You are on page 1of 30

Subscriber access provided by TEXAS A&M INTL UNIV

Article
Structural Phase Stability Control of Monolayer
MoTe with Adsorbed Atoms and Molecules
2

Yao Zhou, and Evan J. Reed


J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b05770 • Publication Date (Web): 25 Aug 2015
Downloaded from http://pubs.acs.org on August 31, 2015

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts
appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered
to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just
Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society.


1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 29 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
Structural Phase Stability Control of Monolayer
9
10
11
12 MoTe2 with Adsorbed Atoms and Molecules
13
14
15
16
17 Yao Zhou and Evan J. Reed*
18
19
20 Department of Materials Science and Engineering, Stanford University, Stanford, California
21
22
23
94305, United States
24
25
26 *E-mail: evanreed@stanford.edu. Tel: [+1] (650) 723 2971. Fax: [+1] (650) 725-4034
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
1
The Journal of Physical Chemistry Page 2 of 29

1
2
3
4
5
6
7
8
9
10 Structural Phase Stability Control of Monolayer
11
12
13
14 MoTe2 with Adsorbed Atoms and Molecules
15
16
17
18
19 ABSTRACT
20
21
22
23 We study the adsorption of some common atoms and molecules onto monolayer MoTe2 and
24
25 the potential for adsorption to induce a structural phase change between the semiconducting 2H-
26
27
28 based and metallic 1T’-based crystal structures of the monolayer. Using density functional
29
30 theory with spin orbit and van der Waals energy contributions, we determined energetically
31
32
33
favorable adsorption positions and orientations on the two crystalline phases of monolayer
34
35 MoTe2. We then obtained the formation energies for these adsorption reactions and found that
36
37 atomic adsorption generally favors 1T’ metallic phases while molecular adsorption favors
38
39
40 semiconducting 2H phases. The phase sensitivity of this material is due to a relatively small
41
42 energy difference, approximately 31 meV per MoTe2 formula unit. We further find that the
43
44 monolayer alloy MoxW1-xTe2 can exhibit some degree of molecular selectivity in phase changes,
45
46
47 potentially providing the basis for molecular sensing applications.
48
49
50
51
52
53
54
55
56 Introduction
57
58
59
60
ACS Paragon Plus Environment
2
Page 3 of 29 The Journal of Physical Chemistry

1
2
3
The synthesis and subsequent discovery of the unusual properties of graphene1 has created
4
5
6 interest in a broader spectrum of materials that can be isolated in a nearly atomically thin form.
7
8 Transition metal dichalcogenide (TMD) monolayers, where the transition metal can be Mo or W
9
10
11
and the chalcogenide can be S, Se or Te, are one such group of monolayers with special
12
13 structural features. Unlike graphene, these TMD monolayers have the potential to exist in more
14
15 than one possible crystal structure. One is the semiconducting 2H phase2 in which the transition
16
17
18 metal atoms are trigonal prismatically coordinated by the chalcogen atoms. The others are the
19
20 metallic 1T phase with transition metal atoms in octahedral coordination, and a distorted version
21
22 of this phase that we will refer to as 1T’ phase where transition metal atoms are also
23
24
25 approximately octahedrally coordinated but with lower symmetry3,4. The 2D lattice of a 1T’
26
27 monolayer is rectangular while the 2D lattices of the 2H and 1T monolayers are hexagonal.
28
29
Earlier DFT-based calculations have been reported for these Mo and W-based monolayer TMDs
30
31
32 indicating that the 2H phase is the lowest energy structure for all except WTe2, which exhibits
33
34 lowest energy in the 1T’ phase4.
35
36
37
38 Chemical control over the crystal structure has been reported in some transition metal
39
40 dichaocogenides5, 6, including the potential for spatial control of the distribution of phases on the
41
42 monolayer7. The interface between semiconducting and metallic phases on MoS2 has been
43
44
45 reported to exhibit favorable electrical contact resistance characteristics, pointing to possible
46
47 electronic device applications for phase engineering8.
48
49
50
Among the Mo- and W- based TMD monolayers, MoTe2 is particularly interesting for phase
51
52
53 change applications because it has the smallest energy difference between 2H phase and 1T’
54
55 phase as shown in Figure 1. We computed a 31 meV energy difference between these phases for
56
57
58
MoTe2 a monolayer at zero stress conditions using a DFT calculation including spin-orbit
59
60
ACS Paragon Plus Environment
3
The Journal of Physical Chemistry Page 4 of 29

1
2
3
coupling effects as discussed in the Methods section. This relatively small energy difference
4
5
6 leads to the reported possibility that tensile uniaxial strains of several percent can change the
7
8 energy ordering of the phases, leading to a structural phase transition4. There is also
9
10
11
experimental evidence that a structural transition from 2H to 1T’ can occur in MoTe2 at high
12
13 temperatures9,10.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Figure 1. Three crystalline phases of monolayer MoTe2 represented in a 1×1 rectangular unit
31
32 cell. 2H phase is trigonal prismatic structure and semiconducting, while 1T phase and 1T’ phase
33
34
35 are octahedral and a distorted octahedral structure, respectively. 1T phase and 1T’ phase are both
36
37 metallic. Top view, side view and 3D view of each phase are shown in the figure.
38
39
40
41
42
43
44 The proximity of monolayer MoTe2 to a structural phase transition suggests the phase stability of
45
46 this material could be sensitive to the local chemical environment due to its nearly atomic
47
48
thickness. This property could lead to complexities or opportunities in controlling the phase
49
50
51 synthesized by a growth process, e.g. chemical vapor deposition (CVD). Some knowledge of the
52
53 effect of adsorbed molecules on phase stability may provide guidance for controlling the
54
55
56
energetically favored phase during synthesis. For example, one would like to know if there exist
57
58
59
60
ACS Paragon Plus Environment
4
Page 5 of 29 The Journal of Physical Chemistry

1
2
3
molecules that can be added to the system during growth that will bias toward a particular phase.
4
5
6 DFT-based adsorption sites for H, O, and F atoms on the 2H phase of MoTe2 monolayer without
7
8 van der Waals corrections and spin orbit coupling have been reported11, indicating minimum
9
10
11
energy binding positions similar to those found here.
12
13
14 Another motivation to study the phase stability in the presence of molecules or adatoms is for
15
16 chemical sensing applications. Two-dimensional materials are good candidates for sensing
17
18
19 materials due to their high surface to volume ratio. Carbon nanotubes12,13 and graphene14,15 have
20
21 been reported to provide solid-state gas sensors with exceptional sensitivity. In the TMD family,
22
23 MoS2 has been reported to be used for glucose and biomolecule detection16, modulation of
24
25
26 photoluminescence by molecular physisorption gating17 or ion intercalation18, and gas detection
27
28 for NH3, NO2, water vapor19, and triethylamine20. Li intercalated in MoS2 has been reported to
29
30
31
yield a structural change from 2H to 1T’21,22,23,24. The MoTe2 monolayer has the potential to
32
33 exhibit more sensitivity to molecular adsorption due to the small energy difference between 2H
34
35 and 1T’ due to the large expected changes in electronic properties across the transition. DFT
36
37
38 calculations of MoTe2 indicate that 2H is semiconducting and 1T’ is metallic or semimetallic,4
39
40 and some experimental evidence for significant electronic changes has been reported10. This
41
42 semiconducting to metallic phase transition induced by molecular or atomic adsorption could
43
44
45 provide a mechanism for enhanced chemical sensitivity or specificity.
46
47
48 Field effect transistors have been reported using single layer MoTe225. Optical properties of
49
50
single- and few-layer MoTe2 have been studied and strong photoluminescence has been observed
51
52
53 in monolayer MoTe226. These studies indicate that few layer MoTe2 is sufficiently stable to
54
55 perform experiments and construct devices.
56
57
58
59
60
ACS Paragon Plus Environment
5
The Journal of Physical Chemistry Page 6 of 29

1
2
3
In this work, we employ density functional methods to explore the potential for adsorbed
4
5
6 molecules and atoms to bias the favored thermodynamic phase of MoTe2. We study the
7
8 formation energies of a range of adsorbed small molecules and adatoms for binding to the
9
10
11
surface of 2H and 1T’ MoTe2 to determine the potential for inducing a phase transformation by
12
13 exposure to these small molecules.
14
15
16
17
18
19
20 Methods
21
22
23 Density functional theory (DFT) was used to study molecule adsorptions on MoTe2 monolayers.
24
25
All DFT calculations were performed with the Vienna Ab Initio Simulation Package (VASP)27,
26
27
28 using the Projected-Augmented Wave (PAW) method28,29. For Mo, Te, H, C, N, O, Cl and F, we
29
30 used standard PAW pseudopotential implementation methods, treating 6, 6, 1, 4, 5, 6, 7 and 7
31
32
33
valence electrons respectively. Since we expected electrons to transfer from alkali atoms to
34
35 MoTe2 monolayers, for Li, Na and K, we used PAW pseudopotentials where semi-core electrons
36
37 are also treated as valence electrons, including 3, 7 and 7 valence electrons respectively. In
38
39
40 addition, for oxygen molecule adsorption, we used a harder O pseudopotential for formation
41
42 energy calculations for O2 in Figure 4 and Figure 5. Electron exchange and correlation effects
43
44 were treated using the generalized gradient approximation (GGA) functional of Perdew, Burke,
45
46
47 and Ernzerhof30. However, GGA tends to underestimate the binding energies31 in the systems
48
49 where nonlocal correlations play an important role. We have studied a range of methods for
50
51
52
including van der Waals interactions within DFT calculations, including the D3 correction
53
54 method of Grimme (DFT-D3)32, van der Waals density functional (vdW-DF)33 and the
55
56 Tkatchenko-Scheffler method (TS-SCS)34. In addition, due to the absence of inversion symmetry
57
58
59
60
ACS Paragon Plus Environment
6
Page 7 of 29 The Journal of Physical Chemistry

1
2
3
in monolayer MoTe2, it is expected that spin-orbit coupling effects will cause band splitting35.
4
5
6 Especially for MoTe2, where we have the heavy element Te, we expect such spin-orbit coupling
7
8 effect should be strong and non-negligible. In Figure 2, we compared these different methods for
9
10
11
the case of oxygen molecule adsorption. In all subsequent calculations (except those involving
12
13 Li, Na, K) we employ the TS-SCS method which gives results comparable to DFT-D3 for the
14
15 case of Figure 2. Due to the reported poor performance of TS-SCS method for alkali metals36, we
16
17
18 used the DFT-D3 method for Li, Na, K atomic adsorption.
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38 Figure 2. Comparison of formation energies of oxygen molecule adsorption on unit cell of 2H-
39
40 MoTe2 monolayer and 1T’-MoTe2 monolayer for different methods. The methods describing van
41
42
43 der Waals interactions (DFT-D3, vdW-DF, TS-SCS) are done with spin-orbit coupling included.
44
45 We include spin-orbit coupling and van der Waals interactions using TS-SCS or DFT-D3 for all
46
47
subsequent adsorption cases in this manuscript.
48
49
50
51
52
53 We used a plane-wave basis set with different energy cutoffs for different atomic and molecular
54
55 adsorption cases. We chose a cutoff greater than 1.3 times the largest ENMAX value given in the
56
57
58 potential POTCAR-file for each adsorption case to ensure convergence. We made an exception
59
60
ACS Paragon Plus Environment
7
The Journal of Physical Chemistry Page 8 of 29

1
2
3
to this choice for systems with O where we used a hard pseudopotential with the 700eV cutoff
4
5
6 recommended in the POTCAR-file. For Li and O2 with a hard O pseudoptential, we used a
7
8 kinetic energy cutoff of 700 eV. For Na, K, H, H2 and Cl we used a 350 eV energy cutoff. For O2
9
10
11
with a regular O pseudopotential, we used a 600 eV energy cutoff. For H2O, NH3, NO, NO2, CO,
12
13 CO2, N2, O, and F we used a 550 eV energy cutoff. We used the same energy cutoff for
14
15 calculations of each adsorption case. We sampled the Brillouin zone using an 18×18×1
16
17
18 Monkhorst-Pack37 k-point grid. Also, we used a Gaussian electronic distribution smearing of 50
19
20 meV. The computational cell is 20Å along the interlayer direction, which gives a vacuum slab
21
22 distance of approximately 16Å between monolayers. We did DFT calculations for two systems,
23
24
25 one with higher adsorbate concentration including a 1×1 unit cell with two formula units of
26
27 MoTe2 and one adsorbate atom or molecule, and another with lower adsorbate concentration
28
29
including a 2×2 unit cell with eight formula units of MoTe2 and one adsorbate atom or molecule.
30
31
32
33 With periodic boundary conditions and when the computational cell is electrically polarized, the
34
35 requirement for the Kohn-Sham potential to be periodic can introduce undesirable or fictitious
36
37
38 electric fields that exist throughout the vacuum region of the cell38,39,40. These were corrected
39
40 using the method of Neugebauer and Scheffler, which provides a correction for the total
41
42 energy40. To determine a likely minimum energy configuration for each adsorbed molecule or
43
44
45 atom, we performed geometry relaxation calculations starting with the molecule at one of three
46
47 initial adsorption sites for the higher adsorbate concentration computational cell containing 2
48
49
formula units of MoTe2. For the 2H phase, the three sites are top of Mo atom, top of Te atom,
50
51
52 and center of hexagon. For the 1T phase, the three sites are the top of Mo atom, and top of two
53
54 different Te atom sites. Multiple calculations were done with different orientations for some
55
56
57
molecules. During relaxations, we found all adsorptions on 1T-MoTe2 relaxed to 1T’-MoTe2.
58
59
60
ACS Paragon Plus Environment
8
Page 9 of 29 The Journal of Physical Chemistry

1
2
3
Preferential binding positions and orientations for each adsorbate on 2H-MoTe2 and 1T’-MoTe2
4
5
6 were taken to be the lowest energy configuration obtained for all initial conditions. To obtain the
7
8 configurations at lower concentration with the 8 formula unit MoTe2 cell, we used these lowest
9
10
11
energy configurations as initial configuration and performed geometry relaxation calculations.
12
13 Starting from the lowest energy geometry, we performed further geometry relaxation including
14
15 van der Waals energy using the TS-SCS method or DFT-D3 method and performed a self-
16
17
18 consistent spin-orbit coupling DFT calculation, except for the H2 case where we only did spin
19
20 restricted calculation due to convergence difficulties . We use the equilibrium lattice constants
21
22 of 2H-MoTe2 monolayer for all cases which assumes that substrate friction may prevent lattice
23
24
25 constants from changing as discussed in the next section. Allowing the lattice constants to relax
26
27 to zero stress may be more applicable to a freely-suspended monolayer and may give different
28
29
results.
30
31
32
33
34
35
36 Results and Discussion
37
38
39
40 We wish to study the potential for the adsorption of small molecules and atoms to the 2H and
41
42 1T’ phases of monolayer MoTe2. In the absence of adsorbates. the 2H phase is energetically
43
44 stable at ambient conditions and the 1T’ phase of the bulk material is reported to be stable at high
45
46
47 temperatures9. We wish to determine the potential for bound molecules and atoms to alter the
48
49 energetically favored phase, 2H or 1T’.
50
51
52
53
The mechanical constraints on the monolayer are expected to play some role in the phase
54
55 stability4 and therefore require consideration. At zero stress conditions, the 1T’ and 2H phases
56
57 of the MoTe2 monolayer exhibit different lattice constants (2.7% and 3.7% strain in the a and b
58
59
60
ACS Paragon Plus Environment
9
The Journal of Physical Chemistry Page 10 of 29

1
2
3
lattice directions of the rectangular cell from 2H phase to 1T’ phase). Therefore, a change in area
4
5
6 of the monolayer would be expected if a transition from one phase to the other occurs at
7
8 conditions of zero stress. However, a monolayer on top of a substrate may experience substrate
9
10
11
friction that could prevent it from changing area if it remains planar. In the latter case, one might
12
13 expect that the lattice constants of the material would remain approximately the same when the
14
15 transition occurs.
16
17
18
19 A third scenario may occur if the monolayer delaminates from the substrate, forming bubbles or
20
21 wrinkles to relax the internal stress. To obtain a simple estimate of the likelihood of
22
23 delamination effects, we used DFT-based calculations of elastic moduli of 2H-MoTe2
24
25

26 monolayer41 to estimate the strain energy density ‫ = ܧ‬ଶ ‫ܥ‬௜௝ ߝ௜ ߝ௝ for a 2H-MoTe2 monolayer at the
27
28
29 lattice constants of the 1T’ phase and obtained a value of 0.064 J/m2. Here, ߝ௜ and ߝ௝ are strains
30
31 in the monolayer and ‫ܥ‬௜௝ are elastic stiffness coefficients. If this energy density is significantly
32
33
34 greater than the energy of adhesion to the substrate, one would expect the monolayer to
35
36 delaminate or ripple to relieve some of this energy42. While the substrate binding energy likely
37
38
39
varies considerably with the type of substrate and the phase of MoTe2, some comparison can be
40
41 made to the interlayer binding energy of graphite 0.19 J/m2 43 which is greater than the strain
42
43 energy density. Therefore one might expect the monolayer to remain planar rather than form
44
45
46 bubbles or wrinkles, but stress relieving wrinkles could occur for sufficiently small substrate
47
48 binding energy.
49
50
51 In the calculations that follow, we make the assumption that the monolayer remains planar across
52
53
54 the transition. We furthermore assume that the lattice constants remain fixed across the
55
56
57
58
59
60
ACS Paragon Plus Environment
10
Page 11 of 29 The Journal of Physical Chemistry

1
2
3
transition, a condition that may be most appropriate for a monolayer constrained by substrate
4
5
6 friction. We use the equilibrium lattice constants of the 2H-MoTe2 monolayer for all cases.
7
8
9 Figure 3 shows the preferential binding positions and orientations of molecules and atoms on
10
11
12 2H-MoTe2 and 1T’-MoTe2 monolayer determined as described in Methods. We calculated
13
14 formation energies for adsorption reactions with 2H-MoTe2 and 1T’-MoTe2 using the energies of
15
16 these structures including the van der Waals and spin-orbit contributions.
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
11
The Journal of Physical Chemistry Page 12 of 29

1
2
3
Figure 3. Most energetically favorable adsorption positions for Li, Na, K, O, Cl, H, F, H2, H2O,
4
5
6 NH3, NO, NO2, CO, CO2, N2, O2 on 2×2 depicted supercell of (a) 2H-MoTe2 monolayer, (b) 1T’-
7
8 MoTe2 monolayer. Top view, side view and 3D view for each adsorption case are shown in
9
10
11 the figure. Dashed black lines are drawn between nearest atoms to aid in visualization.
12
13
14
15
(a)
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30 Li Na K O Cl H2 H2O NH3 NO NO2 CO CO2 N2 O2 no X
31
32
(b)
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
Figure 4. Formation energy of Li, Na, K, O, Cl, H, F, H2, H2O, NH3, NO, NO2, CO, CO2, N2, O2
53
54
55 atomic or molecular adsorption on MoTe2 monolayer with initial structure (a)2H phase (stable
56
57 phase at room temperature) and (b)1T’ phase (stable phase at high temperature). Two different
58
59
60
ACS Paragon Plus Environment
12
Page 13 of 29 The Journal of Physical Chemistry

1
2
3
adsorbate concentrations (one atom or molecule per unit cell or 2×2 supercell of MoTe2) are
4
5
6 considered. Part (a) indicates that binding of molecules would energetically favor 2H phase
7
8 while binding of Li, Na, K, H, O at higher concentration, Cl, and F would potentially cause
9
10
11
phase transition from 2H to 1T’. The magnitude of formation energy depends on the choice of
12
13 reference states (alkali metals and isolated molecules are used in this work). Atomic H and F are
14
15 shown in an inset table due to the large magnitudes associated with the choice of reference states
16
17
18 for these atoms (H2 and F2 respectively).
19
20
21
22
23
24 The blue and red bars in Figure 4(a) show the binding energies of molecules to the 2H phase for
25
26
27 two different concentrations. These bars indicate the energy of the atom or molecule adsorbed on
28
29 2H-MoTe2 monolayer minus the energy of an isolated 2H-MoTe2 monolayer and the energy of
30
31 the reference states of the atom or molecule. Here the reference states are taken to be Li metal,
32
33
34 Na metal, K metal, O2 molecule, Cl2 molecule, H2 molecule, H2O molecule, NH3 molecule, NO
35
36 molecule, NO2 molecule, CO molecule, CO2 molecule, N2 molecule, and F2 molecule. The
37
38
choice of reference states impacts the binding energy to the 2H phase. We calculated reference
39
40
41 state energies using DFT with the same methods as discussed in the Methods section. The blue
42
43 and red bars in Figure 4(a) show that the molecules generally bind to the 2H phase while the
44
45
46
atoms do not. While we expect typical DFT errors in these energies of tenths of eV34, 36, trends
47
48 can be established with these calculations.
49
50
51 The green and yellow bars in Figure 4(a) show the energy change associated with a species
52
53
54 binding to the 2H phase and inducing a phase change to 1T’ for two different concentrations.
55
56 These bars indicate the energy of the atom or molecule adsorbed on 1T’-MoTe2 monolayer
57
58
59
60
ACS Paragon Plus Environment
13
The Journal of Physical Chemistry Page 14 of 29

1
2
3
minus the energy of an isolated 2H-MoTe2 monolayer minus the energy of the reference states of
4
5
6 the atom or molecule. At higher concentrations (green bars), some molecules and atoms have
7
8 the potential to induce a phase change to 1T’. As the concentration decreases (yellow bars), the
9
10
11
energies are positive for all molecules indicating that a phase change to 1T’ is not likely.
12
13
14 The trend with concentration results from the fact that the energy required to convert a region of
15
16 2H to 1T’ increases with the area. The energy per adsorbate required to convert to 1T’ increases
17
18
19 as the concentration decreases. This trend is indicated in the green and yellow bars on the right
20
21 of Figure 4(a) labeled “No X” that show the energies of these reactions with no adsorbate
22
23 molecule.
24
25
26
27 The monolayer phase that is energetically favored upon adsorption of a molecule is determined
28
29 by the difference between the blue and green bars for higher concentrations, and red and yellow
30
31
32
bars for lower concentrations. This energy difference is independent of choice of reference state.
33
34 Figure 4(a) indicates that adsorption of the molecules generally stabilizes the 2H phase, while
35
36 adsorption of adatoms usually stabilizes the 1T’ phase.
37
38
39
40 Figure 4(b) is analogous to Figure 4(a) for 1T’ as the initial phase. All adsorbates studied are
41
42 found to bind to the 1T’ phase (green and yellow bars). For the molecules, a subsequent change
43
44 from the 1T’ to 2H is energetically favored upon binding, while the atoms will generally stabilize
45
46
47 the 1T’ phase upon binding.
48
49
50 Since the phase change is expected to be accompanied by a significant change in electronic
51
52
53
properties, we consider the potential for gas molecular sensing applications and the degree of
54
55 selectivity. Starting in the 2H phase, the lowest energy phase, Figure 4(a) indicates that the
56
57 molecules considered here will bind, but none are expected to result in a phase change to the 1T’
58
59
60
ACS Paragon Plus Environment
14
Page 15 of 29 The Journal of Physical Chemistry

1
2
3
phase. NO2 may be the closest to causing a phase change of the molecules considered. Figure
4
5
6 4(b) indicates that starting in the metastable 1T’ phase, all molecules will bind and favor
7
8 conversion to the 2H phase.
9
10
11
12 Figure 4 assumes that substrate friction fixes the lattice constants across the transition at the 2H
13
14 values. If substrate friction is negligible or the monolayer exhibits stress relieving wrinkles
15
16 (discussed above), a constant stress mechanical constraint may be more relevant. Previous
17
18
19 theoretical work has demonstrated that the boundary of a phase transition in monolayer MoTe2 is
20
21 sensitive to whether the transition occurs at a condition of constant stress or constant area4, in
22
23 analog with constant pressure or constant volume for bulk materials. The energy difference
24
25
26 between the 2H and 1T’ phases of the monolayer at zero stress conditions is 31 meV/f.u., 44
27
28 meV/f.u. lower than the value of 75 meV/f.u. when the lattice constant are fixed at the 2H values.
29
30
31
The closer proximity of the phase boundary at constant stress conditions has the potential to
32
33 increase the likelihood that adsorbates that will induce the 2H to 1T’ transition.
34
35
36 To explore the role of stress relaxation, we did calculations for N2 and Na adsorption under a
37
38
39 zero stress condition at higher concentration. In the case of Na adsorption under zero stress, we
40
41 find that the 1T’ phase is favored by 193 meV/f.u., 44 meV/f.u. higher than the corresponding
42
43 energy for the constant area case. In the case of N2 adsorption under zero stress, we find that the
44
45
46 2H phase is favored by 65 meV/f.u., 13 meV/f.u. lower than the corresponding energy for the
47
48 constant area case. This suggests that the favored phase for both cases is independent of
49
50
mechanical constraint. However, the zero stress condition pushes the phase boundary toward the
51
52
53 1T’ structure and has the potential to cause the 1T’ phase to be favored for some molecular
54
55 adsorbates, including NO where the 2H phase is stabilized by 30 meV/f.u., and NO2 where the
56
57
58
1T’ phase is stabilized by 9 meV/f.u..
59
60
ACS Paragon Plus Environment
15
The Journal of Physical Chemistry Page 16 of 29

1
2
3
A potential route to increasing the molecular selectivity of this material is to consider an alloy
4
5
6 designed such that the energies of the 1H and 1T’ phases are closer or even identical. Alloys of
7
8 single layer materials including MoS2xSe2(x-1) have been reported44,45,46,47. Here we consider the
9
10
11
alloy MoxW1-xTe2 because the lattice constant of the 2H phase of WTe2 (3.552Å and 6.154 Å
12
13 from DFT) is similar to the 2H lattice constant of MoTe2 monolayer4 (3.550Å and 6.149Å from
14
15 DFT). Unlike monolayer MoTe2, monolayer WTe2 is expected to be 1T’ in the lowest energy
16
17
18 state4,3. It is likely that the energy difference between 2H and 1T’ phases of MoxW1-xTe2 can be
19
20 tuned with fraction x without significant mechanical deformation. Inspired by this, we
21
22 performed the calculations of Figure 4 for the alloy Mo0.5W0.5Te2 with transition metal atoms
23
24
25 randomly distributed, and found that the 1T’ phase is energetically favored, independent of the
26
27 adsorbed species. However, as the composition is tuned through the range 0 < x < 0.5, one might
28
29
30 expect some adsorbed molecules to induce a phase change while others do not. Figure 5 shows
31
32 the results of Figure 4(a) with the green and yellow bars shifted to reflect a zero energy
33
34 difference between the 2H and 1T’ phases, i.e. shifted by the green and yellow bars on the right
35
36
37 side of Figure 4(a). Figure 5 indicates that H2O, NO and NO2 molecules have the greatest
38
39 potential to transform the 2H phase to 1T’ phase in a MoxW1-xTe2 monolayer alloy with x
40
41
chosen to give the same energy for 2H and 1T’ phases. This suggests that alloying may provide
42
43
44 some molecular selectivity. The unknown kinetics of the transition could play a role in the
45
46 utility of this mechanism for molecular sensing.
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
16
Page 17 of 29 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23 Figure 5. Estimated formation energy of small molecule adsorption on a MoxW1-xTe2 monolayer
24
25 alloy with composition x chosen to yield the same energy for the 2H and 1T’ phases. The binding
26
27 energies are approximated by those of pure MoTe2 (shown in Figure 4(a)) with the energy
28
29
30 difference between the 2H and 1T’ phases set to zero. Two different adsorbate concentrations
31
32 (one atom or molecule per unit cell or 2×2 supercell of MoTe2) are considered. Of the molecules
33
34
studied, H2O, NO, and NO2 are the most likely to induce a phase transition by adsorption on the
35
36
37 alloy monolayer, pointing to the possibility for some molecule specificity in a sensing
38
39 application.
40
41
42
43
44
45
46 It is also possible to speculate that these results could have application in the preferential growth
47
48 of phases of MoTe2 and its alloys. By introducing gases during the growth or cooling process, it
49
50
51
may be possible to bias the growth toward the 2H or 1T’ phases. Potential complications include
52
53 the spectrum of chemical reactions not considered here that could occur under high temperature
54
55 growth conditions. Other environmental conditions are expected to impact the favored phase,
56
57
58 including strain4 in the monolayer which may result from the growth and cooling process.
59
60
ACS Paragon Plus Environment
17
The Journal of Physical Chemistry Page 18 of 29

1
2
3
4
5
6
7 Conclusions
8
9
10 We have studied the adsorption of some common atoms and molecules onto monolayer MoTe2
11
12
13 and assessed the potential for adsorption to induce a phase change between the semiconducting
14
15 2H and metallic 1T’ crystal structures of the monolayer. We have employed spin orbit effects
16
17 and studied a variety of van der Waals correction techniques for this purpose. We determined
18
19
20 energetically favorable adsorption positions and orientations on the two phases of monolayer
21
22 MoTe2. We find that atomic adsorption generally induces 1T’ metallic phases while molecular
23
24
adsorption induces 2H phases. We further find that the monolayer alloy MoxW1-xTe2 has the
25
26
27 potential to exhibit some degree of molecular selectivity in phase changes by varying the
28
29 composition x, potentially providing the basis for molecular sensing applications due to the large
30
31
32
electronic contrast between 2H and 1T’ phases. In particular, the calculations indicate that it
33
34 may be possible to engineer an alloy such that specific molecules including NO2 and NO will
35
36 induce a phase change to 1T’ while other molecules studied stabilize the 2H phase. An
37
38
39 additional possible application of this work may be the chemical stabilization of a preferred
40
41 phase during the growth process.
42
43
44
45
46 Acknowledgment
47
48 This work was supported in part by the U.S. Army Research Laboratory, through the Army
49
50
High Performance Computing Research Center, Cooperative Agreement W911NF-07-0027, and
51
52
53 by computer and software support provided by the U.S. Army Engineer Research and
54
55 Development Center (ERDC), DoD Supercomputing Resource Center (DSRC) through the DoD
56
57
58
High Performance Computing Modernization Program (HPCMP).
59
60
ACS Paragon Plus Environment
18
Page 19 of 29 The Journal of Physical Chemistry

1
2
3
This work was partially supported by NSF grants EECS-1436626 and DMR-1455050. We
4
5
6 thank Ludwig Bartels and his research team for helpful discussions. We would like to thank
7
8 Alexander Duerloo, Yao Li and Qian Yang for useful comments.
9
10
11
12
13 References
14
15
16
(1) Novoselov, K. S.; Geim, A. K.; Morozov, S. V; Jiang, D.; Zhang, Y.; Dubonos, S. V;
17
18 Grigorieva, I. V; Firsov, A. A. Electric Field Effect in Atomically Thin Carbon Films.
19 Science 2004, 306, 666–669.
20
21 (2) Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S. Electronics and
22 Optoelectronics of Two-Dimensional Transition Metal Dichalcogenides. Nat.
23
24 Nanotechnol. 2012, 7, 699–712.
25
26 (3) Wilson, J. A.; Yoffe, A. D. The Transition Metal Dichalcogenides Discussion and
27 Interpretation of the Observed Optical, Electrical and Structural Properties. Adv. Phys.
28 1969, 18, 193–335.
29
30
31 (4) Duerloo, K.-A. N.; Li, Y.; Reed, E. J. Structural Phase Transitions in Two-Dimensional
32 Mo- and W-Dichalcogenide Monolayers. Nat. Commun. 2014, 5, 4214.
33
34 (5) Chou, S. S.; De, M.; Kim, J.; Byun, S.; Dykstra, C.; Yu, J.; Huang, J.; Dravid, V. P.
35
Ligand Conjugation of Chemically Exfoliated MoS2. J. Am. Chem. Soc. 2013, 135, 4584–
36
37 4587.
38
39 (6) Voiry, D.; Goswami, A.; Kappera, R.; Silva, C. de C. C. e; Kaplan, D.; Fujita, T.; Chen,
40 M.; Asefa, T.; Chhowalla, M. Covalent Functionalization of Monolayered Transition
41 Metal Dichalcogenides by Phase Engineering. Nat. Chem. 2014, 7, 45–49.
42
43
44 (7) Voiry, D.; Mohite, A.; Chhowalla, M. Phase Engineering of Transition Metal
45 Dichalcogenides. Chem. Soc. Rev. 2015, 44, 2702–2712.
46
47 (8) Kappera, R.; Voiry, D.; Yalcin, S. E.; Branch, B.; Gupta, G.; Mohite, A. D.; Chhowalla,
48
49
M. Phase-Engineered Low-Resistance Contacts for Ultrathin MoS2 Transistors. Nat.
50 Mater. 2014, 13, 1128–1134.
51
52 (9) Brown, B. E. The Crystal Structures of WTe2 and High-Temperature MoTe2. Acta
53 Crystallogr. 1966, 20, 268–274.
54
55
56 (10) Vellinga, M. B.; Jonge, R. De; Haas, C. Semiconductor to Metal Transition in MoTe2. J.
57 Solid State Chem. 1970, 302, 299–302.
58
59
60
ACS Paragon Plus Environment
19
The Journal of Physical Chemistry Page 20 of 29

1
2
3
(11) Ma, Y.; Dai, Y.; Guo, M.; Niu, C.; Lu, J.; Huang, B. Electronic and Magnetic Properties
4
5 of Perfect, Vacancy-Doped, and Nonmetal Adsorbed MoSe2, MoTe2 and WS2
6 Monolayers. Phys. Chem. Chem. Phys. 2011, 13, 15546–15553.
7
8 (12) Kong, J.; Franklin, N. R.; Zhou, C.; Chapline, M. G.; Peng, S.; Cho, K.; Dai, H. Nanotube
9 Molecular Wires as Chemical Sensors. Science (80-. ). 2000, 287, 622–625.
10
11
12 (13) Balasubramanian, K.; Kern, K. 25Th Anniversary Article: Label-Free Electrical
13 Biodetection Using Carbon Nanostructures. Adv. Mater. 2014, 26, 1154–1175.
14
15 (14) Schedin, F.; Geim, A. K.; Morozov, S. V; Hill, E. W.; Blake, P.; Katsnelson, M. I.;
16
17
Novoselov, K. S. Detection of Individual Gas Molecules Adsorbed on Graphene. Nat.
18 Mater. 2007, 6, 652–655.
19
20 (15) Leenaerts, O.; Partoens, B.; Peeters, F. M. Adsorption of H2O, NH3, CO, NO2, and NO on
21 Graphene: A First-Principles Study. Phys. Rev. B 2008, 77, 125416.
22
23
24 (16) Wu, S.; Zeng, Z.; He, Q.; Wang, Z.; Wang, S. J.; Du, Y.; Yin, Z.; Sun, X.; Chen, W.;
25 Zhang, H. Electrochemically Reduced Single-Layer MoS2 Nanosheets: Characterization,
26 Properties, and Sensing Applications. Small 2012, 8, 2264–2270.
27
28 (17) Tongay, S.; Zhou, J.; Ataca, C.; Liu, J.; Kang, J. S.; Matthews, T. S.; You, L.; Li, J.;
29
30 Grossman, J. C.; Wu, J. Broad-Range Modulation of Light Emission in Two-Dimensional
31 Semiconductors by Molecular Physisorption Gating. Nano Lett. 2013, 13, 2831–2836.
32
33 (18) Ou, J. Z.; Chrimes, A. F.; Wang, Y.; Tang, S.; Strano, M. S.; Kalantar-zadeh, K. Ion-
34 Driven Photoluminescence Modulation of Quasi-Two- Dimensional MoS2 Nanoflakes for
35
36
Applications in Biological Systems. Nano Lett. 2014.
37
38 (19) Late, D. J.; Huang, Y.; Liu, B.; Acharya, J.; Shirodkar, S. N.; Luo, J.; Yan, A.; Charles,
39 D.; Waghmare, U. V; Dravid, V. P.; et al. Sensing Behavior of Atomically Thin-Layered
40 MoS2 Transistors. ACS Nano 2013, 7, 4879–4891.
41
42
43 (20) Perkins, F. K.; Friedman, A. L.; Cobas, E.; Campbell, P. M.; Jernigan, G. G.; Jonker, B. T.
44 Chemical Vapor Sensing with Monolayer MoS2. Nano Lett. 2013, 13, 668–673.
45
46 (21) Eda, G.; Fujita, T.; Yamaguchi, H.; Voiry, D.; Chen, M.; Chhowalla, M. Coherent Atomic
47 and Electronic Heterostructures of Single-Layer MoS2. ACS Nano 2012, 6, 7311–7317.
48
49
50 (22) Yang, D.; Jiménez Sandoval, S.; Divigalpitiya, W. M. R.; Irwin, J. C.; Frindt, R. F.
51 Structure of Single-Molecular-Layer MoS2. Phys. Rev. B 1991, 43, 12053–12056.
52
53 (23) Gordon, R. A.; Yang, D.; Crozier, E. D.; Jiang, D. T.; Frindt, R. F. Structures of
54
55
Exfoliated Single Layers of WS2, MoS2, and MoSe2 in Aqueous Suspension. Phys. Rev. B
56 2002, 65, 125407.
57
58
59
60
ACS Paragon Plus Environment
20
Page 21 of 29 The Journal of Physical Chemistry

1
2
3
(24) Enyashin, A. N.; Seifert, G. Density-Functional Study of LixMoS2 Intercalates (0⩽x⩽1).
4
5 Comput. Theor. Chem. 2012, 999, 13–20.
6
7 (25) Pradhan, N. R.; Rhodes, D.; Feng, S.; Xin, Y.; Memaran, S.; Moon, B.; Terrones, H.;
8 Terrones, M.; Balicas, L. Field-Effect Transistors Based on Few-Layered a-MoTe2. 2014,
9 5911–5920.
10
11
12 (26) Ruppert, C.; Aslan, O. B.; Heinz, T. F. Optical Properties and Band Gap of Single- and
13 Few-Layer MoTe2 Crystals. Nano Lett. 2014, 14, 6231–6236.
14
15 (27) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy
16
17
Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186.
18
19 (28) Blochl, P. E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953-17979
20
21 (29) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-
22
Wave Method. Phys. Rev. B 1999, 59, 1758–1775.
23
24
25 (30) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made
26 Simple. Phys. Rev. Lett. 1996, 77, 3865–3868.
27
28
(31) Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D. C.; Lundqvist, B. I. Van Der Waals
29
30 Density Functional for General Geometries. Phys. Rev. Lett. 2004, 92, 246401.
31
32 (32) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate Ab Initio
33 Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94
34 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104.
35
36
37 (33) Klime, J.; Bowler, D. R.; Michaelides, A. Van Der Waals Density Functionals Applied to
38 Solids. Phys. Rev. B 2011, 83, 1–13.
39
40 (34) Tkatchenko, A.; DiStasio, R. A.; Car, R.; Scheffler, M. Accurate and Efficient Method for
41
42
Many-Body van Der Waals Interactions. Phys. Rev. Lett. 2012, 108, 1–5.
43
44 (35) Zhu, Z. Y.; Cheng, Y. C.; Schwingenschlögl, U. Giant Spin-Orbit-Induced Spin Splitting
45 in Two-Dimensional Transition-Metal Dichalcogenide Semiconductors. Phys. Rev. B
46 2011, 84, 153402.
47
48
49 (36) Park, J.; Yu, B. D.; Hong, S. Van Der Waals Density Functional Theory Study for Bulk
50 Solids with BCC, FCC, and Diamond Structures. Curr. Appl. Phys. 2015.
51
52 (37) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B
53 1976, 13, 5188–5192.
54
55
56 (38) Bengtsson, L. Dipole Correction for Surface Supercell Calculations. Phys. Rev. B 1999,
57 59, 12301–12304.
58
59
60
ACS Paragon Plus Environment
21
The Journal of Physical Chemistry Page 22 of 29

1
2
3
(39) Makov, G.; Payne, M. Periodic Boundary Conditions in Ab Initio Calculations. Phys. Rev.
4
5 B 1995, 51, 4014–4022.
6
7 (40) Neugebauer, J.; Scheffler, M. Adsorbate-Substrate and Adsorbate-Adsorbate Interactions
8 of Na and K Adlayers on Al(111). Phys. Rev. B 1992, 46, 16067–16080.
9
10
11 (41) Duerloo, K.-A. N.; Ong, M. T.; Reed, E. J. Intrinsic Piezoelectricity in Two-Dimensional
12 Materials. J. Phys. Chem. Lett. 2012, 3, 2871–2876.
13
14 (42) Li, Y.; Duerloo, K.-A. N.; Reed, E. J. Strain Engineering in Monolayer Materials Using
15 Patterned Adatom Adsorption. Nano Lett. 2014, 14, 4299–4305.
16
17
18 (43) Liu, Z.; Liu, J. Z.; Cheng, Y.; Li, Z.; Wang, L.; Zheng, Q. Interlayer Binding Energy of
19 Graphite: A Mesoscopic Determination from Deformation. Phys. Rev. B 2012, 85, 1–5.
20
21 (44) Gong, Y.; Liu, Z.; Lupini, A. R.; Shi, G.; Lin, J.; Najmaei, S.; Lin, Z.; Elías, A. L.;
22
Berkdemir, A.; You, G.; et al. Band Gap Engineering and Layer-by-Layer Mapping of
23
24 Selenium-Doped Molybdenum Disulfide. Nano Lett. 2014, 14, 442–449.
25
26 (45) Mann, J.; Ma, Q.; Odenthal, P. M.; Isarraraz, M.; Le, D.; Preciado, E.; Barroso, D.;
27 Yamaguchi, K.; von Son Palacio, G.; Nguyen, A.; et al. 2-Dimensional Transition Metal
28 Dichalcogenides with Tunable Direct Band Gaps: MoS2(1-X) Se2x Monolayers. Adv. Mater.
29
30 2014, 26, 1399–1404.
31
32 (46) Li, H.; Duan, X.; Wu, X.; Zhuang, X.; Zhou, H.; Zhang, Q.; Zhu, X.; Hu, W.; Ren, P.;
33 Guo, P.; et al. Growth of Alloy MoS2xSe2(1-X) Nanosheets with Fully Tunable Chemical
34 Compositions and Optical Properties. J. Am. Chem. Soc. 2014, 136, 3756–3759.
35
36
37 (47) Chen, Y.; Xi, J.; Dumcenco, D. O.; Liu, Z.; Suenaga, K.; Wang, D.; Shuai, Z.; Huang, Y.-
38 S.; Xie, L. Tunable Band Gap Photoluminescence from Atomically Thin Transition-Metal
39 Dichalcogenide Alloys. ACS Nano 2013, 7, 4610–4616.
40
41
42
43
44 Table of Contents Image
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
22
Page 23 of 29 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
ACS Paragon Plus Environment
23
The Journal of Physical Chemistry Page 24 of 29

1
2
3
4
5
6
7

top view
8
9
10
11
12
13
14
15
16
17
18 side view
19
20
21
22
23
24
25
26
27 3D view
28
29
30
31

2H 1T 1T'
32
33
34
35
36

Te Mo
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

ACS Paragon Plus Environment


Page 25 of 29 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
0.15
9
10
11 0.1
Formation Energy (eV)

12
13
14
15
0.05
16
17
18 0
19
20
21
22
−0.05
23
24
25
26
−0.1
27
28
29 −0.15 2MoTe2(2H) + O2 → 2MoTe2(O2)0.5(2H)
30
31
32 −0.2 2MoTe2(2H) + O2 → 2MoTe2(O2)0.5(1T’)
33
34
35
36 −0.25
37 spin spin-orbit
38 DFT-D3 vdW-DF TS-SCS
39
40
polarized coupling
41
42
43
44
45
46 ACS Paragon Plus Environment
47
48
The Journal of Physical Chemistry Page 26 of 29
(a)
1
2
3
4
5
6
7
8
9
10
11
12 Li Na K O Cl H F H2
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28 H2O NH3 NO NO2 CO CO2 N2 O2
29
30
(b)
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
Li Na K O Cl H F H2
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
H2O NH3 NO NO2 CO CO2 N2 O2

Te Li K Cl F C
O Environment
Na Plus
MoACS Paragon N H
(a)
Page 27 of 29 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Li Na K O Cl H2 H2O NH3 NO NO2 CO CO2 N2 O2 no X
27
28
29
(b)
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 28 of 29

1
2
3
4
5
6
7
8
0
9
10
11
−0.1
Formation Energy (eV)

12
13
14
15
16
17 −0.2
18
19
20
21
22
−0.3
23
24
25
26 −0.4
27 2MoxW1−xTe2(2H) + X → 2MoxW1−xTe2X0.5(2H)
28
29
8MoxW1−xTe2(2H) + X → 8MoxW1−xTe2X0.125(2H)
30
31
−0.5
32
33
2MoxW1−xTe2(2H) + X → 2MoxW1−xTe2X0.5(1T’)
34
35 −0.6 8MoxW1−xTe2(2H) + X → 8MoxW1−xTe2X0.125(1T’)
36
37
38
39 H2 H2O NH3 NO NO2 CO CO2 N2 O2
40
41
42
43
44
45
46 ACS Paragon Plus Environment
47
48
Page 29 of 29 The Journal of Physical Chemistry

1
2
3
4 b
5
6
7
8 c a
9
10
11
12
13
14
15
16
17
18
19
20
2H
21
22
23
24
25
26
27
28
29
30
31 b
32
33
34
35 c a
36
37
38
1T’
39
40
41
42
43
44
45
46 ACS Paragon Plus Environment
47
48

You might also like