansil, A.; Chuang, F.-C. Predicting asymmetry. Phys. Chem. Chem. Phys. 2019, 21, 21022−21029.

31405 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406
ACS Omega http://pubs.acs.org/journal/acsodf Article

(43) Pang, K.; Wei, Y.; Li, W.; Zhou, X.; Jiang, Y.; Yang, J.; Li, X.; Gao, (66) Yeh, C.-H.; Lin, W.-Y.; Jiang, J.-C. Enhancement of
L.; Jiang, Y. Highly sensitive gas sensing material for polar gas molecule chlorobenzene sensing by doping aluminum on nanotubes: A DFT
based on Janus group-III chalcogenide monolayers: A first-principles study. Appl. Surf. Sci. 2020, 514, No. 145897.
investigation. Sci. China: Technol. Sci. 2020, 63, 1566−1576.
(44) Pang, K.; Wei, Y.; Xu, X.; Li, W.; Yang, J.; Zhang, G.; Li, X.; Ying,
T.; Jiang, Y. Modulation of the electronic band structure of silicene by
polar two-dimensional substrates. Phys. Chem. Chem. Phys. 2020, 22,
21412−21420.
(45) Kou, L.; Frauenheim, T.; Chen, C. Phosphorene as a Superior
Gas Sensor: Selective Adsorption and Distinct I-V Response. J. Phys.
Chem. Lett. 2014, 5, 2675−2681.
(46) Babar, V.; Sharma, S.; Schwingenschlögl, U. Highly Sensitive
Sensing of NO and NO2 Gases by Monolayer C3N. Adv. Theory Simul.
2018, 1, No. 1700008.
(47) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient
Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868.
(48) Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid
metals. Phys. Rev. B 1993, 47, 558−561.
(49) Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of
the liquid-metal-amorphous-semiconductor transition in germanium.
Phys. Rev. B 1994, 49, 14251−14269.
(50) Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio
total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996,
54, 11169−11186.
(51) Kresse, G.; Furthmüller, 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.
(52) Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B
1994, 50, 17953−17979.
(53) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the
projector augmented-wave method. Phys. Rev. B 1999, 59, 1758−1775.
(54) Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone
integrations. Phys. Rev. B 1976, 13, 5188−5192.
(55) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and
accurate ab initio parametrization of density functional dispersion
correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010,
132, No. 154104.
(56) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping
function in dispersion corrected density functional theory. J. Comput.
Chem. 2011, 32, 1456−1465.
(57) Alidoust, N.; Bian, G.; Xu, S. Y.; Sankar, R.; Neupane, M.; Liu, C.;
Belopolski, I.; Qu, D. X.; Denlinger, J. D.; Chou, F. C.; Hasan, M. Z.
Observation of monolayer valence band spin-orbit effect and induced
quantum well states in MoX2. Nat. Commun. 2014, 5, No. 4673.
(58) Cheng, C.; Sun, J. T.; Chen, X. R.; Fu, H. X.; Meng, S. Nonlinear
Rashba spin splitting in transition metal dichalcogenide monolayers.
Nanoscale 2016, 8, 17854−17860.
(59) Fan, X.; Singh, D. J.; Zheng, W. Valence Band Splitting on
Multilayer MoS2: Mixing of Spin-Orbit Coupling and Interlayer
Coupling. J. Phys. Chem. Lett. 2016, 7, 2175−2181.
(60) Steiner, S.; Khmelevskyi, S.; Marsmann, M.; Kresse, G.
Calculation of the magnetic anisotropy with projected-augmented-
wave methodology and the case study of disordered Fe1−xCox alloys.
Phys. Rev. B 2016, 93, No. 224425.
(61) Limas, N. G.; Manz, T. A. Introducing DDEC6 atomic
population analysis: part 2. Computed results for a wide range of
periodic and nonperiodic materials. RSC Adv. 2016, 6, 45727−45747.
(62) Manz, T. A. Introducing DDEC6 atomic population analysis:
part 3. Comprehensive method to compute bond orders. RSC Adv.
2017, 7, 45552−45581.
(63) Manz, T. A.; Limas, N. G. Introducing DDEC6 atomic
population analysis: part 1. Charge partitioning theory and method-
ology. RSC Adv. 2016, 6, 47771−47801.
(64) Ahmadi, A.; Hadipour, N. L.; Kamfiroozi, M.; Bagheri, Z.
Theoretical study of aluminum nitride nanotubes for chemical sensing
of formaldehyde. Sens. Actuators, B 2012, 161, 1025−1029.
(65) Yeh, C.-H.; Hsiao, Y.-J.; Jiang, J.-C. Dopamine sensing by boron
and nitrogen co-doped single-walled carbon nanotubes: A first-
principles study. Appl. Surf. Sci. 2019, 473, 59−64.

31406 https://dx.doi.org/10.1021/acsomega.0c04938
ACS Omega 2020, 5, 31398−31406

You might also like