Attempts to reconcile simulated photoelectron spectra of MoVO4 clusters are complicated by the presence of very low energy barriers in the potential energy surfaces (PESs) of the lowest energy spin states and isomers. Transition state structures associated with the inversion of terminal oxygen ligands are found to lie below, or close to, the zero point energy of associated modes, which themselves are found to be of low frequency and thus likely to be significantly populated in the experimental characterization. Our simulations make use of Boltzmann averaging over low-energy coordinates and full mapping of the PES to obtain simulations in good agreement with experimental spectra. Furthermore, molecular orbital analysis of accessible final spin states reveals the existence of low energy two-electron transitions in which the final state is obtained from a finite excitation of an electron along with the main photodetachment event. Two-electron transitions are then used to justify the large difference in intensity between different bands present in the photoelectron spectrum. Owing to the general presence of terminal ligands in metal oxide clusters, this study identifies and proposes a solution to issues that are generally encountered when attempting to simulate transition metal cluster photoelectron spectroscopy.

A multitude of studies over many decades has revealed that catalytic activity at transition metal oxide surfaces occurs at edge and point defects resulting from oxygen vacancies.1,2 Transition metal oxides are effective redox promoters as they can readily cycle through two or more oxidation states. This characteristic is often enhanced in bimetallic systems, facilitating high degrees of reactivity and selectivity.3–7 However, a detailed experimental study of such systems in situ is complicated by the small ratio of defect sites to stoichiometric solid resulting in low signal-to-noise ratios. Furthermore, surfaces undergo continuous structural changes such as defect formation, metal reduction, metal oxidation, reagent adsorption, reagent desorption, surface restructuring, and subsurface oxygen diffusion.8–10 Defect sites have incomplete valencies and can exhibit strained electronic structures and atypical metal oxidation states. Thus the study of transition metal oxide clusters is largely motivated by their utility as model systems for defects on catalytic surfaces.11 

Photoelectron spectroscopy has proven to be a powerful tool for characterizing the electronic and molecular structures of metal oxide clusters. Detailed insight into ground and low-lying electronic states of the neutral species can be obtained from the resulting spectra. Observed vibrational excitations reflect the structural changes associated with changing the electron population in the orbital associated with electron detachment. However, analysis is often complicated by the number of possible spin states and structural isomers/conformers. Furthermore, as spectral congestion grows with system size, it becomes increasingly difficult to differentiate fundamental, overtone, and combination bands. Theory-based spectral simulations are a critical tool to aid assignments of species by calculation of electronic states and nuclear configurations, along with associated vibrational modes and transition intensities required for spectral simulation.12–15 

The application of theoretical methods to systems such as metal oxide clusters is further complicated by orbital degeneracies which result in multiconfigurational electronic structure and are intrinsically connected with the observed catalytic activity. Therefore, modeling of spectra formally requires a multi-reference approach; however, application of such methods is typically impractical owing to steep computational scaling. Despite the issues in obtaining a correct description of electronic structure, a large number of studies have employed hybrid Density Functional Theory (DFT) to calculate low energy minima, Adiabatic Detachment Energies (ADEs), Vertical Detatchment Energies (VDEs), and vibrational modes in metal oxide clusters.16–25 Development of efficient approaches for a balanced description of dynamic and static correlation remains an active research area.26,27

An example of profound failure of DFT to appropriately describe a metal oxide cluster based on photoelectron spectral simulation is MoVO4. Definitive assignment of the anion and resulting neutral structures and electronic states was not possible because of an absolute lack of agreement between computation-based simulations and the well-resolved experimental spectrum, which displays a nearly vertical transition with several short but well-resolved vibrational progressions.16 Previous work has indicated that Yamaguchi’s Approximate Projection (AP) method28–39 can correct for spin contamination errors in metal oxide clusters.28 However, the relevant states of MoVO4 displayed low levels of spin contamination. Correct spectral simulation is still not achieved due to the presence of close lying electronic states, low frequency vibrational modes, and geometric symmetry breaking.

In this investigation, we seek to resolve the spectral simulation issues. We therefore attempt to definitely assign the species involved in the observed photodetachment spectrum. In doing so, we reveal the behavior of transition metal oxide clusters to form neutral states accessed by two-electron transitions40–43 and are able to rationalize the observed behavior based on a simple molecular orbital picture.

In order to examine the MoVO4 potential energy surface, the same procedure was followed as our recent study on MoNbO2.28 Following previous work,16–20 isomers are labeled and denoted by dividing the molecule into basins, with the boundary of each basin defined by a line running through each metal center perpendicular to the metal-metal bond. In a bimetallic system, there are three basins and each basin contains oxygen atoms labeled depending on whether they are at the molybdenum terminus (X), bridging the two metals (Y), or at the vanadium terminus (Z). The nomenclature for each structure then is XYZ, where each letter gives the number of oxygens in each basin. For example, 211 indicates that there are two terminal oxygens bound to molybdenum, one bridging oxygen, and one terminal oxygen bound to vanadium. From each isomer, all possible configurations were obtained by considering the highest symmetry point group and all subgroups. Energies of each structure were then obtained for singlet, triplet, quintet, and septet spin states for the anion cluster and doublet, quartet, sextet, and octet spin states for the neutral cluster, generating a total of 432 initial geometries.

All calculations were carried out using a local development version of Gaussian.44 The employed model chemistry was B3PW91 with the SDDPlusTZ basis set, which is comprised of the SDD relativistic pseudopotentials on the molybdenum and vanadium centers to describe core electrons, an augmented Stuttgart/Dresden (SDD) basis set for valence metal orbitals, and aug-cc-pVTZ to describe orbitals on the oxygen atoms. Analysis of the initial and final Kohn-Sham determinants was carried out to ensure that they corresponded to a stable solution.45 Harmonic vibrational modes were computed at the same level of theory used for geometry optimization, and these modes were used in the simulation of Franck-Condon (FC) profiles.46,47 The spectra were simulated at 100 K using the FC approximation with Dushinsky rotations.13–15 Gaussian widths were added to the stick spectra with widths of 0.0062 eV (50 cm−1) and 0.0124 eV (100 cm−1).

In order to reproduce the observed photoelectron spectrum, the models used in the simulations must produce accurate relative spin-state energies, geometries, and force constants. Thus the discussion below is broken into three main sections. We first discuss the identification of the global ground state, which allows us to identify available spin-channels in the photodetachment process. From this, we are able to compute the VDE and ADE resulting from the associated transitions. Second, we discuss the issue of very low energy barriers between different conformational isomers in both MoVO4 anion and neutral states and how the lack of a well defined minimum energy structure complicates the calculation of FC factors. Simulation of vibrational progressions accounting for the presence of multiple low energy barriers demonstrates assignment validity. Finally, we discuss the presence of multiple broken-symmetry electronic structure solutions that suggest the presence of close lying states and use a Molecular Orbital (MO) picture along with Natural Ionization Orbital (NIO) analysis48 to determine the relevant states for comparison with experiment.

In previous work by Mann et al., it was found that two anion structures have nearly identical energies and that either could be the global anion minimum.16 These two structures were the 220 5B1 state and the 121 3A′ state which they calculated to be only 0.06 eV apart. Our extensive search (using a slightly different model chemistry than employed in the original study) did not yield any additional candidate structures. Mulliken spin density analysis and Natural Bond Orbital (NBO) analysis indicated that both unpaired electrons in the 3A′ structure were localized to the vanadium center occupying 3d-like orbitals. An additional electronic state was identified for the 121 structure, 3A″, in which both unpaired electrons occupy orbitals localized on molybdenum. However, this heavily spin contaminated state can be dismissed as irrelevant to this study due to its significantly high energy (+0.558 eV relative to 220 5B1).

In contrast to previous results using the B3LYP functional, our calculations using B3PW91 predicted the 121 3A′ state to be 0.271 eV higher in energy than the 220 5B1 state (Table I). Previous work on FeO2 has indicated the average deviation of B3PW91 energies from multi-reference configuration interaction (MRCI) calculations to be on the order of 0.2 eV.49 Assuming transferability of this result to MoVO4, the B3PW91 results indicate that 220 5B1 is the global minimum. While the 220 5B1 structure is nearly spin pure, the 121 3A′ state exhibits mild spin contamination as determined by the value of S^2. These low levels of spin contamination indicate that the broken symmetry density functional scheme is suitable, in contrast to other recent studies involving other transition binary metal oxide clusters.28 As both B3PW91 and B3LYP studies found no other structures within 0.39 eV of the global minimum, we focus on 220 5B1 and 121 3A′ states for the remainder of this discussion.

TABLE I.

Relative energies, nuclear symmetry, S2, and intermetal bond length of low energy spin states/isomers computed with UB3PW91/SDDPlusTZ.

ElectronicNuclearRelative
StructurestatesymmetryS^2energy (eV)R(Mo–V) (Å)
220 5B1 C2ν 6.009 0.000 2.743 
121 3A′ Cs 2.338 0.271 2.733 
121 3A″ Cs 2.844 0.558 2.770 
ElectronicNuclearRelative
StructurestatesymmetryS^2energy (eV)R(Mo–V) (Å)
220 5B1 C2ν 6.009 0.000 2.743 
121 3A′ Cs 2.338 0.271 2.733 
121 3A″ Cs 2.844 0.558 2.770 

Further evidence that 220 5B1 is an unsuitable candidate for the initial state of the observed experimental spectrum can be obtained by comparing experimental and theoretical spectroscopic parameters, which indicate that transitions from the 220 5B1 state are not consistent with the observed spectrum. Table II shows the computed VDE and ADE for all spin channels accessible from the lowest energy ground states. The 220 neutral ground state is the 4B1 state, and the ADE for the 220 5B1 → 220 4B1 + e transition (2.189 eV) is smaller than the experimentally determined value (2.66 eV). The strong peak intensity at the progression onset of the X and A bands in the experimental spectrum suggests that there is a small structural difference between anion and neutral geometries. The calculated relaxation energy on the neutral energy surface is an order of magnitude larger than the experiment (0.364 eV vs 0.06 eV), resulting in a much broader simulated FC profile than observed experimentally. The transition energy of the alternative spin channel, involving the 220 5B1 → 220 6B1 + e transition, is significantly higher.

TABLE II.

Adiabatic (ADE) and vertical (VDE) detachment energies arising from ionization of the lowest energy anionic structures computed using UB3PW91/SDDPlusTZ. VDE of transitions to 4A″ (2et) and 2A″ (1et2) collapsed to a lower energy state and could not be obtained.

StructureInitial stateFinal stateVDE (eV)ADE (eV)
220 5B1 4B1 2.553 2.189 
220 5B1 6B1 5.264 4.909 
121 3A′ 2A″ (1et) 2.650 2.402 
121 3A′ 4A″ (1et) 2.626 2.464 
121 3A′ 2A″ (2et) 2.736 2.314 
121 3A′ 4A″ (2et) ⋯ 2.365 
121 3A′ 2A″ (1et2) ⋯ 2.537 
Experiment     
X band   2.74 2.66 
A band   2.88 2.82 
StructureInitial stateFinal stateVDE (eV)ADE (eV)
220 5B1 4B1 2.553 2.189 
220 5B1 6B1 5.264 4.909 
121 3A′ 2A″ (1et) 2.650 2.402 
121 3A′ 4A″ (1et) 2.626 2.464 
121 3A′ 2A″ (2et) 2.736 2.314 
121 3A′ 4A″ (2et) ⋯ 2.365 
121 3A′ 2A″ (1et2) ⋯ 2.537 
Experiment     
X band   2.74 2.66 
A band   2.88 2.82 

In the case of the transitions initiated from 121 3A′, two spin channels are energetically accessible: 121 2A″ and 121 4A″. Three different 121 2A′ minima were found, one of which is not accessible from the anion through a one-electron transition, while two 121 4A′ minima were located, one of which is not one-electron accessible. The lowest energy states of both the doublet and quartet are not accessible through one-electron transitions. Analysis of the electronic structure of the species involved in these transitions indicated that the relevant transitions are the 121 3A′ → 121 2A″ (1et) + e and 121 3A′ → 121 4A″ (2et) + e transitions (vide infra). In this section, we discuss the structures and properties of the clusters that must be accounted for in order to simulate the photodetachment spectrum before returning to the nature of the electronic states of these transitions.

Fig. 1 shows the structures involved in the lowest energy transitions. The 220 5B1 structure has C2ν symmetry and is well defined assuming a single potential energy well around the minimum. The 121 3A′ structure has Cs symmetry and, in contrast to the 220 5B1 structure, has flexible terminal oxygen atoms with low energy barriers to cis-trans isomerization across the plane formed by the two metal centers and two bridging oxygens. Mo–V–O are collinear and the motion associated with this angle bending is 104 cm−1, while the V–Mo–O bend is out of the plane with frequency 50 cm−1. The barrier for isomerization of the V–Mo–O bend is very small at just 0.42 kcal mol−1 (146 cm−1), well below the zero point energy.

FIG. 1.

Structures of the species involved in the low energy photodetachment transitions obtained using UB3PW91/SDDPlusTZ.

FIG. 1.

Structures of the species involved in the low energy photodetachment transitions obtained using UB3PW91/SDDPlusTZ.

Close modal

The bending of the terminal oxygens to break C2ν symmetry can be understood by considering the ligand environment of each metal. The oxygen atoms around Mo and V form an effective MX3 environment in a trigonal planar orientation. The ligand field splitting of the metal d-shell in such an environment gives rise to a degenerate dxy and dx2y2 pair being higher in energy than the degenerate dxz and dyz set and the dz2 orbital.50,51 Depending on the nature of the ligand, the order of the dxz and dyz pair and the dz2 orbital will change. σ donating ligands favor the dxz and dyz pair while π donating ligands will give the dz2 orbital as the most stable in the d manifold. The presence of Jahn-Teller distortion, achieved by bending of the terminal oxygen out of the plane, will depend on the order of the dxz/dyz pair and the dz2 orbital. If the metal center is d1, the symmetry distortion will occur when the dxz and dyz pair is lowest in energy; if the dz2 orbital is lowest in energy, only d2 metals will undergo symmetry distortion. As the oxygen ligands are both σ (pz) and π (px, py) donating, the exact order of these lowest energy d orbitals depends upon the particular geometry and the ratio of ligand electron donation types to the metal center.

The neutral structures associated with the transitions reflect the complexity observed in the anions from which they originate. The lowest energy vibrational modes in the 220 4B1 structure are 108 cm−1 and 229 cm−1 associated with wagging and bending of the two terminal oxygens. As with the 220 5B1 anion ground state, the 220 4B1 structure is described well by a single harmonic well. The three 121 doublet minima all show out-of-plane bending of both terminal oxygen atoms, including the Mo–V–O angle that is in the plane of the metals and bridging oxygens in the 121 3A′ anion structure. In the 121 2A″ (1et) structure, which was identified as the relevant state for comparison with experiment, the V–Mo–O bend has a barrier at 180° of 5.81 kcal mol−1 (harmonic frequency of the associated vibrational mode is 39 cm−1). While this barrier is relatively high, comparison with the 121 3A′ potential energy surface (PES) shows that it lies along the valley connecting the two minima in the 121 3A′ structure and so is likely to be obtained as the initial structure on the neutral state after photodetachment—the FC structure (Fig. 2, center). Thus 121 2A″ (1et) is a double well system with two possible equivalent paths (Fig. 2, left). Similar double well systems have been identified and studied in tungsten oxide clusters where the presence of the double-well in the excited anion state renders FC based simulations inadequate and some form of conformational averaging must be used instead.19 A similar pattern is observed on the 121 4A″ (2et) PES, although the two equivalent minima lie along the Mo–V–O bending coordinate orthogonal to the double-well coordinate in the 121 2A″ (1et) structure (Fig. 2, right). The 121 4A″ Mo–V–O coordinate has a barrier of 2.65 kcal mol−1 with a vibrational frequency of 116 cm−1. There is a very small barrier to inversion of the V–Mo–O angle of just 0.01(47) kcal mol−1 to a trans minimum which lies 0.01(13) kcal mol−1 above the cis minimum. Given that the energy difference is so small, inversion of this coordinate can be considered barrierless.

FIG. 2.

Potential energy surfaces of the 121 structures identified as candidates for the observed experimental Frank-Condon profiles plotted with respect to bending of the terminal oxygens. All three surfaces show a double well indicating that simulations of Franck-Condon profiles require structural averaging with regard to these coordinates.

FIG. 2.

Potential energy surfaces of the 121 structures identified as candidates for the observed experimental Frank-Condon profiles plotted with respect to bending of the terminal oxygens. All three surfaces show a double well indicating that simulations of Franck-Condon profiles require structural averaging with regard to these coordinates.

Close modal

Given the low barriers for inversion, we considered the possibility that differences between the experimental and theoretical ADE and VDE values may stem from the fact that the minimum energy structures do not give a correct representation of the thermally accessible ensemble of structures. While the 220 5B1 → 220 4B1 + e simulation is relatively straightforward, as previously discussed, the presence of multiple/flat minima in the 121 3A′ → 121 4A″ (2et) + e and 121 3A′ → 121 2A″ (1et) + e transitions complicates the simulation of FC progressions as they assume the structures are described by a single harmonic well. In order to simulate these species, we follow the protocol used by Waller et al.19 described in Sec. II whereby we optimize to the critical point closest to the Boltzmann weighted average of the 121 3A′ structures and ignore imaginary and very low frequency modes.

Figure 3 shows a comparison of the calculated FC spectrum with experiment for the three likely transitions identified above. The simulated spectra are shown with added Gaussian widths of 50 cm−1 (blue) and 100 cm−1 (red). The simulations were carried out at a temperature of 100 K. Clearly the 121 3A′ → 121 2A″ (1et) + e transition gives good agreement with the experimental A band. In particular, the profile shows a very intense peak at the onset consistent with experiment, the progressions have similar peak spacings, and the higher energy bands are broader due to an increase in the density of transitions. The 220 5B1 → 220 4B1 + e and 121 3A′ → 121 4A″ (2et) + e transitions are both broad with similar peak spacing in vibrational progressions and hence either can be considered to be in reasonable agreement with the X band. The fact that the 121 3A′ → 121 4A″ (2et) + e transition is a two-electron transition provides an excellent explanation of the high intensity difference between the A and X bands. Therefore, as the one-electron 220 5B1 → 220 4B1 + e transition should be of comparable intensity to the 121 3A′ → 121 2A″ + e transition, we use the relative intensities of the A and X bands to tentatively exclude the 220 5B1 → 220 4B1 + e transition, along with experimental evidence that both transitions arise from a common anion.16 Furthermore, the 220 5B1 → 220 4B1 + e transition is beyond the typical 0.3 eV error range for ADE and VDE energies.18,52 Thus we assign the X band to the 121 3A′ → 121 4A″ (2et) + e transition.

FIG. 3.

Computed Franck-Condon progressions of the transitions identified as potentially giving rise to the observed transitions (black sticks) compared to the orientational averaged experimental spectrum (top, black line). Simulated spectra were computed using harmonic modes obtained at the UB3PW91/SDDPlusTZ level of theory without accounting for spin contamination with spectral Gaussian linewidths of 50 cm−1 (blue) and 100 cm−1 (red).

FIG. 3.

Computed Franck-Condon progressions of the transitions identified as potentially giving rise to the observed transitions (black sticks) compared to the orientational averaged experimental spectrum (top, black line). Simulated spectra were computed using harmonic modes obtained at the UB3PW91/SDDPlusTZ level of theory without accounting for spin contamination with spectral Gaussian linewidths of 50 cm−1 (blue) and 100 cm−1 (red).

Close modal

To obtain a direct comparison with experimental results, the simulated 121 3A′ → 121 4A″ (2et) + e and 121 3A′ → 121 2A″ (1et) + e transitions were fitted to experiment minimizing the mean squared error of a combined spectrum with a spectral Gaussian linewidth of 50 cm−1 (Fig. 4). The resulting spectrum was obtained with a mean squared error of 0.005. The 121 3A′ → 121 2A″ (1et) + e and 121 3A′ → 121 4A″ (2et) + e transitions were shifted by 0.316 and 0.294 eV and scaled by 0.812 and 0.525, respectively. The composite profile gives good agreement between the experimental and simulated progressions and lends weight to the transition assignments made above.

FIG. 4.

Combined simulated Franck-Condon progression spectra of the 121 3A′ → 121 2A″ (1et) + e and 121 3A′ → 121 4A″ (2et) + e transitions shifted by 0.316 and 0.294 eV, respectively, and scaled by 0.812 and 0.525, respectively, to minimize the mean squared error of 78 evenly distributed energy values with the experimental profile (MSE =5.11×103). Simulated spectra were computed using harmonic modes obtained using B3PW91/SDDPlusTZ and Gaussian linewidth of 50 cm−1.

FIG. 4.

Combined simulated Franck-Condon progression spectra of the 121 3A′ → 121 2A″ (1et) + e and 121 3A′ → 121 4A″ (2et) + e transitions shifted by 0.316 and 0.294 eV, respectively, and scaled by 0.812 and 0.525, respectively, to minimize the mean squared error of 78 evenly distributed energy values with the experimental profile (MSE =5.11×103). Simulated spectra were computed using harmonic modes obtained using B3PW91/SDDPlusTZ and Gaussian linewidth of 50 cm−1.

Close modal

Again, three 121 2A″ minima were located: (i) [B] 121 2A″ (1et) with two unpaired electrons on vanadium and a single anti-parallel spin electron on molybdenum (Fig. 5); (ii) [C] 121 2A″ (2et) with two parallel electrons on molybdenum and a single anti-parallel electron on vanadium (Fig. 5(c)); and (iii) [D] 121 2A″ (1et2) with two spin paired electrons on molybdenum and a single electron on vanadium (Fig. 5(d)). [C] 121 2A″ (2et) was found to be lowest in energy by 0.088 eV compared to [B] 121 2A″ (1et), while [D] 121 2A″ (1et2) was higher in energy by 0.135 eV. Attempts to find a further doublet state with α and β V electrons and an α Mo electron collapsed to the [D] 121 2A″ (1et2) state and was thus discarded as being too high in energy. NIO analysis48 indicated that transition from [A] 121 3A′ to [B] 1212A″ (1et) and [D] 121 2A″ (1et2) states results from one-electron transitions from molybdenum and vanadium d-orbitals, respectively. On the other hand, transition to the [C] 121 2A″ (2et) state is accessed via a two-electron transition (Fig. 6).

FIG. 5.

Molecular orbital diagram showing electron arrangement in molybdenum and vanadium transition metal centers after one-electron photodetachment. Electron transfer in the quartet can proceed directly between the metal centers. In the doublet direct electron transfer between metal centers leads to the electronic state with low spin on the molybdenum which is higher in energy. In order to access the electronic state with high spin on the molybdenum which is lower in energy an electron must be transferred through the ligand. However, this breaks the super-exchange interaction and results in violation of Pauli repulsion.

FIG. 5.

Molecular orbital diagram showing electron arrangement in molybdenum and vanadium transition metal centers after one-electron photodetachment. Electron transfer in the quartet can proceed directly between the metal centers. In the doublet direct electron transfer between metal centers leads to the electronic state with low spin on the molybdenum which is higher in energy. In order to access the electronic state with high spin on the molybdenum which is lower in energy an electron must be transferred through the ligand. However, this breaks the super-exchange interaction and results in violation of Pauli repulsion.

Close modal
FIG. 6.

Selected NIOs (δ>0.5) computed from the 3A′ → 2A″ (1et) + e transition (top) and the 3A′ → 2A″ (2et) + e transition (bottom). Indicated in the figure are the occupation change number δelec for each NIO, where a value of −1.00 indicates the detached electron hole. In the case of the 3A′ → 2A″ (2et) + e transition, loss of an α electron residing within the vanadium d-shell is accompanied by a β electron transition from the molybdenum d-shell to the same d-shell orbital on vanadium from which the electron was ionized. Furthermore there is a significant redistribution of the α d-shell electron density to counter the hole left on molybdenum. The transition was thus identified as a two-electron transition in which one electron ionization is accompanied by promotion of a second electron. Orbital surfaces shown at a density contour value of 0.04 a.u.

FIG. 6.

Selected NIOs (δ>0.5) computed from the 3A′ → 2A″ (1et) + e transition (top) and the 3A′ → 2A″ (2et) + e transition (bottom). Indicated in the figure are the occupation change number δelec for each NIO, where a value of −1.00 indicates the detached electron hole. In the case of the 3A′ → 2A″ (2et) + e transition, loss of an α electron residing within the vanadium d-shell is accompanied by a β electron transition from the molybdenum d-shell to the same d-shell orbital on vanadium from which the electron was ionized. Furthermore there is a significant redistribution of the α d-shell electron density to counter the hole left on molybdenum. The transition was thus identified as a two-electron transition in which one electron ionization is accompanied by promotion of a second electron. Orbital surfaces shown at a density contour value of 0.04 a.u.

Close modal

There are two 121 4A″ minima: (1) [E] 121 4A″ (1et) which retains two unpaired electrons on vanadium and has a single parallel spin electron on molybdenum (Fig. 5(e)); and (2) [F] 121 4A″ (2et) which has the opposite electronic configuration with two unpaired electrons on molybdenum and an unpaired parallel spin electron on vanadium (Fig. 5(f)). As with the doublet, NIO analysis confirmed that the transition to the [F] 121 4A″ (2et) structure is a two-electron transition while the [E] 121 4A″ (1et) structure results from a one-electron transition (Fig. 7).53 As shown in Table II, [F] 121 4A″ (2et) is lower in energy than [E] 121 4A″ (1et) by 0.099 eV. We note that attempts to compute VDEs to the 4A″ (2et) state collapsed to the 4A″ (1et) electronic structure. Exploration of the PES located a transition state between the [E] 121 4A″ (1et) and [F] 121 4A″ (2et) structures corresponding to lateral motion of the bridging oxygens between the two metal centers only 0.025 eV above the [E] 121 4A″ (1et) structure and 0.124 eV above the [F] 121 4A″ (2et) structure. The transition structure that interconverts the two electronic states implies the existence of a crossing of the two states of interest. Furthermore, our calculations suggest that the intersystem crossing point is quite low in energy and we expect interconversion between these two states to occur relatively rapidly on the experimental time scale.

FIG. 7.

Selected NIOs (δ>0.5) computed from the 3A′ → 4A″ (1et) + e transition (top) and the 3A′ → 4A″ (2et) + e transition (bottom). Indicated in the figure are the occupation change number δelec for each NIO, where a value of −1.00 indicates the detached electron hole. In both transitions, the ionized β electron resides in the molybdenum d-shell. The 3A′ → 4A″ (1et) + e transition also involves a significant relaxation of α density to counter the ionized electron hole. The 3A′ → 4A″ (2et) + e transition superficially looks similar to the one-electron transition except that the α electron relaxation is greater—to the extent that it can be considered an electron transfer from the vanadium d-shell to the molybdenum d-shell. As described in the text, the 4A″ (2et) state could not be obtained at the 3A′ geometry because it collapsed to the 4A″ (1et) state. Thus the NIOs shown are obtained from the closest 4A″ (2et) state structure to the 3A′ geometry we could obtain. The use of these different geometries results in slight asymmetry in the occupation change eigenvalues because they also account for some geometry dependent density changes. Orbital surfaces shown at a value of 0.04.

FIG. 7.

Selected NIOs (δ>0.5) computed from the 3A′ → 4A″ (1et) + e transition (top) and the 3A′ → 4A″ (2et) + e transition (bottom). Indicated in the figure are the occupation change number δelec for each NIO, where a value of −1.00 indicates the detached electron hole. In both transitions, the ionized β electron resides in the molybdenum d-shell. The 3A′ → 4A″ (1et) + e transition also involves a significant relaxation of α density to counter the ionized electron hole. The 3A′ → 4A″ (2et) + e transition superficially looks similar to the one-electron transition except that the α electron relaxation is greater—to the extent that it can be considered an electron transfer from the vanadium d-shell to the molybdenum d-shell. As described in the text, the 4A″ (2et) state could not be obtained at the 3A′ geometry because it collapsed to the 4A″ (1et) state. Thus the NIOs shown are obtained from the closest 4A″ (2et) state structure to the 3A′ geometry we could obtain. The use of these different geometries results in slight asymmetry in the occupation change eigenvalues because they also account for some geometry dependent density changes. Orbital surfaces shown at a value of 0.04.

Close modal

Fig. 5 illustrates how the available detachment processes can be interpreted in terms of different valence bond structures and how the ability to transition between electronic states adiabatically can be rationalized using a simple orbital picture. Analysis of [C] 121 2A″ (2et), [D] 121 2A″ (1et2), and [F] 121 4A″ (2et) structures involves Mo(IV) and V(IV) oxidation states, while [B] 2A″ (1et) and [E] 4A″ (1et) involve Mo(V) and V(III) oxidation states. Thus the metals are either d1 or d2 and are able to couple magnetically through bridging oxygen p-orbitals. In the case of the 121 4A″ states, the electronic structure of the [F] 4A″ (2et) structure (Fig. 5(f)) is obtained from the [E] 4A″ (1et) structure (Fig. 5(e)) by transfer of an electron from vanadium to molybdenum. As all electrons in the metal d-orbitals have the same spin, spin-conserving electron transfer likely occurs without significant energy penalty, coupled through nuclear vibrations of the bridging oxygen centers. The lateral position of the bridging oxygens indicates the position of the transferred electron. The low energy barrier and the lower energy of the minimum likely result in the [F] 4A″ (2et) structure (Fig. 5(f)) to be favored in this case. After photodetachment, it is expected, therefore, that the [F] 4A″ (2et) structure corresponding to the two-electron transition would be present, though, spectroscopic measurements would be significantly less intense than for one-electron transitions.

It is possible that another triplet ground state exists with triplet-coupling of Mo electrons forming an (overall) triplet with the two V electrons being singlet-coupled. This would allow direct access to the [F] 121 4A″ (2et) state upon photodetachment. However, we were unable to locate such an electronic structure, and although it is not possible to conclusively state that this alternative triplet state is not present, within the constraints of the tractable and available theory at our disposal this appears to be a reasonable conclusion. This conclusion is further supported by the experimental photoelectron spectrum which would not display the weaker X band intensity if the transition was one-electron accessible.

As shown in Fig. 5, the antiferromagnetically coupled metal centers facilitate a superexchange mechanism in the 2A″ structures. Similar to the situation observed in the quartet, [B] 121 2A″ (1et) (Fig. 5(b)) and [D] 121 2A″ (1et2) (Fig. 5(d)) structures lie on the same PES connected by a spin-conserving electron transfer. [C] 121 2A″ (2et) (Fig. 5(c)) is the lowest energy structure; however, there is no direct, one-electron pathway to concurrently detach an electron and rearrange the electronic structure. Instead, such a change requires the exchange of electron density between metal centers from the lowest VDE state [B] 2A″ (1et). Owing to the extensive electron rearrangement occurring in response to electron detachment, we expect the [C] 2A″ (2et) state is not readily accessed. The [B] 121 2A″ (1et) structure (Fig. 5(b)) corresponding to a one-electron transition is therefore the relevant minimum even though the [C] 121 2A″ (2et) (Fig. 5(c)) structure is lower in energy.

To summarize, analysis of the different accessible spin channels suggests that transition from the anion 3A′ to the neutral 2A″ state is a one-electron transition, while transition to the neutral 4A″ state arises from a two-electron transition. The different selection rules for one-electron and two-electron transitions provide an explanation of the difference in intensity between the experimental X and A bands.

The transitions responsible for the observed FC profile of MoVO4 were assigned as the 121 3A′ → 121 4A″ + e transition for the X band and the 121 3A′ → 121 2A″ + e transition for the A band. A potential third transition, resulting from 220 5B1 → 220 4B1 + e, was excluded due to experimental evidence suggesting that the X and A bands arise from a common isomer and discrepancies between computed VDE and ADE and observed experimental values. Although the 220 5B1 structure was computed to be lower in energy that the 121 3A′ structure, further justification for the exclusion of the 220 5B1 → 220 4B1 + e transition is based on the tendency of hybrid functionals to overstabilize higher spin states.

Discrepancies between simulated FC profiles in previous reports and the experimental photoelectron spectrum have been shown to result from significant geometric distortion along either very low frequency vibrational modes or over low energy barriers resulting from Jahn-Teller distortions. The presence of flat or multiple minima on the PES causes FC simulations, which assume a single well-defined minimum in both the ground and excited states, to be invalid and yield poor agreement with experimental and simulated FC profiles. In order to correctly simulate FC profiles, Boltzmann weighted structural averaging was used as well as exclusion of imaginary and low frequency vibrational modes from the FC simulation.

A further consideration is the presence of a number of broken symmetry solutions in the neutral states that complicate the assignment of the observed transitions. Broken symmetry density functional theory calculations were able to provide both the energy and electron distribution of these states. Within a single-reference approach such as DFT, transition structures may qualitatively indicate regions where different states approach each other and adiabatic transitions between states can occur. In MoVO4, the pertinent transition vector involves the lateral shifting of the bridging oxygens and is associated with electron transfer between the metal centers. Rationalizating electron transfer mechanisms using an orbital analysis, it was possible to identify the 121 3A′ → 121 4A″ + e transition as a two-electron process while the 121 3A′ → 121 2A″ + e transition was determined to be a one-electron event. Accounting for differences in transition types gave agreement with the observed experimental intensity difference between X and A bands.

See supplementary material for geometries, energies, and spin densities of computed anion and neutral MoVO4 structures as well as for the geometry used to compute natural ionization orbitals on the 121 3A′ → 1214A″ (2et) transition.

We gratefully acknowledge financial support from the University of California, Merced, and computing time on the Multi-Environment Computer for Exploration and Discovery (MERCED) cluster which is supported by National Science Foundation Grant No. ACI-1429783.

1.
A. G.
Anshits
,
E. N.
Voskresenskaya
, and
L. I.
Kurteeva
,
Catal. Lett.
6
,
67
(
1990
).
2.
R. A.
Van Santen
and
M.
Neurock
,
Molecular Heterogeneous Catalysis: A Conceptual and Computational Approach
(
Wiley-VCH
,
Weinheim, Germany
,
2009
).
3.
M.
Sadakane
,
N.
Watanabe
,
T.
Katou
,
Y.
Nodasaka
, and
W.
Ueda
,
Angew. Chem., Int. Ed.
46
,
1493
(
2007
).
4.
M.
Nößler
,
R.
Mitrić
,
V.
Bonačić-Koutecký
,
G. E.
Johnson
,
E. C.
Tyo
, and
A. W.
Castleman
,
Angew. Chem., Int. Ed.
49
,
407
(
2010
).
5.
M.
Ziolek
and
I.
Nowak
,
Catal. Today
78
,
543
(
2003
).
6.
C.
Tagusagawa
,
A.
Takagaki
,
A.
Iguchi
,
K.
Takanabe
,
J. N.
Kondo
,
K.
Ebitani
,
T.
Tatsumi
, and
K.
Domen
,
Catal. Today
164
,
358
(
2011
).
7.
P.
Botella
,
A.
Dejoz
,
M. C.
Abello
,
M. I.
Vázquez
,
L.
Arrúa
, and
J. M.
López Nieto
,
Catal. Today
142
,
272
(
2009
).
8.
U. S.
Ozkan
and
R. B.
Watson
,
Catal. Today
100
,
101
(
2005
).
9.
R.
Schlögl
,
A.
Knop-Gericke
,
M.
Hävecker
,
U.
Wild
,
D.
Frickel
,
T.
Ressler
,
R. E.
Jentoft
,
J.
Wienold
,
G.
Mestl
,
A.
Blume
,
O.
Timpe
, and
Y.
Uchida
,
Top. Catal.
15
,
219
(
2001
).
10.
E.
Hoefs
,
J. Catal.
57
,
331
(
1979
).
11.
D. K.
Böhme
and
H.
Schwarz
,
Angew. Chem., Int. Ed.
44
,
2336
(
2005
).
12.
K.
Bravaya
and
A. I.
Krylov
,
J. Phys. Chem. A
117
,
11815
(
2013
).
13.
F.
Santoro
,
R.
Improta
,
A.
Lami
,
J.
Bloino
, and
V.
Barone
,
J. Chem. Phys.
126
,
084509
(
2007
).
14.
F.
Santoro
,
A.
Lami
,
R.
Improta
,
J.
Bloino
, and
V.
Barone
,
J. Chem. Phys.
128
,
224311
(
2008
).
15.
F.
Santoro
,
A.
Lami
,
R.
Improta
, and
V.
Barone
,
J. Chem. Phys.
126
,
184102
(
2007
).
16.
J. E.
Mann
,
D. W.
Rothgeb
,
S. E.
Waller
, and
C. C.
Jarrold
,
J. Phys. Chem. A
114
,
11312
(
2010
).
17.
B. L.
Yoder
,
J. T.
Maze
,
K.
Raghavachari
, and
C. C.
Jarrold
,
J. Chem. Phys.
122
,
094313
(
2005
).
18.
S. E.
Waller
,
J. E.
Mann
,
D. W.
Rothgeb
, and
C. C.
Jarrold
,
J. Phys. Chem. A
116
,
9639
(
2012
).
19.
S. E.
Waller
,
M.
Ray
,
B. L.
Yoder
, and
C. C.
Jarrold
,
J. Phys. Chem. A
117
,
13919
(
2013
).
20.
N. J.
Mayhall
,
D. W.
Rothgeb
,
E.
Hossain
,
K.
Raghavachari
, and
C. C.
Jarrold
,
J. Chem. Phys.
130
,
124313
(
2009
).
21.
D. M.
Neumark
,
Annu. Rev. Phys. Chem.
52
,
255
(
2001
).
22.
T.
Andersen
,
K. R.
Lykke
,
D. M.
Neumark
, and
W. C.
Lineberger
,
J. Chem. Phys.
86
,
1858
(
1987
).
23.
G.
Meloni
,
M. J.
Ferguson
, and
D. M.
Neumark
,
Phys. Chem. Chem. Phys.
5
,
4073
(
2003
).
24.
J. B.
Kim
,
M. L.
Weichman
, and
D. M.
Neumark
,
Phys. Chem. Chem. Phys.
15
,
20973
(
2013
).
25.
G.
Meloni
,
S. M.
Sheehan
, and
D. M.
Neumark
,
J. Chem. Phys.
122
,
074317
(
2005
).
26.
A. J.
Garza
,
C. A.
Jiménez-Hoyos
, and
G.
Scuseria
,
J. Chem. Phys.
140
,
244102
(
2014
).
27.
L. M.
Thompson
and
H. P.
Hratchian
,
J. Chem. Phys.
141
,
034108
(
2014
).
28.
L. M.
Thompson
and
H. P.
Hratchian
,
J. Phys. Chem. A
119
,
8744
(
2015
).
29.
L. M.
Thompson
and
H. P.
Hratchian
,
J. Chem. Phys.
142
,
054106
(
2015
).
30.
K.
Yamaguchi
,
F.
Jensen
,
A.
Dorigo
, and
K. N.
Houk
,
Chem. Phys. Lett.
149
,
537
(
1988
).
31.
S.
Yamanaka
,
M.
Okumura
,
M.
Nakano
, and
K.
Yamaguchi
,
J. Mol. Struct.: THEOCHEM
310
,
205
(
1994
).
32.
H. P.
Hratchian
,
J. Chem. Phys.
138
,
101101
(
2013
).
33.
T.
Saito
and
W.
Thiel
,
J. Phys. Chem. A
116
,
10864
(
2012
).
34.
L.
Noodleman
,
D. A.
Case
, and
A.
Aizman
,
J. Am. Chem. Soc.
110
,
1001
(
1988
).
35.
L.
Noodleman
and
D. A.
Case
, in
Advances in Inorganic Chemistry
, edited by
R.
Cammack
and
A. G.
Sykes
(
Academic Press
,
Orlando
,
1992
), p.
423
.
36.
J. M.
Mouesca
,
L.
Noodleman
,
D. A.
Case
, and
B.
Lamotte
,
Inorg. Chem.
34
,
4347
(
1995
).
37.
L.
Noodleman
,
J. G.
Norman
, Jr
, and
J. G.
Norman
,
J. Chem. Phys.
70
,
4903
(
1979
).
38.
L.
Noodleman
,
J. Chem. Phys.
74
,
5737
(
1981
).
39.
L.
Noodleman
,
C. Y.
Peng
,
D. A.
Case
, and
J. M.
Mouesca
,
Coord. Chem. Rev.
144
,
199
(
1995
).
40.
J. O.
Kafader
,
M.
Ray
, and
C. C.
Jarrold
,
J. Chem. Phys.
143
,
034305
(
2015
).
41.
J. O.
Kafader
,
M.
Ray
, and
C. C.
Jarrold
,
J. Chem. Phys.
143
,
064305
(
2015
).
42.
C.-Y.
Cha
,
G.
Ganteför
, and
W.
Eberhardt
,
J. Chem. Phys.
99
,
6308
(
1993
).
43.
H.-J.
Zhai
,
W.-J.
Chen
,
X.
Huang
, and
L.-S.
Wang
,
RSC Adv.
2
,
2707
(
2012
).
44.
M. J.
Frisch
,
G. W.
Trucks
,
H. B.
Schlegel
,
G. E.
Scuseria
,
M. A.
Robb
,
J. R.
Cheeseman
,
G.
Scalmani
,
V.
Barone
,
B.
Mennucci
,
G. A.
Petersson
,
H.
Nakatsuji
,
M.
Caricato
,
X.
Li
,
H. P.
Hratchian
,
A. F.
Izmaylov
,
J.
Bloino
,
B.
Janesko
,
F.
Lipparini
,
G.
Zheng
,
J. L.
Sonnenberg
,
W.
Liang
,
M.
Hada
,
M.
Ehara
,
K.
Toyota
,
R.
Fukuda
,
J.
Hasegawa
,
M.
Ishida
,
T.
Nakajima
,
Y.
Honda
,
O.
Kitao
,
H.
Nakai
,
T.
Vreven
,
J. A.
Montgomery
, Jr.
,
J. E.
Peralta
,
F.
Ogliaro
,
M.
Bearpark
,
J. J.
Heyd
,
E.
Brothers
,
K. N.
Kudin
,
V. N.
Staroverov
,
T.
Keith
,
R.
Kobayashi
,
J.
Normand
,
K.
Raghavachari
,
A.
Rendell
,
J. C.
Burant
,
S. S.
Iyengar
,
J.
Tomasi
,
M.
Cossi
,
N.
Rega
,
J. M.
Millam
,
M.
Klene
,
J. E.
Knox
,
J. B.
Cross
,
V.
Bakken
,
C.
Adamo
,
J.
Jaramillo
,
R.
Gomperts
,
R. E.
Stratmann
,
O.
Yazyev
,
A. J.
Austin
,
R.
Cammi
,
C.
Pomelli
,
J. W.
Ochterski
,
R. L.
Martin
,
K.
Morokuma
,
V. G.
Zakrzewski
,
G. A.
Voth
,
P.
Salvador
,
J. J.
Dannenberg
,
S.
Dapprich
,
P. V.
Parandekar
,
N. J.
Mayhall
,
A. D.
Daniels
,
O.
Farkas
,
J. B.
Foresman
,
J. V.
Ortiz
,
J.
Cioslowski
, and
D. J.
Fox
, gaussian Development Version, Revision I.02+,
Gaussian, Inc.
,
Wallingford CT
,
2014
.
45.
H. B.
Schlegel
and
J. J. W.
McDouall
, “
Do you have SCF stability and convergence problems?
,”
Computational Advances in Organic Chemistry: Molecular Structure and Reactivity
(
Springer
,
Netherlands, Dordrecht
,
1991
).
46.
G.
Scuseria
,
R. E.
Stratmann
,
J. C.
Burant
, and
M. J.
Frisch
,
J. Chem. Phys.
106
,
10175
(
1997
).
47.
K.
Raghavachari
,
J. A.
Pople
, and
H. B.
Schlegel
,
Int. J. Quantum Chem.
16
,
225
(
1979
).
48.
L. M.
Thompson
,
H.
Harb
, and
H. P.
Hratchian
,
J. Chem. Phys.
144
,
204117
(
2016
).
49.
F.
Grein
,
Int. J. Quantum Chem.
109
,
549
(
2009
).
50.
M. A.
Halcrow
,
Chem. Soc. Rev.
42
,
1784
(
2013
).
51.
M. J.
Muldoon
,
Dalton Trans.
39
,
337
(
2010
).
52.
J. N.
Harvey
, in
Principles and Applications of Density Functional Theory in Inorganic Chemistry I
, edited by
N.
Kaltsoyanis
and
J. E.
McGrady
(
Springer
,
Berlin, Germany
,
2004
).
53.

As we were unable to obtain the 4A″ (2et) electronic state at the 3A′ geometry because it collapses to the 4A″ (1et) state, the density was described by using a nearby point on the potential energy surface where we could converge the 4A″ (2et) state. This nearby structure was obtained by taking the 3A′ state minimum geometry and altering the bridging oxygen atoms to their position in the 4A″ (2et) state. The electronic structure of this new geometry could be converged without collapsing to the 4A″ (1et) state.

Supplementary Material