Chih-Kai
Lin
*ab,
Jun
Li
c,
Zheyan
Tu
d,
Xiangyuan
Li
c,
Michitoshi
Hayashi
e and
Sheng Hsien
Lin
ab
aInstitute of Atomic and Molecular Sciences, Academia Sinica, Taipei, Taiwan 10617, ROC. E-mail: ethene@gate.sinica.edu.tw
bDepartment of Applied Chemistry, National Chiao Tung University, Hsinchu, Taiwan 30010, ROC
cCollege of Chemical Engineering, Sichuan University, Chengdu, 61005, P. R. China. E-mail: xyli@scu.edu.cn
dCollege of Chemistry, Sichuan University, Chengdu, 61005, P. R. China. E-mail: tuzheyan@gmail.com
eCenter for Condensed Matter Sciences, National Taiwan University, Taipei, Taiwan 10617, ROC
First published on 15th September 2011
In this work geometric optimization and potential energy surface (PES) scan for the TiO2 molecule were carried out to find out possible stable structures by using quantum chemical calculations. The ground state (S0 1 1A1) has the only equilibrium with a symmetric bent structure, whereas the linear conformer is unstable. The first singlet excited state (S1 1 1B2), on the other hand, possesses two potential minima where one is bent around the Franck–Condon region and the other is linear. The second singlet excited state (S2 1 1A2) has only one minimum which is linear, and at which geometry the first two excited states degenerate. Other low-lying singlet and triplet excited states have been investigated accordingly. Among the functionals applied in the present DFT calculations, B3LYP gave results most close to available experimental data, while some recently developed ones including HSE06, ωB97X-D and CAM-B3LYP were not satisfactory for this small system. Ab initio methods such as CASSCF, CASPT2 as well as CCSD-related calculations have been applied, and the latter two showed good results similar to DFT/B3LYP.
It has been suggested that the TiO2 molecule has a bent structure in the ground state according to the measured permanent dipole moment in a very early study.1 The infrared spectrum of the molecule was then recorded in a neon matrix, indicating two vibrational motions, the symmetric stretching mode (ν1, of the a1 irreducible representation in the C2ν point group) at 962.0 cm−1 and the asymmetric stretching mode (ν3, b2) at 934.8 cm−1.2 Later the two modes were revised to 946.9 and 917.1 cm−1, respectively.3 However, the low-frequency bending mode (ν2, a1) in the 300 cm−1 range was not observed until very recently.4 The Ti–O bond length (rTiO) and the O–Ti–O bending angle (θOTiO) have been determined as 1.651 Å and 111.6°, respectively, in the argon matrix.5 Theoretical investigations by using density functional theory (DFT), complete active space self-consistent field (CASSCF) and coupled cluster (CC) calculations have all shown consistent results with the experimental ones for the ground state.6–11
The situation becomes much more complicated for the excited states. Garkusha et al. recorded the electronic absorption spectrum of TiO2 in a 6 K neon matrix by using halogen and xenon arc lamps, presenting the band origin at 19084 cm−1 (524 nm, 2.37 eV) for the S1 1 1B2 ← S0 1 1A1 transition and the progressions of all three modes of the excited state. In addition, they noticed another progression beginning at 15
924 cm−1 (628 nm, 1.97 eV) which was assigned to the linear isomer in its 1 1Σ+g ground state.12 Steimle and Maier's group performed resonance-enhanced multiphoton ionization (REMPI) and laser induced fluorescence (LIF) spectra of the gas-phase molecule, showing that the band origin slightly shifted to 18
655 cm−1 (536 nm, 2.31 eV) but the vibronic progressions were too complicated to recognize at first glance.4 The newest report by the same group has re-located the 0–0 transition to 17
591 cm−1 (568 nm, 2.18 eV) and assigned several tens of vibronic peaks.13 There are as yet many minor peaks remaining unassigned, implying that some other low-lying states or conformers might also contribute to the absorption spectrum through possible vibronic and/or spin–orbit couplings.
Theoretical characterizations on the excited states of TiO2 are quite rare. There are only two computational reports concerning the first singlet excited state (S1 1 1B2) till now. Ramana and Phillips studied the state at the configuration interaction (CI) level, predicting that it lies just 1.63 eV above the ground state, which is much lower than the experimental value, and the equilibrium structure is linear with the bond length 1.73 Å.14 Grein performed DFT optimization and obtained a bent equilibrium structure with the bond length and bending angle as 1.703 Å and 96.3°, respectively.9 The adiabatic transition energy to this state was calculated as 2.14 eV, which is close to the most recent experimental report.13
For the second singlet excited state (S2 1 1A2), two DFT results with different functionals indicated that the bent structure is stable. The bond lengths were calculated both as 1.72 Å and the bending angles as 143.9° and 153.7°, respectively.9,10 The adiabatic excitation energies for this state were reported as 2.50 eV and 1.85 eV, respectively, where the latter was apparently underestimated. On the other hand, a brief potential energy scan depending on the bending angle showed a very wide and shallow potential basin for this state,9 implying that the structure shall no longer be a harmonic oscillator and might even be unstable.
In this work, we aimed to identify the possible stable structures of the ground and low-lying excited states. The geometric and electronic properties of TiO2 monomer would be unveiled with the full optimization and a thorough scan on the potential energy surfaces at the DFT, CAS and CC levels, providing us more insights into the complicated spectroscopic problems.
![]() | ||
Fig. 1 Three categories of TiO2 mono-molecular isomers: (a) symmetric bent form, (b) cyclic form and (c) end-on form. |
For the optimization and vibrational frequency calculation of the ground state, S0 1 1A1, CASSCF, CASPT2 and CCSD(T) have been carried out. The augmented correlation-consistent polarized valence triple-zeta (aug-cc-pVTZ) basis set was generally applied. The Los Alamos effective core potentials (ECPs) plus MBS or DZ (that is, LanL2MB or LanL2DZ)16 of the Ti atom were also adopted in some tests. For clarity only aug-cc-pVTZ results are presented here, while the comparison between different basis sets is listed in the ESI.† In CASSCF calculations, the choice of the largest active space included all 16 valence electrons and 14 orbitals, forming the (16,14) active space. A smaller one excluded the inner orbitals mainly composed of the O2s atomic orbitals, giving the (12,12) active space. The global minimum on the potential energy surface was searched according to previously reported geometric data, and other possible local minima were considered as well.
In addition to ab initio methods mentioned above, DFT with several different functionals were applied. Besides the commonly accepted B3LYP functional, some recently developed functionals that attracted much interest have also been tested including HSE06,17,18 which takes screened short-range Coulomb potential and has been applied in TiO2 crystal systems, and CAM-B3LYP19 and ωB97X-D,20 which add in long-range corrections. The results of these new functionals were compared to verify if they could give proper or even better descriptions.
For the first and second singlet excited states, S1 1 1B2 and S2 1 1A2, time-dependent DFT (TD-DFT), CASSCF, CASPT2 and equation-of-motion CCSD (EOM-CCSD) were applied to localize the minimum and to calculate vibrational modes. The search for equilibrium structures was first around the Franck–Condon (vertical excitation) region and then expanded to other conformers.
In order to understand the characters of the low-lying excited states, a systematic scan on the potential energy surfaces was carried out by changing the geometry of the molecule. Keeping the point group symmetry as C2ν, the Ti–O bond length and the O–Ti–O bending angle were varied from 1.50 Å to 2.00 Å and from 80° to 180°, respectively. TD-DFT was applied to locate singlet states up to S5. CASSCF followed by CASPT2 was done in an eight-state-averaged manner, with two states of each irreducible representation (1A1, 1B1, 1B2 and 1A2) included.
In the last part of this work, triplet excited states such as T1 1 3B2 and T2 1 3A2 were also studied. Spin–orbit couplings were calculated to verify possible interactions between neighbor singlet and triplet states. Additionally, cyclic and end-on isomers were looked at briefly. All the quantum chemical calculations were performed with GAUSSIAN09,21MOLPRO200622 and ACESII23 computational packages.
In this work the optimization and vibrational frequency calculation of the S0 1 1A1 ground state were performed at levels including Hartree–Fock (HF), Møller–Plesset second-order perturbation (MP2), CASSCF, CASPT2, CCSD(T) and DFT with several different functionals. Referring to Table 1, the geometric data obtained by HF and MP2 significantly differed from the experimental reference as reported previously,6 while the other methods performed rather well with deviations within 0.01 Å (0.6%) for the bond length, 2° (2%) for the bending angle and 0.6 debye (9%) for the permanent dipole moment. The choice of basis set and the size of active space showed minor effects in these calculations (Table S1 in the ESI).
Level of theory | Equilibrium geometry | Vibrational frequencies | Dipole moment | |||
---|---|---|---|---|---|---|
r TiO /Å | θ OTiO /° | ν 1 (a1)/cm−1 | ν 2 (a1)/cm−1 | ν 3 (b2)/cm−1 | μ/debye | |
a Ref. 2. b Ref. 3. c Ref. 5. d Ref. 4. | ||||||
HF | 1.615 | 118.2 | 1133.9 | 335.1 | 1046.6 | 7.48 |
MP2 | 1.685 | 107.4 | 885.9 | 318.9 | 922.2 | 8.59 |
CASSCF(12,12) | 1.654 | 112.6 | 1005.8 | 329.6 | 945.1 | 6.92 |
CASSCF(16,14) | 1.648 | 113.2 | 1021.8 | 326.5 | 940.0 | 6.98 |
CASPT2(12,12) | 1.672 | 111.3 | 924.4 | 328.0 | 926.5 | 6.97 |
CCSD(T) | 1.652 | 111.2 | 988.5 | 333.7 | 955.2 | 6.79 |
B3LYP | 1.642 | 111.7 | 1023.0 | 341.9 | 975.2 | 6.73 |
HSE06 | 1.630 | 111.7 | 1046.3 | 347.9 | 992.8 | 6.77 |
ωB97X-D | 1.629 | 112.5 | 1055.1 | 349.7 | 1000.4 | 6.86 |
CAM-B3LYP | 1.629 | 112.2 | 1059.5 | 344.9 | 1004.7 | 6.84 |
Expt. | 110 ± 15a | 962.0a | 934.8a | |||
113 ± 5b | 946.9b | 917.1b | ||||
1.651c | 111.6c | 330 ± 6d | 6.33d |
The three vibrational modes, i.e. symmetric stretching (ν1), bending (ν2) and asymmetric stretching (ν3), have been reported from experimental studies. The calculated harmonic frequencies, also shown in Table 1, were slightly higher by a factor within 5%, which is reasonable. In a critical comparison, CCSD(T) presented the best performance on reproducing both geometric and vibrational properties, and CASSCF, CASPT2 as well as DFT (with all tested functionals) showed acceptable agreements in describing the equilibrium structure.
Furthermore, it was suggested that a symmetric linear structure would exist with the 1 1Σ+g ground state in the D∞h point group.12 The search for such a stable structure was carried out by fixing the bending angle at 180°, and the converged results were obtained with energies about 2 eV higher than the bent equilibrium. However, as shown in Table 2 (and Table S2), they all corresponded to a saddle point instead of a true minimum, giving imaginary frequencies to the bending mode. It indicated that there is no stable linear conformer in the ground state, which was implied in a previous energy scan.9 This point shall be demonstrated more clearly with a thorough potential energy surface scan in the next sections. It is noticed that while the gas-phase linear molecule is unstable, the possibility of stabilization of such a molecule in the inert gas matrix could not be entirely excluded.12,13
Level of theory | Adiabatic energya | Equilibrium geometry | Vibrational frequencies | |||
---|---|---|---|---|---|---|
ΔEad/eV | r TiO /Å | θ OTiO /° | ν 1 (σ+g)/cm−1 | ν 2 (πu)/cm−1 | ν 3 (σ−g)/cm−1 | |
a With respect to the ground state equilibrium calculated at the same level of theory; without zero-point energy (the same in the following tables). | ||||||
MP2 | 1.852 | 1.750 | 180.0 | 760.5 | (340i) | 783.5 |
CASSCF(12,12) | 1.507 | 1.722 | 180.0 | 881.6 | (340i) | 719.4 |
CASSCF(16,14) | 1.419 | 1.713 | 180.0 | — | — | — |
CASPT2(12,12) | 1.705 | 1.735 | 180.0 | — | — | — |
CCSD(T) | 1.689 | 1.716 | 180.0 | 866.8 | (284i) | 889.2 |
B3LYP | 1.834 | 1.707 | 180.0 | 896.7 | (380i) | 901.5 |
HSE06 | 1.871 | 1.698 | 180.0 | 915.2 | (382i) | 908.5 |
ωB97X-D | 1.874 | 1.696 | 180.0 | 929.0 | (392i) | 914.1 |
CAM-B3LYP | 1.821 | 1.695 | 180.0 | 932.9 | (380i) | 910.8 |
State | Excitationa | TD-B3LYP | TD-HSE06 | TD-ωB97X-D | TD-CAM-B3LYP | EOM-CCSD | CASSCF(16,14) | CASPT2(16,14) | MRCI a |
---|---|---|---|---|---|---|---|---|---|
a Major excitation configurations and MRCI data from ref. 9. b Experimental vertical excitation energy to this state is 2.46 eV according to ref. 2. | |||||||||
1 1B2b | 6b2–10a1 | 2.672 | 2.966 | 3.352 | 2.939 | 2.367 | 1.981 | 2.386 | 2.43 |
1 1A2 | 6b2–4b1 | 3.256 | 3.517 | 3.730 | 3.506 | 3.019 | 3.654 | 3.150 | 3.09 |
2 1B2 | 6b2–11a1 | 3.337 | 3.645 | 3.934 | 3.658 | 3.191 | 4.021 | 3.365 | 3.21 |
2 1A1 | 9a1–10,11a1 | 3.520 | 3.799 | 4.179 | 3.755 | 3.294 | 3.167 | 3.480 | 3.13 |
1 1B1 | 3b1–10,11a1 | 3.811 | 4.114 | 4.450 | 4.110 | 3.547 | 3.294 | 3.601 | 3.57 |
2 1B1 | 9a1–4b1 | 4.028 | 4.259 | 4.417 | 4.212 | 3.682 | 4.658 | 3.979 | 3.74 |
2 1A2 | 1a2–10,11a1 | 4.116 | 4.433 | 4.755 | 4.409 | 3.880 | 3.862 | 3.948 | 4.07 |
3 1A1 | 9a1–11a1 | 4.218 | 4.476 | 4.651 | 4.450 | 3.993 | 4.321 | 4.586 | 3.83 |
3 1B2 | 5b2–10,11a1 | 4.403 | 4.720 | 5.011 | 4.661 | 4.255 | 4.574 | 4.250 | 4.19 |
1 3B2 | 6b2–10a1 | 2.556 | 2.834 | 3.248 | 2.808 | 2.324 | 2.069 | 2.495 | 2.40 |
1 3A2 | 6b2–4b1 | 3.133 | 3.376 | 3.608 | 3.376 | 2.996 | 3.598 | 2.991 | 3.07 |
2 3B2 | 6b2–11a1 | 3.181 | 3.428 | 3.748 | 3.457 | 3.114 | 4.017 | 3.253 | 3.20 |
1 3A1 | 9a1–10, 11a1 | 3.226 | 3.435 | 3.651 | 3.322 | 3.072 | 3.160 | 3.425 | 3.12 |
1 3B1 | 3b1–10, 11a1 | 3.551 | 3.774 | 4.042 | 3.741 | 3.332 | 3.214 | 3.612 | 3.43 |
2 3A1 | 9a1–11a1 | 3.600 | 3.789 | 4.232 | 3.892 | 3.657 | 4.201 | 3.886 | 3.85 |
2 3B1 | 9a1–4b1 | 3.706 | 3.920 | 4.299 | 3.923 | 3.474 | 4.286 | 3.616 | 3.59 |
3 3A1 | 3b1–4b1 | 3.845 | 4.048 | 4.412 | 4.080 | 3.843 | 4.479 | 4.449 | 4.00 |
2 3A2 | 1a2–10, 11a1 | 3.885 | 4.148 | 4.508 | 4.107 | 3.696 | 3.665 | 3.932 | 3.81 |
3 3B2 | 5b2–10, 11a1 | 4.130 | 4.338 | 4.551 | 4.258 | 4.044 | 4.720 | 4.122 | 4.02 |
According to more reliable methods like EOM-CCSD and CASPT2, the lowest two singlet excited states are unquestionably S1 1 1B2 and S2 1 1A2, and the triplet ones are T1 1 3B2 and T2 1 3A2. The sequence of following excited states slightly differs by levels of theory. Nevertheless the difference is insignificant since the energy gaps between neighboring states are quite small, i.e. within 0.2 eV.
![]() | ||
Fig. 2 TD-B3LYP/aug-cc-pVTZ potential energy curves of low-lying (a) singlet and (b) triplet states of TiO2. The relative energy is with respect to the equilibrium of the 1 1A1 ground state. |
Level of theory | Vertical energya | Adiabatic energya | Equilibrium geometry | Vibrational frequencies | Dipole moment | |||
---|---|---|---|---|---|---|---|---|
ΔEFC/eV | ΔEad/eV | r TiO /Å | θ OTiO /° | ν 1 (a1)/cm−1 | ν 2 (a1)/cm−1 | ν 3 (b2)/cm−1 | μ/debye | |
a With respect to the ground state equilibrium calculated at the same level of theory. b Ref. 9, with the 6-311++G(3df) basis set. c Ref. 2. d Ref. 12. e Ref. 4. f Ref. 13. | ||||||||
CASSCF(12,12) | 2.223 | 2.036 | 1.703 | 97.7 | 890.6 | 228.3 | (555i) | 3.14 |
CASSCF(16,14) | 2.168 | 1.918 | 1.700 | 95.6 | 908.9 | 241.3 | (323i) | 2.68 |
CASPT2(12,12) | 2.454 | 2.139 | 1.727 | 93.7 | — | — | — | 3.92 |
EOM-CCSD | 2.370 | 2.324 | 1.662 | 100.4 | 1009.1 | 198.3 | 499.0 | 4.31 |
TD-B3LYP | 2.672 | 2.579 | 1.673 | 100.2 | 945.6 | 212.5 | 374.7 | 3.57 |
TD-HSE06 | 2.966 | 2.868 | 1.662 | 99.5 | 966.9 | 209.6 | 390.3 | 3.58 |
TD-ωB97X-D | 3.352 | 3.278 | 1.670 | 111.7 | 920.7 | 115.9 | 339.7 | 5.19 |
TD-CAM-B3LYP | 2.939 | 2.874 | 1.646 | 110.0 | 10918.8 | (494i) | 10494.3 | 15.24 |
BPW91 b | — | 2.14 | 1.703 | 96.3 | 875 | 196 | 480 | 5.07 |
Expt. | 2.46c | 2.37d | 836d | 201d | 498d | |||
2.31e | 1.704e | 100.1e | 2.55e | |||||
2.18f | 876f | 184f | 316f |
In the case of TD-DFT calculations, B3LYP and HSE06 functionals achieved similar geometry although the latter somewhat overestimated the excitation energy. ωB97X-D gave apparent overestimates in energies, and CAM-B3LYP failed to verify this equilibrium. As for ab initio methods, the major problems are that CASSCF always converged with an imaginary-frequency vibrational mode for this state (and for all following excited states investigated in this work), and that CASPT2 could not afford frequency calculation. Despite these failures, the TD-DFT and EOM-CCSD data as well as experimental reports are enough to convince us of the existence and stability of such an equilibrium. Moreover, it is noticed that CASSCF (incidentally) well reproduced the permanent dipole moment which was measured as 2.55 debye4 while TD-B3LYP, TD-HSE06 and CASPT2 gave overestimates of about 50%, although this had only a minor influence on the energetic and geometric results.
In addition to the bent equilibrium, other possible stable conformers of this state have been searched. A second minimum was found with a symmetric linear structure at all levels of theory. To our surprise, this conformer is nearly isoenergetic with the bent one, showing that the “linearization energy” is almost zero or even negative as listed in Table 5, which is much smaller than previously estimated 0.42 eV.9 Within the computational error, the linear form might be even more stable than the bent form. This implies that the linear conformer of the first singlet excited state, rather than that of the ground state, could be a candidate that contributes to the minor peaks as well as to the red of band origin on the S1 1 1B2 ← S0 1 1A1absorption spectrum.
Level of theory | Adiabatic energya | Linearization energyb | Equilibrium geometry | Vibrational frequencies | |||
---|---|---|---|---|---|---|---|
ΔEad/eV | ΔElin/eV | r TiO /Å | θ OTiO /° | ν 1 (σ+g)/cm−1 | ν 2 (πu)/cm−1 | ν 3 (σ−g)/cm−1 | |
a With respect to the ground state equilibrium calculated at the same level of theory. b With respect to the local minimum around the Franck–Condon region of the same electronic state surface. c Ref. 9, with the 6-311++G(3df) basis set. | |||||||
CASSCF(12,12) | 2.026 | −0.010 | 1.735 | 180.0 | 782.9 | (59i) | 173.4 |
CASSCF(16,14) | 2.039 | 0.121 | 1.735 | 180.0 | 782.7 | (61i) | 292.6 |
CASPT2(12,12) | 2.153 | 0.014 | 1.749 | 180.0 | — | — | — |
EOM-CCSD | 2.387 | 0.063 | 1.684 | 180.0 | 887.9 | 297.3 | 879.4 |
TD-B3LYP | 2.609 | 0.030 | 1.700 | 180.0 | 837.2 | 253.8 | 658.8 |
TD-HSE06 | 2.865 | −0.003 | 1.688 | 180.0 | 853.2 | 215.8 | 701.6 |
TD-ωB97X-D | 2.976 | −0.302 | 1.684 | 180.0 | 877.3 | 225.7 | 789.7 |
TD-CAM-B3LYP | 2.812 | −0.062 | 1.682 | 180.0 | 873.2 | 233.9 | 820.7 |
BPW91 c | 2.56 | 0.42 | 1.728 | 180.0 | — | — | — |
In our work, we could not find any stable bent conformer at any TD-DFT level. The optimization process in fact converged to the linear form, at which geometry the S2 1 1A2 and S1 1 1B2 states become degenerate and correlate to the 1 1Δu state of the D∞h point group. The only stable form found for the S2 1 1A2 state was therefore essentially the same as the linear structure of the S1 1 1B2 state. EOM-CCSD gave the same consequence as TD-DFT which are summarized in Table 6. It is noticed that these outcomes are slightly different from Table 5, originating from computational errors under different initial symmetry settings (C2νvs. D∞h). CASSCF and CASPT2, on the other hand, indicated a non-linear equilibrium point with the bond angle of ∼103°. Unfortunately the calculated vibrational frequency of the asymmetric stretching mode was imaginary from CASSCF and unavailable from CASPT2, hence the existence of a stable bent conformer of this electronic state remains questionable.
Level of theory | Vertical energya | Adiabatic energya | Equilibrium geometry | Vibrational frequenciesb | Dipole moment | |||
---|---|---|---|---|---|---|---|---|
ΔEFC/eV | ΔEad/eV | r TiO /Å | θ OTiO /° | ν 1 (a1)/cm−1 | ν 2 (a1)/cm−1 | ν 3 (b2)/cm−1 | μ/debye | |
a With respect to the ground state equilibrium calculated at the same level of theory. b In the case of linear structure, the three vibrational modes have the symmetry of σ+g, πu and σ−g, respectively. c Degenerate with S1 1 1B2, both belonging to the doubly degenerate state 1 1Δu of the D∞h point group for the linear structure. d Ref. 9, with the 6-311++G(3df) basis set. | ||||||||
CASSCF(12,12) | 4.081 | 3.629 | 1.750 | 103.0 | 633.3 | 320.4 | (2718i) | 4.64 |
CASSCF(16,14) | 4.056 | 3.527 | 1.745 | 102.0 | — | — | — | 3.73 |
CASPT2(12,12) | 3.872 | 3.401 | 1.774 | 104.2 | — | — | — | 5.34 |
EOM-CCSD | 2.964 | 2.387 | 1.684 | 180.0c | 887.9 | 297.3 | 879.4 | 0.00 |
TD-B3LYP | 3.256 | 2.609 | 1.701 | 180.0c | 837.1 | 254.0 | 658.7 | 0.00 |
TD-HSE06 | 3.517 | 2.865 | 1.689 | 180.0c | 853.0 | 216.0 | 701.3 | 0.00 |
TD-ωB97X-D | 3.730 | 2.980 | 1.679 | 179.8 | 853.5 | 237.8 | 751.8 | 0.06 |
TD-CAM-B3LYP | 3.506 | 2.812 | 1.683 | 179.9 | 873.0 | 234.4 | 820.5 | 0.01 |
BPW91 d | — | 2.50 | 1.724 | 143.9 | 767 | 127 | 575 | — |
![]() | ||
Fig. 3 Two-dimensional TD-B3LYP/aug-cc-pVTZ scan on the potential energy surfaces of the lowest singlet states of TiO2. |
It is clearly seen that the S0 1 1A1 ground state possesses only one minimum which has a symmetric bent structure. The linear conformer is located on a saddle point of the ground state surface: it is a minimum along the symmetric stretching coordinate but a barrier top along the bending coordinate. The S1 1 1B2 state, on the other hand, possesses two minima, one being around the Franck–Condon region (r = 1.673 Å, θ = 100.2° and ΔEad = 2.579 eV) and the other being linear (r = 1.700 Å, θ = 180.0° and ΔEad = 2.609 eV). As noted above, these two structures are almost isoenergetic. The linearization energy is just 0.03 eV, although there is a barrier about 0.25 eV between the two conformers.
The energy of the S2 1 1A2 state slides down smoothly as the bending angle increases up to 180°, at which point the state reaches its only minimum degenerate with the 1 1B2 state (r = 1.700 Å, θ = 180.0° and ΔEad = 2.609 eV). The S3 2 1B2 state has only one minimum whose bending angle is larger than the ground equilibrium (r = 1.680 Å, θ = 130.6° and ΔEad = 3.066 eV). In fact this geometry is located in a potential pit, which is much like the coupling region between the two 1B2 surfaces (S1 and S3) with the energy difference of ∼0.2 eV. The potential shapes of S4 2 1A1 and S5 1 1B1 states are somewhat similar to those of S0 1 1A1 and S1 1 1B2 states, respectively, although the potential wells of the higher excited states are much shallower.
It should be noticed that a potential minimum obtained in such a two-dimensional scan does not always guarantee a truly stable structure because there is still a third degree of freedom, i.e. the coordinate of asymmetric stretching motion, which has not been scanned. To check whether the geometry at the minimum is stable, the full optimization has to be executed. In this work we have made sure that the minimum points on the S0 to S3 surfaces do correspond to stable structures, and thus the two-dimensional potential scan could provide meaningful insights into the characters of these states.
In addition to TD-DFT, CASSCF potential energy surface scan was carried out in the same geometric range, and in order to count in dynamical electron correlation, CASPT2 was also applied. The active space was constructed of 12 electrons and 12 orbitals in the scan. The state-average calculation included two 1A1, two 1B1, two 1B2 and two 1A2 states, which are the lowest eight states around the ground equilibrium. It was noticed that the CAS procedure depends subtly on the averaged states, and the results may alter a lot if other “unwanted” states, e.g. the 3 1A1 state which was not counted in, lie below any of the chosen eight states when the geometry is far from the ground equilibrium. This effect caused the surfaces to be rougher than those obtained by TD-DFT, and the data of the highest states were unreliable. Fortunately it had a relatively smaller influence on the ground and lower excited states. The potential energy surfaces of the first six singlet states scanned by multi-reference multi-state CASPT2(12,12) are generally similar to the TD-DFT results, and hence are illustrated as Fig. S1 in the ESI†.
State | Level of theory | Vertical energya | Adiabatic energya | Equilibrium geometry | Vibrational frequenciesb | Dipole moment | |||
---|---|---|---|---|---|---|---|---|---|
ΔEFC/eV | ΔEad/eV | r TiO /Å | θ OTiO /° | ν 1 (a1)/cm−1 | ν 2 (a1)/cm−1 | ν 3 (b2)/cm−1 | μ/debye | ||
a With respect to the global minimum (S0 1 1A1 equilibrium) calculated at the same level of theory. b In the case of linear structure, the three vibrational modes have the symmetry of σ+g, πu and σ−g, respectively. c Ref. 9, with the 6-311++G(3df) basis set. | |||||||||
T 1 1 3B2 | CASSCF(16,14) | 2.003 | 1.781 | 1.694 | 95.3 | 925.7 | 235.6 | (830i) | 1.97 |
EOM-CCSD | 2.324 | 2.135 | 1.660 | 101.9 | 1003.7 | 174.2 | 460.5 | 4.12 | |
TD-B3LYP | 2.556 | 2.500 | 1.671 | 103.0 | 937.3 | 175.9 | 370.7 | 3.49 | |
BPW91 c | — | 2.19 | 1.700 | 96.6 | 882 | 192 | 457 | — | |
T 2 1 3A2 | CASSCF(16,14) | 3.887 | 3.402 | 1.741 | 102.8 | — | — | — | 3.55 |
EOM-CCSD | 2.996 | 2.402 | 1.684 | 180.0 | 886.7 | 288.0 | 862.1 | 0.00 | |
TD-B3LYP | 3.133 | 2.609 | 1.699 | 180.0 | 839.8 | 251.6 | 663.0 | 0.00 | |
BPW91 c | — | 2.53 | 1.722 | 143.7 | 772 | 131 | 561 | — |
Since the neighboring states of the same spatial symmetry but different spin multiplicity are nearly isoenergetic, it is queried if these states could interact with each other through spin–orbit coupling and consequently induce level-splitting and numerous absorption/emission spectral peaks which fit the experimental observation. This could be verified by the spin–orbit coupling Hamiltonian, , where the summation runs over all electrons, ξ is the potential parameter, and
and Ŝ are the orbital and spin angular momentum operators, respectively. It is found that the x, y and z components of
·Ŝ transform as B2, B1 and A2, respectively, in the C2ν point group. As a result, the spin–orbit coupling between states of the same spatial term but different spins is zero, e.g.
where Φ0 denotes the zeroth-order adiabatic electronic wavefunction, since the totally symmetric term in the spatial part of the operator is not available.
How about the spin–orbit coupling between states of different spatial terms? In this work it has been calculated in a perturbative manner at the CASSCF level including five states: the ground, the first two singlet excited and the first two triplet excited states. The spin–orbit coupled results showed negligible shifts in energy eigenvalues of these states, say, less than 0.2 cm−1. Therefore the couplings could be omitted from this system, indicating that the effect contributes extremely little to spectral peaks. A related phenomenon is that the phosphorescence emission, which is preceded by singlet–triplet intersystem crossing through spin–orbit coupling, should be rather weak and hardly observed.
Isomer | State | Level of theory | Adiabatic energya | Equilibrium geometry | Vibrational frequenciesc | Dipole moment | ||||
---|---|---|---|---|---|---|---|---|---|---|
ΔEad/eV | r TiO /Å | r OO /Å | θ/°b | ν 1 (a1)/cm−1 | ν 2 (a1)/cm−1 | ν 3 (b2)/cm−1 | μ/debye | |||
a With respect to the global minimum (S0 1 1A1 equilibrium) of the symmetric bent isomer. b Bending angle refers to θOTiO for cyclic isomers and θTiOO for end-on isomers. c In the case of end-on isomers, the three vibrational modes have the symmetry of σg, π and σg, respectively. d Ref. 10, with the 6-311+G(d) basis set. e Ref. 15, with the 6-311+G(d) basis set. f Ref. 9, with the 6-311+G(3df) basis set. | ||||||||||
Cyclic | 1 3A1 | B3LYP | 3.859 | 1.821 | 1.453 | 47.0 | 953.0 | 704.0 | 520.3 | 4.92 |
B1LYP d | 3.884 | 1.820 | 1.454 | 47.1 | — | — | — | — | ||
BPW91 e | 4.460 | 1.829 | 1.459 | 47.0 | 915 | 665 | 493 | 4.59 | ||
BPW91 f | 4.160 | 1.826 | 1.450 | 46.8 | 939 | 676 | 504 | — | ||
1 1A1 | B3LYP | 4.405 | 1.785 | 1.473 | 48.8 | 936.0 | 681.2 | 629.7 | 3.84 | |
B1LYP d | 4.491 | 1.785 | 1.474 | 48.0 | — | — | — | — | ||
BPW91 f | 4.720 | 1.789 | 1.467 | 48.4 | 930 | 663 | 633 | 3.70 | ||
End-on | 1 3Π | B3LYP | 5.349 | 1.712 | 1.305 | 180.0 | 584.5 | 217.1 | 1145.9 | 6.34 |
B1LYP d | 5.292 | 1.719 | 1.304 | 180.0 | — | — | — | — | ||
BPW91 e | 5.510 | 1.721 | 1.306 | 180.0 | 582 | 210 | 1219 | — | ||
1 1Σg | B3LYP | 6.435 | 1.658 | 1.285 | 180.0 | 668.3 | 276.5 | 1281.9 | 5.28 | |
B1LYP d | 6.467 | 1.656 | 1.288 | 180.0 | — | — | — | — |
Comparing the performances of different levels of theory, all methods achieved very similar results for the ground state. As for excited states, EOM-CCSD appear to be most reliable in both energetic and geometric data. CASSCF fails in excitation energies attributed to lack of dynamical electron correlation, whereas CASPT2 effectively revises this point. TD-DFT with the B3LYP functional is trustworthy, while HSE06, ωB97X-D and CAM-B3LYP functionals always overestimate excitation energies, implying that introducing correction factors which are important in periodic crystal or long-range systems into this simple molecular monomer might not be suitable.
In order to clarify the cause of spectral peaks to the red of the S1 1 1B2 ← S0 1 1A1 absorption system around 2.3 eV, influences from triplet states and other categories of isomers have been further considered. The triplet excited states were found quite close in energy and similar in structure to their corresponding singlet excited states. Spin–orbit couplings between states of the same spatial symmetry are zero and those between any other low-lying states are negligible. Furthermore, cyclic and end-on isomers of titanium dioxide were briefly examined, and all resulted energies are much higher than the equilibria of the S0 to S2 states of the symmetric bent/linear isomer. In consequence, these factors could not give satisfactory interpretation of the spectral peaks to the red. The cause of those low-energy absorption transitions then could be attributed only to singlet states with the symmetric geometry. We suggest that the linear conformer of the ground state might be excluded for its instability, while the linear one of the first singlet excited state might contribute according to its possible low energy.
Footnote |
† Electronic Supplementary Information (ESI) available. See DOI: 10.1039/c1ra00478f/ |
This journal is © The Royal Society of Chemistry 2011 |