Supporting information for Electron dynamics of solvated Titanium

advertisement
1
Supporting information for
2
3
Electron dynamics of solvated Titanium Hydroxide
4
Dayton J. Vogel, Dmitri S. Kilin*
5
University of South Dakota, Department of Chemistry, Vermillion, SD, 57069
6
*Corresponding author email: Dmitri.Kilin@usd.edu
7
8
The electronic nature was calculated using density functional theory (DFT) run
9
within the Vienna ab initio simulation package (VASP)[1] using the Perdew-Burke-
10
Ernzerhof (PBE)[2] form of the generalized gradient approximation (GGA). The
11
electronic nature was implemented with a plane wave basis set and projected
12
augmented wave (PAW)[3] potentials. DFT calculations provide data for the
13
optimized geometry, ground state electronic structure, absorption spectra, and
14
molecular dynamics of the model using a Kohn-Sham (KS) basis set.
15
16
Optical transitions are induced an oscillating electromagnetic field propagating a
17
system from a ground state to an excited state. Such transitions are determined by
18
matrix elements of the leading term in the electron-photon interaction
19
20
𝑯𝒊𝒏𝒕 =
⃗ )𝟐
(𝒑+𝒆𝑨
𝟐𝒎
𝒑𝟐
= 𝟐𝒎 +
⃗
⃗ 𝒆𝑨
𝒑
𝒎
+
𝐞𝟐 ⃗𝑨𝟐
𝟐𝒎
(S1)
Where 𝑝 is the momentum operator, e is electric charge, 𝐴 is the vector potential.[4]
21
S-1
22
⃗ 𝑖𝑗 are calculated from Kohn-Sham orbitals (KSO) as
The transition dipole moments 𝐷
23
detailed previously.[5] The KSO are eigenfunctions of the KS effective Hamiltonian
24
FKS with energies 𝜀𝑗 , given at position r by:
25
⟨𝒓|𝒋⟩ = 𝝓𝒋 (𝒓) = ∑𝑮<𝑮𝒄𝒖𝒕 𝑪𝒋,𝑮⃗ 𝒆−𝒊𝑮⃗ 𝒓
26
where 𝐶𝑗,𝐺 is a KSO in the momentum representation, G is a grid point in reciprocal
27
space and Gcut is the cutoff value for the magnitude of G vectors in a Fourier
28
expansion.[6] The transition dipole moment, D jk , for the jk transition is given by:
29
30
̂|𝒌⟩ =
𝑫𝒋𝒌 = ⟨𝒋|𝑫
∑𝑮<𝑮𝒄𝒖𝒕 𝑪∗𝒋,𝑮 𝑫𝑮,𝑮′ 𝑪
𝒌,𝑮′
(S2)
(S3)
Where
′
31
𝑫𝑮,𝑮′ = −𝒆 ∫ 𝒅𝟑 𝒓𝒆𝒊𝑮𝒓 𝒓𝒆−𝒊𝑮 𝒓
(S4)
32
The dipole matrix elements have been re-expressed in terms of matrix elements of
33
the momentum operator:
34
35
⟨𝒋|𝒑̂|𝒌⟩ = 𝒊ℏ ∑𝑮,𝑮′ <𝑮𝒄𝒖𝒕 𝑪∗𝒋,𝑮 𝑮𝑪
𝒌,𝑮′
(S5)
to obtain
36
37
38
⟨𝒋|𝑫̂|𝒌⟩ = ⟨𝒋|𝒑̂|𝒌⟩𝒊ℏ𝒆/𝒎𝒆 (𝜺𝒌 − 𝜺𝒋 )
(S6)
for calculations.[5]
39
40
The gamma point, Γ, was chosen to represent the k-point sampling based on the
41
substantial energy gap of the model. Such an approximation is justified for wide-gap
42
insulators and periodic models in vacuum. As the bandgap is larger than 4eV for this
43
⃗ ) in TiO2
model it is considered a wide-gap insulator. Dispersion curves of 𝜀𝑖 (𝑘
S-2
44
nanowires has been found to be “flat.”[7] Dispersion corrections implemented by
45
Grimme are being considered in future calculations.[8]
46
47
Computational details include: the energy cutoff value was chosen to be 300eV, a
48
MD trajectory length of 1ps to represent typical nuclear motion, and a time step of
49
1fs was chosen for the Ab initio MD. Standard procedure from VASP was used to
50
integrate nuclear motion trajectories. Equilibration was reached by modeling
51
nuclear interaction with a thermostat at various ambient temperatures according to
52
Eq. 6. The temperatures simulated were 100K, 200K, 300K, and 350K.
53
54
Computational details include: the energy cutoff value was chosen to be
55
300eV, a MD trajectory length of 1ps to represent typical nuclear motion, and a time
56
step of 1fs was chosen for the Ab initio MD. Standard procedure from VASP was
57
used to integrate nuclear motion trajectories. Equilibration was reached by
58
modeling nuclear interaction with a thermostat at various ambient temperatures
59
according to Eq. 6. The temperatures simulated were 100K, 200K, 300K, and 350K.
60
61
Decoherence can be important for very quick initial time, t<10 fs. However,
62
after 10fs decoherence is of secondary importance. The equation of motion (EOM)
63
for 𝜌ij splits on two independent subsets where populations i=j and 𝑖 ≠ 𝑗.
64
Decoherence is important for 𝑖 ≠ 𝑗, but we focus on i=j. The decoherence effects are
65
important factors and have been implemented into recent surface hopping
66
procedures.[9] Correcting for decoherence is expected to slightly speed up the
S-3
67
relaxation. The transitions with smaller subgap energies will be enhanced to a
68
greater measure, thus obeying the energy gap law.
69
Excitation States
Oscillator
Excitation
Initial
Final
Strength
Energy
HO-23
LU+2
1.30866221
5.4454
HO-22
LU+2
1.04703001
5.3963
HO-8
LU+2
0.95182345
6.7146
HO-19
LU+2
0.62432678
5.5258
HO-40
134
0.60306177
6.8978
HO-26
LU+2
0.59854254
5.3011
70
71
Table S1. The six highest oscillator strength values are given. Initial and final states for
72
each transitions, corresponding to each oscillator strength, are given as well as the energy
73
needed to generate the excitation.
74
75
S-4
76
Figure S 1. An insert highlighting the ligand field splitting (close to tetrahedral) found between
77
0eV and 2eV of the DOS in Figure 1 C. The non-shaded area under the curve represents the
78
electronic states are unoccupied. The vertical lines represent individual state energies. Solid
79
lines represent states of Ti 3d character while the dashed line represents a state of oxygen
80
character.
81
82
83
Figure S2. Partial density of states for KSO range of HO to LU+5. Orbital character
84
shows Ti 3d nature to LU, LU+1, LU+3, LU+4, LU+5. LU+2 state shows that the main
85
orbital character comes from s orbitals. The left panel shows the data points while the
86
right panel visualizes the trend between KSOs orbital character.
87
88
The Redfield tensor, Rijlm, expressed in Eq. 12 is comprised of tensor components
89
according to
Rijkl = G+ljik + G-ljik - dlj å G+immk - dik å G-lmmj
90
91
m
m
(S7)
where
92
G+ijkl = ò dt M ijkl (t ) e-iwklt
(S8)
93
G-ijkl = ò dt M ijkl (t ) e
(S9)
S-5
-iwijt
94
The Redfield tensor is derived and reported in previous work, according to A.
95
Redfield.[10]
A
B
C
D
96
97
Figure S3. The Redfield Tensor is displayed in panels A, B, C, and D for ambient
98
temperature simulations at 100K, 200K, 300K, and 350K, respectively. The x and y
99
axis are given for orbital indexes j and i, respectively. The z component of the graph
100
gives the amplitude of Rijij fs-1.
101
Electron Relaxation Time (ps)
Ti(OH)4
2.479
<001> NW
2.812
<001> Surface
0.216
Hole Relaxation Time (ps)
Ti(OH)4
0.237
<001> NW
1.181
102
S-6
<001> Surface
0.307
103
Table S2. Charge carrier relaxation times for both electron and hole. Relaxation
104
times for the solvated Ti(OH)4, <001> nanowire (NW), and <001> anatase surface
105
are given in picoseconds.
106
107
Computation of the Ti(OH)4 gas phase molecule were conducted using time
108
dependent density functional theory (TDDFT) to use in comparison to the presented
109
DFT results.
110
111
Figure S4. Absorption spectra of Ti(OH)4 gas phase molecule computed with
112
TDDFT, with B3LYP hybrid functional and LanL2Dz basis.
113
Orbitals contributing to
excitation
Initial
Final
24
25
Excitation Energy (eV)
3.319
23
25
3.6866
21
25
3.9426
S-7
22
25
114
21
25
4.0589
22
25
23
26
Table S3. Excitation Energies calculated using TDDFT for the Ti(OH)4 gas phase
115
molecule.
116
117
The excitation energies calculated using TDDFT are comparable to the band gap
118
energies calculated using DFT. The excitation energies calculated using TDDFT are
119
expected to be smaller than the band gap energies found using DFT. TDDFT takes
120
electron and hole interaction into account when calculating the energies of possible
121
excitations.
122
123
References
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
1.
2.
3.
4.
5.
6.
7.
8.
Kresse, G. and J. Furthmuller, Efficient iterative schemes for ab initio totalenergy calculations using a plane-wave basis set. Phys Rev B Condens Matter,
1996. 54(16): p. 11169-11186.
Perdew, J.P., K. Burke, and M. Ernzerhof, Generalized Gradient Approximation
Made Simple. Physical Review Letters, 1996. 77(18): p. 3865-3868.
Blochl, P.E., Projector augmented-wave method. Phys Rev B Condens Matter,
1994. 50(24): p. 17953-17979.
Gerry, C.C. and P. Knight, Introductory quantum optics. 2005, Cambridge, UK ;
New York: Cambridge University Press. xiii, 317 p.
Vazhappilly, T., D.S. Kilin, and D.A. Micha, Photoabsorbance and Photovoltage
of Crystalline and Amorphous Silicon Slabs with Silver Adsorbates. Journal of
Physical Chemistry C, 2012. 116(48): p. 25525-25536.
Martin, R.M., Electronic structure : basic theory and practical methods. 1st pbk.
ed. 2008, Cambridge, UK ; New York: Cambridge University Press. xxii, 624 p.
Huang, S. and D.S. Kilin, Electronic structure and hot carrier relaxation in
⟨001⟩ anatase TiO2 nanowire. Molecular Physics, 2013: p. 1-7.
Grimme, S., Accurate description of van der Waals complexes by density
functional theory including empirical corrections. J Comput Chem, 2004.
25(12): p. 1463-73.
S-8
143
144
145
146
147
9.
10.
Jaeger, H.M., S. Fischer, and O.V. Prezhdo, Decoherence-induced surface
hopping. The Journal of Chemical Physics, 2012. 137(22): p. -.
Redfield, A.G., On the Theory of Relaxation Processes. IBM Journal of Research
and Development, 1957. 1: p. 19-31.
S-9
Download