ACS Publications. Most Trusted. Most Cited. Most Read
Oxygen Exchange on Vanadium Pentoxide
My Activity

Figure 1Loading Img
  • Open Access
C: Chemical and Catalytic Reactivity at Interfaces

Oxygen Exchange on Vanadium Pentoxide
Click to copy article linkArticle link copied!

  • Yuanqing Wang
    Yuanqing Wang
    Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4−6, D-14195 Berlin, Germany
    BasCat, UniCat BASF Jointlab, Technische Universität Berlin, D-10623 Berlin, Germany
  • Frank Rosowski
    Frank Rosowski
    BasCat, UniCat BASF Jointlab, Technische Universität Berlin, D-10623 Berlin, Germany
    Heterogeneous Catalysis, BASF SE, Process Research and Chemical Engineering, Carl-Bosch-Straße 38, D-67065 Ludwigshafen, Germany
  • Robert Schlögl
    Robert Schlögl
    Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4−6, D-14195 Berlin, Germany
    Max-Planck-Institut für Chemische Energiekonversion, Stiftstrase 34−36, D-45470 Mülheim, Germany
  • Annette Trunschke*
    Annette Trunschke
    Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4−6, D-14195 Berlin, Germany
    *Email: [email protected]
Open PDFSupporting Information (1)

The Journal of Physical Chemistry C

Cite this: J. Phys. Chem. C 2022, 126, 7, 3443–3456
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcc.2c00174
Published February 9, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

The isotopic exchange of 18O2 on polycrystalline V216O5 was studied by Raman spectroscopy at different temperatures between 300 and 580 °C and in the presence of different mixtures of oxygen with ethane, propane, or n-butane in the gas phase. Supported by DFT calculations, a method was developed to determine which of the three differently coordinated oxygen atoms in the crystal structure of V2O5 (vanadyl oxygen O1, 2-fold-coordinated oxygen O2, and three-coordinated oxygen O3) are involved in the exchange with 18O2 from the gas phase. Thus, it was found that the band at 994 cm–1, which is commonly exclusively assigned to a V═16O1 stretching (Ag) vibration, also contains contributions of an 16O1–V–16O2 stretching vibration (B2g). If only the O1 position is exchanged, the B2g component shifts to 964.2 cm–1, while if both O1 and O2 are exchanged, a shift to 953.4 cm–1 is expected. In contrast, the Ag component shifts only to 955 cm–1, regardless of whether only the O1 position or all three oxygen atoms are exchanged. On this basis, it was found that oxygen exchange at 573 °C in absence of an alkane involves O1 and O3 atoms, whereas in the presence of propane all three oxygen atoms are exchanged. In the latter case, the overall exchange rate appears to be limited by bulk diffusion. At typical reaction temperatures for the oxidative dehydrogenation of propane between 320 and 430 °C, no exchange occurs in pure oxygen. In presence of ethane or propane, only O1 is partly exchanged possibly at the surface and/or in a near-surface region. Under the typical reaction conditions of oxidative dehydrogenation of propane at 400 °C, there is hardly any variation in the spectra, and the small changes observed after long times on stream only affect O1, which, considering the sensitivity of the measurement method, leaves open whether the Mars–van Krevelen mechanism is indeed the predominant reaction mechanism under the conditions of oxidative dehydrogenation of alkanes on V2O5.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2022 The Authors. Published by American Chemical Society

1. Introduction

Click to copy section linkSection link copied!

Activation of molecular oxygen on the surface of metal oxides is of fundamental importance in heterogeneous catalysis, (1−3) solid oxide fuel cells, (4) and gas sensor technology. (5) Direct oxidation of inert alkane molecules with gas-phase oxygen over metal oxide catalysts not only is of economic interest but poses a challenge in scientific understanding due to the complexity of the catalysts as well as the reaction networks. (6) The selective oxidation of alkanes requires both the activation of C–H bonds in the alkane molecule and the activation of molecular oxygen. Reactive oxygen species formed in the process of charge transfer between the hydrocarbon and the oxygen molecule include electrophilic or nucleophilic intermediates, respectively. The process of O2 dissociation and final incorporation of oxygen into the oxide lattice involves the polarized adsorbed O2 molecule, differently adsorbed charged molecular species (superoxide O2 and peroxide O22–), and charged atomic species (O) with electrophilic character. The reaction ultimately leads to the incorporation of nucleophilic O2– ions into the lattice (eq 1), (7)
(1)
where the symbols * stand for a free adsorption site, g for gas phase, s for surface, and l for lattice. It was discussed that surface and volume defects are important for both adsorption and dissociation of oxygen. (8−13)
Control over the occurrence and distribution of the different oxygen species under the reaction conditions is key to achieving high selectivity for certain reaction products and avoiding complete combustion to unwanted CO2 during alkane oxidation. Different types of oxygen species are required depending on whether an olefin is to be formed by oxidative hydrogen abstraction or whether oxygenates are the desired products. The synthesis of oxygenates containing C–C double bonds requires the simultaneous presence of nucleophilic and electrophilic oxygen species, which can coexist at different locations on the catalyst surface. Ultimately, it is a matter of avoiding total oxidation to carbon dioxide, which is also the product of oxygen incorporation and C–C bond cleavage.
Experimental discrimination of oxygen species actually involved in hydrocarbon oxidation is complicated due to the elevated reaction temperatures that lead to rapid interconversion of the intermediates, which appear in eq 1. (1) Furthermore, the low concentration of active sites, (14,15) the limited sensitivity of suitable spectroscopic techniques such as vibrational (16) and EPR spectroscopies, or the superposition of a multitude of oxygen species in photoelectron spectra of working catalysts (17−19) pose additional challenges. (20,21)
Isotope exchange experiments were used to investigate oxygen exchange between the gas phase and oxides under various conditions. (22,23) Three different types of oxygen exchange reactions are distinguished in the postulated mechanisms. The R0 mechanism does not involve lattice oxygen (M–OL), but gas-phase 16O2 exchanges one oxygen atom with gas-phase 18O2 under formation of mixed 16O18O, while the atomic fraction of the two oxygen isotopes in the gas phase remains constant. In the R1 and R2 exchange mechanisms, one or two 16O lattice oxygen atoms, respectively, participate under formation of 16O18O and 16O2, respectively, in the presence of 18O2 in the gas phase. Winter, (24,25) Nováková et al., (22,26) and Boreskov (27) derived expressions to distinguish the exchange type based on the exchange rate. V2O5 was found to follow a combination of R1 (in the lower temperature range of 480–540 °C) and R2 (in the higher temperature range of 520–600 °C) mechanism. (25) In other works, only the R2 mechanism was found above 450 °C. (28,29) Over supported vanadium oxides catalysts the R2 activity decreases in the order V2O5/TiO2 > V2O5/Al2O3 ∼ V2O5/SiO2, which corresponds to the order in which the activity in butane oxidation decreases. (28,30) However, the R2 activity cannot be solely attributed to surface vanadia species, because 17O NMR provided evidence for participation of the oxygen from the support in the exchange mechanism, (31) suggesting that oxygen exchange rate and reactivity are not necessarily linked. Based on density functional theory (DFT) calculations, the involvement of a surface ozonide complex O3 in the R1 mechanism, and a 4-atom complex with the O2 and O22– surface species in the R2 mechanism over VOx/TiO2, respectively, were proposed. (32,33) Some authors indicated that the presence of a reducing gas or the degree of reduction of the oxide has an impact on the oxygen exchange. (29,34)
Kinetic measurements of oxygen exchange, however, do not disclose which lattice oxygen is involved in the exchange. Three different types of oxygen atoms can be distinguished in bulk V2O5: (1) vanadyl oxygen O1; (2) 2-fold-coordinated oxygen O2; (3) three-coordinated oxygen O3 (Figure 1). The combination of oxygen exchange with vibrational spectroscopy provides the opportunity to distinguish the type of oxygen that participates in the exchange mechanism. In addition, another disadvantage of the purely kinetic studies of oxygen exchange is eliminated, since the vibrational spectroscopic experiments can usually be carried out in situ and not in vacuum and in the absence of alkanes.

Figure 1

Figure 1. Unit cell of V2O5: O1, terminal oxygen; O2, bridging oxygen; O3, chain oxygen.

In early studies, Hirota and Kera et al. investigated the oxygen exchange between gaseous C18O2 or 18O2 over V2O5 by infrared spectroscopy. (35−37) A new band appearing at 962 cm–1 was ascribed to the V═18O1 stretching in contrast to V═16O1 at 1019 cm–1, however, the low resolution of the infrared spectra complicated the identification of other vibrational modes, which might be simultaneously present.
Raman spectroscopy is a versatile tool in analyzing the interaction of O2 with metal oxides, because the stretching frequency of molecular oxygen species, which can be measured directly, decreases with decreasing bond order caused by increasing charge transfer from 1556 cm–1 for the free oxygen molecule (38) to 1200–1000 cm–1 for superoxide and 1000–800 cm–1 for peroxide species adsorbed on metal oxide surfaces. (39−41) Furthermore, Raman spectroscopy in combination with theory and isotope exchange experiments has been used to analyze the molecular structure of supported vanadium oxide species. (42−47)
Oyama et al. investigated the oxygen exchange over silica-supported vanadium oxide catalysts with different vanadia loadings that contained mixtures of surface vanadia species characterized by a vanadyl stretching vibration at 1042 cm–1 and segregated crystalline V2O5 particles exhibiting the typical V═16O1 mode of bulk V2O5 at 997 cm–1 by Raman spectroscopy. (48) The silica-supported vanadyl groups were exchanged at 347 °C, whereas the vanadyl groups in the V2O5 crystallites were not affected. The analysis was not trivial because the 18O substitution of the silica-supported vanadyl species shifted the original V═16O band at 1042 cm–1 to 997 cm–1, the same position as that for the bulk V═16O1 vibrations in the segregated crystalline V2O5 particles.
The shift was confirmed in studies of highly dispersed vanadium oxide species supported on zirconia and silica, respectively, investigating catalysts that do not exhibit spectroscopic features of bulk V2O5 particles. (42,43) It is, however, problematic to assign the bands of silica-supported vanadium oxide in the range of 1000–1030 cm–1 exclusively to vanadium oxide stretching vibrations because of strong coupling of the V–O modes with the modes of Si–O–V interphase bonds, (45) which also complicates the assignment of the bands due to exchange with 18O2. (44)
Ono et al. investigated oxygen exchange in the presence of n-butane and propane at 430–520 °C on ill-defined V2O5/SiO2 catalysts that exhibited a Raman spectrum, which was entirely dominated by segregated V2O5 crystallites. (49,50) The bands of V2O5 at 998 and 703 cm–1 were shifted to 964 and 685 cm–1, respectively, upon dosing of 18O2. Oxygen atoms at the V–O3 site (703 cm–1) were found to exchange easier than that at the V═O1 site (998 cm–1) in the presence of n-butane, while in the presence of propane, exchange at the V═O1 site was easier. The difference was ascribed to differences in crystal orientation, lacking any convincing evidence due to the ill-defined, polycrystalline structure of the catalysts.
Taking all this together, it can be said that Raman spectroscopic studies of oxygen exchange on vanadium pentoxide have so far only been carried out on heterogeneous systems. A clear assignment of the Raman bands was therefore not possible. Thus, the changes in the spectra caused by the oxygen exchange under different conditions were difficult to interpret. This motivated us to investigate crystalline V2O5 again in depth with Raman spectroscopy. DFT calculations were used to support the assignment of the Raman modes. The spectral changes resulting from the partial exchange of 16O atoms on the surface and in the bulk of V2O5 with 18O2 from the gas phase at different temperatures and in the presence of different gas atmospheres could thus be clearly elucidated. These studies are model investigations for processes on vanadium-containing mixed metal oxides, which are known for their high performance in alkane and electrocatalytic water oxidation. (51−53) The work contributes to a better understanding of oxygen activation and exchange on vanadium pentoxide, which is important for targeted design with respect to surface or bulk properties of oxidation catalysts. In addition, the study provides reference experiments for operando investigations of vanadium oxide-based catalysts.

2. Methods

Click to copy section linkSection link copied!

2.1. Properties of V2O5

According to X-ray diffraction (XRD) (Figure S1 in the Supporting Information and Table S1) and X-ray fluorescence spectroscopy (XRF) analysis, the studied V2O5 (internal ID: 22055) was phase pure (orthorhombic phase, space group Pmmn (No. 59), point group D2h, ICSD 60767; for lattice parameters, see Table S1), contained no impurities, and it had a specific surface area of 4.2 m2/g. V2O5 was thermally stable under Ar up to 669 °C followed by melting as indicated by an endothermic peak (Figure S2). Melting was accompanied by mass loss due to the evolution of oxygen and sublimation. STEM images (Figure S3) revealed a rod-like shape of the V2O5 particles with a size ranging from several hundred nanometers to 3 μm.

2.2. Basic Characterization of V2O5

The XRD measurements were performed in Bragg–Brentano geometry on a Bruker AXS D8 Advance II θ/θ diffractometer, using Ni-filtered Cu Kα1+2 radiation and a position sensitive energy dispersive LynxEye silicon strip detector. X-ray fluorescence spectroscopy (XRF) was used for elemental analysis of V2O5, applying a Bruker S4 Pioneer X-ray spectrometer. For sample preparation, the mixture of V2O5 (0.1 g) and lithium tetraborate (8.9 g, >99.995%, Aldrich) was fused into a disk using an automated fusion machine (Vulcan 2MA, Fluxana). Nitrogen adsorption was performed at −196 °C using the Autosorb-6B analyzer (Quantachrome) after outgassing the material in vacuum for 2 h at 200 °C with an external preparation unit (Autosorb Degasser). All data treatments were performed using the Quantachrome Autosorb software package ASWin. The specific surface area SBET was calculated according to the multipoint BET method in the p/po = 0.05–0.15 pressure range assuming an N2 cross sectional area of 16.2 Å2. Thermogravimetry (TG) and differential thermal analysis (DTA) were carried out on a Netzsch STA449 Jupiter thermanalyzer in 70 N mL/min Ar total flow applying a heating rate of 10 °C/min. Around 80 mg of the sample was placed into the corundum crucible (0.2 mL) without lid. The gas phase analysis during heating was monitored with a quadrupole mass spectrometer (Pfeiffer, QMS200 Thermostar). Evolution of O2 was recorded by monitoring the change of intensity of the ion currents (m/z = 32). The high angle annular dark field scanning transmission electron microscopy (HAADF-STEM) images were taken on an aberration-corrected JEOL JEM-ARM200F transmission electron microscope operated at 200 kV. The UV/vis spectrum of V2O5 was measured using a Cary 5000 UV–vis spectrometer (Agilent) equipped with a Praying Mantis diffuse reflection accessory (Harrick). The data were obtained in percent reflectance and then converted to the Kubelka–Munk function F(R). Spectralon was used as a white standard in the range between 800 and 250 nm.

2.3. Raman Spectroscopy

In a typical Raman measurement, polycrystalline V2O5 was either compressed on a glass slide for ex situ measurements or placed in the sample holder of a Linkam CCR1000 reaction cell (Linkam Scientific Instruments LTD), or a high temperature reaction chamber from Harrick Scientific Products Inc., respectively, for in situ measurements. Except for the experiments presented in Figure 7 using the Harrick Raman chamber, all the in situ measurements were conducted in the Linkam cell. The spectra were measured using a customized multiwavelength Raman spectroscopy system (Figure S4) assembled by S&I Spectroscopy & Imaging GmbH (Warstein, Germany). Nine different lasers (785, 633, 532, 488, 457, 442, 355, 325, and 266 nm) were used as excitation sources. Laser powers were tuned by neutral density filters (785 nm, 0.3 mW; 633 nm, 0.2 mW; 532 and 488 nm, 0.5 mW; 266 nm, 0.3 mW; all other lasers, 0.4 mW) to prohibit sample damage. Exposure times (a few seconds to 600 s) were optimized accordingly with regard to a high signal-to-noise ratio. We were aware that V2O5 is sensitive to UV lasers (325 and 266 nm), and long exposure times with relative high laser power could lead to progressive appearance of Raman features around 758, 896, 910, and 1009 cm–1 (Figure S5). The laser-induced damage is relevant to structural change due to the removal of vanadyl oxygen as evidenced by the disappearance of the Raman band at 994 cm–1. The entrance slit was set to 100 μm for all measurements. The scattered light was dispersed on 300, 600, or 2400 grooves/mm gratings depending on laser wavelength and resolution requirements, and finally collected by one of the two liquid-nitrogen-cooled CCD detectors (PyLoN:2K and PyLoN:100 from Princeton Instruments). The primary beam was eliminated by using corresponding edge filters. To compare Raman spectra measured by using different excitation lasers, the true Raman scattering intensities were extracted by correcting instrument effects and optical absorption properties of V2O5. A standard lamp (deuterium and halogen light source, Ocean Optics DH-2000) with a known emission spectrum was used to correct instrument effects (objectives, filters, gratings, detectors, etc.). In addition, the absorption of light by V2O5 was taken into account by the G(R) function that can be calculated from the measured diffuse reflectance spectrum of V2O5. (54) A detailed description of the Raman spectroscopy setup and the intensity correction is given in the Supporting Information (Figures S4 and S6–S8).

2.4. Oxygen Isotope Exchange

A 20–50 mg V2O5 sample was placed in the Linkam CCR1000 reaction cell unless otherwise mentioned and pretreated by heating the sample at a heating rate of 5 °C/min to the desired temperature under a flow of 20% 16O2 (99.999%) balanced by helium (or nitrogen) with a total flow rate of 10 mL/min. Then 16O2 was switched to 18O2 (97 at% enrichment from the supplier, Linde) to conduct oxygen exchange for 2 h. The fractions of 16O16O and 16O18O, respectively, that were present as impurities in the 18O2 gas cylinder were determined by mass spectrometry and were 1.2% and 2.3%, respectively (Figure S9). Other gases used were propane (Westfalen AG, 99.95%), deuterium labeled propane (Sigma-Aldrich, 99 atom % D), ethane (Westfalen AG, 99.95%), and n-butane (Westfalen AG, 99.95%). Raman spectra were measured continuously throughout the exchange process with each spectrum taking 6 or 8 min (532 nm). There is a second thermocouple inserted into the catalyst bed in the Linkam CCR1000 reaction cell to measure the actual temperature. For the Harrick Raman chamber, the heating element and the thermocouple are separated from the reaction part, and thus a calibrated temperature was given.

2.5. Operando Raman Spectroscopy

A 20 mg V2O5 sample was placed in the Linkam CCR1000 reaction cell to perform the operando Raman experiments. V2O5 was first heated from room temperature to 420 °C at a heating rate of 5 °C/min under 20% 16O2 balanced by helium with a total flow rate of 10 mL/min. After reaching 420 °C, the sample was further heated under various reaction feeds (C3H8/16O2/He). The effluent gas was connected to a micro-GC (Agilent 490) to perform online gas product analysis. A combination of MS5A (10 m length) and PPQ columns (10 m length), connected to a thermal conductivity detector (TCD), was used to analyze the oxidation products in the gas phase. The conversion of propane X and the product selectivity S were calculated based on the sum of the detected products. Raman measurements were performed simultaneously by using the 532 nm laser during propane oxidation over V2O5.

2.6. DFT Calculations

Phonon calculations were performed using Quantum Espresso 5.4.0 (55) to obtain vibrational frequencies. The crystal structure of V2O5 (Figure 1, 9012221.cif (56)) was obtained from the Crystallography Open Database (COD, Web site: http://www.crystallography.net/cod/index.php), which was used as the initial input structure. The structure was optimized by variable cell relaxation calculations (Tables S1 and S2). The calculations were carried out by using the Perdew–Burke–Ernzerhof (PBE) functional and ultrasoft pseudopotentials. A van der Waals-inclusive correction (DFT-D) (57,58) was applied to take dispersion forces between layers into account. The plane wave kinetic energy cut off values of 60 and 480 Ry were adopted for V2O5 for the wave functions and the charge densities, respectively. A Methfessel–Paxton smearing of 0.01 Ry was used to improve calculation performance. Here, 6 × 6 × 6 Monkhorst–Pack grids of k-point sampling was chosen for V2O5. (59)

3. Results and Discussion

Click to copy section linkSection link copied!

3.1. Experimental and Calculated Raman Spectra of V2O5

Generally, the Raman spectrum of V2O5 can be divided into (i) the high energy range from 1000 to 500 cm–1 where V–O stretching vibrations arise, (ii) the medium energy range from 500 to 300 cm–1 where V–O–V bending vibrations arise, and (iii) the low energy range ν ∼ <300 cm─1 where collective vibrations of (V2O5)n units arise. Based on the symmetry of the crystal structure (space group Pmmn (No. 59), point group D2h), analysis using the Bilbao crystallographic server (Web site: http://www.cryst.ehu.es/) results in 39 optical modes
in which 21 modes are Raman active
According to the light absorption properties of V2O5 in the UV/vis range (Figure 2a), the experimental Raman spectra of V2O5 recorded using nine different lasers can be grouped into off-resonance Raman spectra (785 and 633 nm), one preresonance Raman spectrum (532 nm), and resonance Raman spectra (488 to 266 nm). As shown in Figure 2b, 10 distinct bands were resolved experimentally. Note that the band at 102 cm–1 is not observed with the UV lasers (266, 325, and 355 nm) due to cutting off by the edge filter. The excitation wavelength has no significant effect on the band position. The peak maximum at 1001.4 ± 0.54 cm–1 of the standard polystyrene was deviating by less than 3 cm–1 with all lasers (Figure S7). Raman shifts indicated in Figure 2b were determined based on the spectrum of V2O5 measured with the 532 nm laser. In addition, six weak bands were observed at 1025, 962, 800, 660, 356, and 232 cm–1 with the 785 nm laser (Figure S10, parts b and c).

Figure 2

Figure 2. UV/vis spectrum (a) and Raman spectra (b) of V2O5 collected at room temperature (ex situ). The laser wavelength used for the Raman measurements is shown next to the Raman spectra. The Raman spectra were corrected with respect to instrumental effects, taking into account the known response curve of a white lamp and the absorption spectrum of V2O5 (see Figure S8 for details of the correction procedure and Figure S10 for uncorrected spectra); All spectra were normalized to the corresponding maximum band intensity ([0,1]), which differs for the different excitation energies; The band positions indicated were determined for the spectrum measured with the excitation wavelength 532 nm; The intensity ratio of the band at 994 cm–1 to the band at 144 cm–1 as a function of excitation wavelength is plotted in part a.

With the increase of photon energy (decrease of excitation wavelength), the intensity ratio of the band at 994 cm–1 to the band at 144 cm–1 increases and exhibits a maximum at 355 nm (Figure 2a) suggesting resonance enhancement of the band at 994 cm–1 by using excitation energies in the UV range. (60) However, the probing depths changes with the excitation wavelength as well. Therefore, the change in the intensity ratio may also reflect structural gradients within the depth profile. UV lasers are more surface sensitive. (61) Consequently, the relative low intensities of bands at lower wavenumbers in the spectra recorded with the UV lasers, which are due to long-range order, can be interpreted in terms of restructuring and formation of a near surface region, which exhibits less long-range order.
Various experimental and theoretical studies contributed to an assignment of the peaks observed in the Raman spectrum of V2O5. (62−66) However, some inconsistencies occur in the literature. Therefore, phonon calculations were performed to refine the assignment. The calculated lattice parameters of the optimized structure agree very well with the lattice parameters of the investigated V2O5 material determined by XRD (Table S1), indicating that the model represents reality quite well. The atomic coordinates of the optimized calculated V2O5 structure are given in Table S2. Experimentally determined peaks measured by using the 532 nm laser are compared with calculation results of the present and a previous study (66) in Table 1 and Figure S11. In Figure S11, the atomic displacement patterns of all calculated Raman active phonon modes are shown. The peak positions calculated in the present work agree well with the experiment. The largest deviation is 14 cm–1. However, not all calculated modes are resolved experimentally (resolution ∼0.5 to ∼7 cm–1, please see Figure S6). Each of the six bands found at 197, 283, 303, 481, 700, and 994 cm–1 is caused by two different calculated modes (Table 1). The most pronounced band at 144 cm–1 is composed of three different modes calculated at 133.7, 142.5, and 144.4 cm–1. Two modes calculated at 214.2 and 944.7 cm–1 were not observed experimentally, most likely due to low Raman intensity.
Table 1. Experimentally Determined (Excitation Wavelength 532 nm) Peaks, Calculated Raman Active Phonon Modes (Unit: cm–1), and Vibrational Assignment
modeexptlacalcdb V216O516O1–16O2–16O3calcdccalcdb V216/18O518O1-16O2-16O3calcdb V218O518O1-18O2-18O3assignment
Ag10296.6107.794.793.0Tz translation
 197192.3192.8184.7181.2skeleton bending
 303303.8294.1297.4286.5Rx libration
 404400.8381.7396.9383.6skeleton bending
 481467.3527.8466.3447.2O2–V–O3 bending
 527526.1542.5525.4512.0O2–V–O3 bending
 994998.11028.9955.0955.0V═O1 stretching
B1g144142.5167.3138.6138.4Rz libration
 283280.6276.7270.7270.5O1–V–O3 bending
 700691.9772.3691.0653.0V–O3 stretching
B2g144133.7146.0131.4129.4Ry libration
 197191.8199.8184.9183.1Tx translation, Ry libration
 303307.1298.8299.5291.3Ry libration
 356350.7361.3346.2340.1skeleton bending
 481482.4528.7480.7454.4displacement of O3
 n.d.d944.71010.3935.6900.7V–O2 stretching
 994997.61030.1964.2953.4O1–V–O2 stretching
B3g144144.4168.7140.5140.2Ty translation
 n.d.214.2218.6214.9203.0Rx libration
 283281.9279.0272.4271.6O1–V–O2 bending
 700691.5772.4690.5652.6V–O3 stretching
a

Experimental Raman spectra are shown in Figure 2b, the peak positions were determined using the 532 nm laser.

b

Calculated Raman active phonon modes, this work

c

Calculated Raman active phonon modes from reference. (66)

d

n.d.: not detected.

Table 1 provides the assignment of all experimental modes according to the present DFT calculations. The band at 994 cm–1 is frequently exclusively attributed to the V═O1 stretching mode. (65) Our results show that the 994 cm–1 band has noticeable contribution from O1–V–O2 stretching, hence, it involves O1 and O2 atoms.
The band observed at 527 cm–1 was assigned to V–O2–V symmetric stretching by Abello et al. (63) In contrast, the same band was attributed to V–O3 bond stretching and V–O3–V angle deformation by Clauws et al. (64) and Brázdová et al. (65) Based on the present calculations (Table 1), the mode involves both O2 and O3, and it is related to O2–V–O3 bending, which is in agreement with Zhou et al. (66)
Major inconsistencies exist concerning the involvement of the three different oxygen atoms in the V2O5 structure to the bands observed at 404 and 481 cm–1, which were ascribed to O2 and O3 atoms displacements, respectively, by Abello et al. (63) Clauws et al., however, ascribed the bands to O1 and O2 atoms displacements, respectively. (64) In turn, Brázdová et al. proposed the two bands to O3 and O2 atom displacement, respectively. (65) Here we assign the band at 404 cm–1 to skeleton bending involving all three oxygen atoms and the vanadium atom (Table 1). The band at 481 cm–1 originates from O2–V–O3 bending and the displacement of the O3 atom (Table 1).
The four weak and broad bands located at around 1025, 962, 800, and 660 cm–1, which were clearly observed by using the infrared laser (785 nm) (Figure S10b), have been tentatively assigned to overtones or combination bands in previous studies. (63,64) In contrast, Brázdová et al. (65) assigned the band at 962 cm–1 to a fundamental band (V–O2–V stretching, B2g mode). A band ascribed to V–O2 stretching was predicted at 944.7 cm–1 in our study (Table 1), which, however, deviates considerably from the experimentally observed band at 962 cm–1. The band at around 1025 cm–1 may arise from surface VOx species self-supported on V2O5, because VOx species supported on oxides similarly exhibit a broad feature in the range between 1012 and 1027 cm–1. (67) The assignment of the band observed at 232 cm–1 with the 785 nm laser excitation (Figure S10c) remains unclear so far. The closest calculated value of 214.2 cm–1 is due to a collective vibration of the V2O5 lattice.
The assignment of the bands is an important prerequisite for the interpretation of the shift of bands observed during isotope exchange.

3.2. Oxygen Isotope Exchange

Isotope exchange of oxygen on the surface and in the bulk of polycrystalline V216O5 with 18O2 in the gas phase was studied by Raman spectroscopy at different temperatures and chemical potentials. The temperatures of ca. 320, 430, and 570 °C applied in the present study were chosen below, at and above typical reaction temperatures in selective oxidation of alkanes over vanadium oxide based catalysts, respectively. All temperatures are distinctly above the Tammann temperature TTam(V2O5) of 204 °C (477 K). At the Tammann temperature, TTam ≈ 0.5Tmelt [K], atoms or ions become sufficiently mobile and migration from the bulk to the surface and surface liquefaction are possible. The exchange was probed in synthetic air and in the presence of alkanes and the type of oxygen atom exchanged was concluded based on the results of the DFT calculations.

3.2.1. Prediction of Raman Shifts in Partially and Fully Exchanged V2O5 by DFT Calculations

The changes in the Raman spectrum of V2O5 were predicted by DFT calculations for two cases:
(i)

only 16O1 is exchanged by 18O

(ii)

all three types of oxygen atoms are exchanged by 18O

The calculated peak positions for the two cases are listed in Table 1.
(i)

The exchange of the O1 atom will have an impact on the position of the experimentally determined band at 994 cm–1, which contains contributions of two different phonon modes as explained above. The Ag mode is related to O1 (V═O1 stretching) and the B2g mode is related to O1 (major contribution) and O2 atoms (minor contribution) in an O1–V–O2 stretching vibration (Table 1). The calculated values of Ag at 998.1 cm–1 and B2g at 997.6 cm–1 in V216O5 differ slightly and are experimentally not resolved, giving rise to only one experimentally observed band at 994 cm–1 in V216O5. However, when 16O atoms are exchanged by 18O only at the V═O1 site, the calculated Ag and B2g modes shift differently by 43.1 cm–1 to 955.0 cm–1 and by 33.4 cm–1 to 964.2 cm–1, respectively (Table 1). Hence, the appearance of two new peaks at lower energy is expected. It should be possible to resolve these peaks experimentally due to the difference of about 10 cm–1. In addition, the O1 atom contributes to the B1g mode with an O1–V–O3 bending vibration (calculated shift at 280.6 cm–1) and to the B3g mode with an O1–V–O2 bending vibration (calculated shift at 281.9 cm–1). The two modes are giving rise to the experimentally observed band of V216O5 at 283 cm–1 (Table 1). The calculated Raman shifts are moved by 9.9 and 9.5 cm–1, respectively, to 270.7 and 272.4 cm–1, respectively, when just the O1 atom is exchanged. Hence, it is expected that this band will not split into two components in the experimental spectrum and will shift only slightly to lower energy. In particular, when the oxygen exchange is not complete, the shift will be negligible.

(ii)

When all types of oxygen atoms are exchanged to 18O, all bands are shifted to lower energies and are different from case i. Ag and B2g modes (994 cm–1) shift in a similar manner by ca. 44 cm–1 to 955.0 and 953.4 cm–1, respectively.

3.2.2. Experimentally Observed Changes in the Raman Spectra

3.2.2.1. Temperature Effect in the Presence of Oxygen Only
No change in the Raman spectrum was detectable when a flow of 20% 16O2 in helium was exchanged by a flow of 20% 18O2 in helium at 431 °C (Figure 3).

Figure 3

Figure 3. Raman spectra of V2O5 measured at room temperature before and after isotopic oxygen exchange; Gray and red spectra denote the cases before and after exchange at 322, 431, and 573 °C, respectively, using 18O2 (20% in He) or a mixture of propane and oxygen (C3H8 (1%) + 18O2 (19%) in He) for 2 h. All the spectra were normalized to [0,1]. The positions of all spectra were aligned with respect to the band at 994 cm–1. Laser: 532 nm. Heating rate: 5 °C/min. Total flow rate: 10 mL/min.

At 573 °C, however, the peak at 994 cm–1 lost intensity. Taking into account the results of the DFT calculations of partially (O1) and fully (O1, O2, O3) exchanged V2O5, the Raman spectra recorded at room temperature after the isotope exchange experiments at different temperatures were deconvoluted to get insight into the contributing components in a qualitative manner (Figure 4). The deconvolution procedure is explained in detail the Supporting Information. The fitting parameters are provided in Table S3.

Figure 4

Figure 4. Deconvolution of Raman spectra collected at room temperature after treatment at different temperatures and in different gas atmospheres (the same Raman spectra as shown in Figure 3) in the V–O3 (a) as well as V═O1 and O1–V–O2 stretching vibration region (b). Temperatures and gas atmospheres are specified in the right top corner of each section. (c) Corresponding representations of vibrational motions of phonon modes.

Figure 4a, middle, shows the V–O3 stretching vibrations, whereas Figure 4b, middle, shows V═O1 and O1–V–O2 stretching vibrations after exchange in 20% 18O2 in helium at 573 °C. Two components at higher energy (features >1000 cm–1 assigned above to surface VOx species) become visible now at 1007 and 1029 cm–1, which might be attributed to V═18O(s) and V═16O(s), respectively (Figure 4b, middle). The two components appear also after exchange in the presence of propane (see Section 3.2.2.2 and Figure 4b, top). Interestingly, the peak width of V═16O(s) species in terms of fwhm (full width at half-maximum) is much larger than that of V═18O(s) (∼50 vs ∼15 cm–1, Table S3). The large peak width in the spectrum before oxygen exchange reflects the heterogeneity of the VOx species at the surface. This may mean that the reactivity of the energetically different surface species is also different. The relatively small peak width of V═18O(s) suggests that only a portion of the V═16O(s) species is reactive and has been exchanged. Assuming an isotope ratio of 1.0452, one would expect a band at ∼1050 cm–1 for this reactive V═16O(s) species. The exposure of the V2O5 to C3H8 + 16O2 feed alone does not cause any spectral modification (Figure S12), suggesting that the observed change in the presence of oxygen at high temperature is indeed due to a partial oxygen exchange and not, for example, due to the formation of defects.
In addition, a new band appeared at around 962 cm–1 (18O1–V–16O2) with a shoulder around 954 cm–1 (V═18O1) (Figure 4b, middle). The deconvolution of the original spectra is shown in the Supporting Information, Figure S13. The band at 702 cm–1 moved to 690 cm–1, suggesting a partial exchange of the O3 atom, which affects the V–O3 stretching vibration (Figure 4a, middle). Shifts of bands below 530 cm–1 are hardly visible. The observations suggest that oxygen atoms at the V═O1 and V–O3 sites were partially exchanged in the presence of 20% 18O2/He at 573 °C.
The temporal evolution of spectra recorded in situ during the experiment at 573 °C is shown in Figure 5a. The gradual increase of the new band around 962 cm–1 is clearly observed reflecting the increase of the exchange extent. The band positions of original bands are not affected, showing that the extent of exchange (O1 atoms) has no influence on the band position.

Figure 5

Figure 5. In situ Raman spectra of V2O5 at 573 °C under 20% 18O2 in He (a), 1% C3H8 + 19% 18O2 in He (b), and 1% C3D8 + 19% 18O2 in He (c) as a function of time. Conditions of exchange experiments are described in the caption of Figure 3. The numbers show the time that had passed since the start of the experiment at the moment of recording. Laser: 532 nm.

3.2.2.2. Oxygen Isotope Exchange in the Presence of Propane
The following experiment explores how the chemical potential of the gas phase influences the oxygen exchange at different temperatures. For this purpose, propane was added to the oxygen flow. The presence of propane, which undergoes oxidation under the applied conditions (propane conversion X = 36% at 573 °C), will change the defect concentration on the surface and in the bulk of the vanadia. A low ratio of propane to oxygen (1:19) was used to prevent severe reduction of V2O5, in particular at the high reaction temperatures.
When a flow of C3H8:18O2:He in a ratio of 1:19:80 was applied at 573 °C, the intensity of the band at 994 cm–1 decreased significantly and all other bands moved to lower wavenumbers (Figures 3, 4a, 4b(top), and 5b). A new band appeared at around 954 cm–1. This is different from the case using 20% 18O2 in helium (Figures 4a, 4b(middle), and 5a), where the development of a band at 962 cm–1 with a shoulder around 955 cm–1 was observed. Specifically, the extent of exchange of O3 atoms is higher than under only oxygen flow as indicated by the larger shift of the position of the peak near 700 cm–1 and the appearance of a component of full V–18O3 in this band (Figure 4a(top)). The deconvolution analysis further reveals that the newly formed band at 954 cm–1 possibly contains three components from different combinations of exchanged O1 and O2 atoms due to V–O1 and O1–V–O2 stretching vibrations (Figure 4b(top)). The presence of a component corresponding to 18O1–V–16/18O2 (coupling of O2 atoms) suggests a faster exchange of O1 atoms with 18O as compared to O2 atoms. Bearing in mind the calculated Raman spectrum of V218O5 (Table 1), the results suggest that essentially all three types of oxygen atoms were exchanged in the presence of propane at 573 °C. In a control experiment, 16O2 instead of 18O2 was used (Figure S12). According to that, propane oxidation at this high reaction temperature does not change the structure of V2O5 so drastically as to affect the Raman spectrum. Consequently, the change in band positions presented in Figure 3 is due to 18O incorporation into V2O5. These results directly prove that the presence of a reductant accelerates exchange rates. (29,34) Assuming the R2 exchange mechanism over V2O5, two surface vacancies are required to make exchange happen. (28) It is likely that propane effectively generates vacancies on the surface of V2O5 and greatly increases the probability of having two neighboring vacancies. DFT calculations also imply that reduced surfaces of metal oxides are energetically more favorable for oxygen activation than unreduced surfaces. (40) It was pointed out that the formation energy of surface oxygen vacancies is reduced by the presence of surface OH groups due to a weakening of the V–O bond in V–OH. (68) To verify this, deuterium labeled propane was used in the isotopic oxygen exchange experiment. The exchange proceeds indeed at a slightly lower rate in the case of using deuterium labeled propane as compared to normal propane (Figure 5c). The exchange extent at each time step is lower than that using normal propane. The observed isotope effect suggests that hydrogen is involved in the oxygen activation process, possibly by reducing the energy to form oxygen vacancies at the surface.
To further quantify the exchange process, the exchange extent ΔA,
(2)
where A refers to the band area, as a function of time (Figure 5) was analyzed at 573 °C in the different gas atmospheres (Figure 6). The analysis does not allow the differentiation between the three types of oxygen atoms. Instead, the total extent of exchange is estimated. In general, isotopic oxygen exchange comprises surface exchange and bulk diffusion. (69−71) Surface exchange takes place at the interface between the solid and the gas. The propane molecule would also be involved in the surface exchange by reacting with labeled surface oxygen species and generating oxygen vacancies. Bulk diffusion of oxygen can proceed through vacancies or interstitial sites. (72) In the following, the two borderline cases
(3)
and
(4)
where kex refers to the surface oxygen exchange coefficient, D is the oxygen diffusion coefficient, and a is the radius of the spherical catalyst particles, have been considered. The assumption of spherical particles is a rather crude simplification in view of the plate-like nature of the V2O5 particles (Figure S3). However, a comparison of relative changes is possible. The detailed derivation of eqs 3 and 4 and the fitting parameters (Table S4) are provided in the Supporting Information. As Figure 6 shows, oxygen exchange at 573 °C remains limited to the surface in the absence of propane (Figure 6, black lines and data points). Studies of oxygen exchange on amorphous V2O5 at 400–550 °C have shown that, in the absence of a reducing agent, the rate of exchange is controlled by the surface reaction. (73) In the presence of propane, however, the overall exchange rate appears to be limited by the bulk diffusion (Figure 6, red and blue lines and data points). Apparently, propane at 573 °C accelerates oxygen exchange at the surface, so that oxygen diffusion in the bulk becomes a comparatively slow process.

Figure 6

Figure 6. Extent of oxygen exchange as a function of time; The lines are the fitted results using eqs 3 (dashed) and 4 (solid). The experimental data points were taken from the experiments shown in Figure 5. Note that the time used in the graph is the average time taken to record each spectrum.

The main bands positions were not shifted when the same experiment was performed at 431 °C (Figures 34a, and 4b(bottom)). The temperature 431 °C is a typical reaction temperature applied in alkane oxidation over V-based catalysts. Just a small intensity difference near 962 cm–1 was found indicating that the oxygen exchange is just detectable by Raman spectroscopy at this temperature and mainly V═O1 sites were exchanged (Figure 4b, bottom). In contrast, oxygen exchange at the same temperature in pure oxygen could not be detected by Raman spectroscopy. Due to the small extent of exchange at 431 °C, it was not possible to perform a kinetic analysis in analogy to the experiment at 573 °C (Figure 6). However, since oxygen diffusion in the bulk is even more difficult at lower temperatures, it is very likely that this process does not play a role in the lower temperature range.
Finally, the exchange in the presence of propane was also performed at 322 °C. No change in the Raman spectrum was observed under these conditions after 2 h (Figure 3).
The direct comparison of the present in situ Raman experiments with kinetic studies of isotopic oxygen exchange by mass spectroscopy in the past is not possible, as these experiments were carried out in vacuum at low oxygen partial pressures of a few millibar, (28,29) which were lower than or in the order of the magnitude of the equilibrium oxygen partial pressure that develops above V2O5 at elevated temperatures. (74−76) Online mass spectra recorded during a temperature programed experiment in the Raman cell in the presence of 20% 18O2 in He show that the onset temperature for observing mixed 16O18O in the gas phase was ∼424 °C (Figure S9), and for 16O2 at ∼450 °C, which is in agreement with the observation that no oxygen exchange was observed by Raman spectroscopy at 431 °C in 20% 18O2 in He (Figure 3). The Raman spectroscopy appears a little less sensitive than the mass spectroscopy. Furthermore, an increase in the relative concentration of 16O atoms in the gas phase oxygen also occurred at ∼424 °C, which confirms that the oxygen exchange between the gas phase and the solid starts at temperatures higher than 420 °C in absence of a reducing agent, because in the case of a full gas-phase R0 mechanism a constant atomic ratio of 16O would be expected. The formation of 16O18O further implies that oxygen dissociation (eq 1) on V2O5 is reversible, which is different from a previous work, in which no formation of 16O18O was found from C3H816O218O2 mixtures on V16Ox/Zr16O2 at about 427 °C. (77)
3.2.2.3. Influence of the Hydrocarbon Chain Length
The effect of the nature of the alkane on the isotope oxygen exchange was examined at 400 °C applying an alkane:oxygen ratio of 10:5. These conditions correspond to the typical temperature and feed composition used in the oxidative dehydrogenation of alkanes. Ethane, propane, and n-butane were compared. In case of ethane and propane, a small band near 962 cm–1 occurs after 2 h under these conditions, while all the other bands maintain their positions (Figure 7a). Apparently, mainly a small fraction of V═O1 sites was exchanged. By using n-butane, a new band at 946 cm–1 is already formed after 2 min (Figure 7b). No change in the Raman spectra was observed after such a short time in the experiments with ethane and propane. As compared to the spectrum obtained under 16O2 at the same temperature, the band at 700 cm–1 (shifted to 693 cm–1 due to thermal effect) maintains its position as well as other bands. V2O5 was fully reduced, however, after only 12 min, and no clear features can be found in the in situ Raman spectra anymore (Figure 7b). The more reducing conditions in n-butane lead most likely to a higher concentration of surface oxygen vacancies. The newly formed band at 946 cm–1 is attributed either to isotope oxygen exchange (see band at 953.4 cm–1 for 18O1–V–18O2 stretching in Table 1) or to an intermediate (suboxide) during reduction of V2O5.

Figure 7

Figure 7. Effect of the nature of the alkane on isotope oxygen exchange. (a) Raman spectra of V2O5 measured at room temperature before and after isotopic oxygen exchange experiments under 10% C3H8 + 5% 18O2 and 10% C2H6 + 5% 18O2, respectively, at 400 °C. (b) In situ Raman spectra of V2O5 at 400 °C under 16O2 and 10% C4H10 + 5% 18O2 as a function of time. Numbers on the spectra indicate the starting time of measurement. Asterisks represent cut cosmic ray signals. Raman spectra were normalized to [0,1]. Laser: 532 nm. Exchange experiments were conducted at 400 °C for 2 h in the Harrick Raman chamber. Heating rate: 5 °C/min. Total flow rate: 10 mL/min.

4. General Discussion and Conclusions

Click to copy section linkSection link copied!

V2O5 was studied using multiwavelength Raman spectroscopy. Reassignment of the experimentally observed active Raman vibrational modes using DFT calculations revealed that the band at 994 cm–1, which in previous studies was attributed exclusively to the vanadyl oxygen stretching vibration (V═O1), contains additional contributions from O2 atoms. Based on the present combined experimental and theoretical approach, a method was developed to unambiguously determine the nature of the oxygen atom in the crystal structure of V2O5 (O1, O2, or O3) exchanging with 18O2 in the gas phase. The new assignment makes it possible to determine experimentally which type of oxygen atom can be exchanged under which reaction conditions. The exchange position depends strongly on the temperature and the chemical potential in the gas phase. An overview of the cases studied in the present work is given in Table 2. The conditions investigated are important for the selective and total oxidation of alkanes.
Table 2. Types of Oxygen Atoms in V2O5 Exchanged at Different Temperatures and in Different Gas Atmospheres
 gas phase composition (vol %)temperature of oxygen exchange (°C)
 alkaneO2inert573431400322
O2/He02080O1, O3n.d.ab
C2H6/O210585O1
C3H8/O211980O1, O2, O3O1n.d.
10585O1
C4H10/O210585O1, O2c
a

n.d.: not detectable by Raman spectroscopy.

b

–: not analyzed.

c

Observation difficult due to superimposed reduction of V2O5

For example, in the presence of propane, but under quite oxidative conditions (1% C3H8, 19% O2), only O1 oxygen atoms are exchanged at 431 °C. This temperature is in the range where selective oxidation reactions of alkanes are typically carried out. Using eq 2, a degree of exchange of ∼6.8% is measured by Raman spectroscopy after 2 h under these conditions. Since the probing depth of Raman spectroscopy is limited and ranges between a few and several hundreds of nanometers depending on the energy of the exciting laser, the overall exchange extent of the entire sample is much smaller than this value. The very low degree of exchange, which was only detectable after prolonged reaction time (2 h), indicates that the oxygen isotope exchange occurs mainly at the surface and/or in a near-surface region under the applied conditions. With increasing temperature in an oxidative atmosphere or in the presence of reducing agents, O3 sites and eventually O2 sites are exchanged. All types of oxygen (O1, O2, and O3) are involved in the exchange at very high temperatures (573 °C) and in the presence of propane in the feed. Under these conditions, a kinetic analysis utilizing spectral changes as a function of time was possible. The analysis shows that oxygen vacancies generated by propane facilitate the incorporation of gaseous oxygen into the lattice of V2O5 by accelerating the surface exchange process. Hence, rapid oxygen exchange at 573 °C in the presence of propane appears to be limited by bulk diffusion.
The exchange of oxygen in V2O5 becomes energetically increasingly difficult in the following order: O1 < O3 < O2. The terminal vanadyl defects are the easiest to form in agreement with predictions by theory. (9,78) Under typical reaction conditions of propane oxidation (400–450 °C), mainly vanadyl oxygen O1 is involved in the oxygen exchange and the very low exchange extent suggests that the exchange occurs primarily on the surface and/or the near surface region. With a classical feed composition (C3H8:O2 = 2:1) and temperature (400 °C) of the oxidative dehydrogenation of propane, hardly any change is seen in the spectrum and this small change only affects O1. In this temperature range, the working catalyst in the Raman cell produces propene by oxidative dehydrogenation of propane and CO2 as the product of total oxidation (Figure S14). The selectivity to propene under these conditions and at low conversions, which are typically set for kinetic studies, is between 55 and 60%. The catalyst was working in steady-state (Figure S14), and consequently, no changes in the Raman spectra are observed with time on stream. Also under other reaction conditions, no changes were detected after a total reaction time of 7 h (Figure S14).
In vanadium oxide based systems, it is commonly assumed that the formation of propene and other oxidation products occurs according to the Mars–van Krevelen mechanism, (79) which would involve consumption of surface oxygen atoms by incorporation of surface or lattice oxygen into water or oxygen-containing oxidation products. As far as propane oxidation on vanadium oxide catalysts is concerned, these conclusions are mainly based on kinetic modeling, (77,80) or DFT calculations at 0 K. (81) In the present work, the basis for the assignment of the Raman bands of V2(16O,18O)5 was provided. The isotope exchange experiments presented here cast doubt on whether the Mars–van Krevelen mechanism prevails under typical conditions of selective oxidation of propane, since a very small change in the spectra was observed only after 2 h of operation at 400 °C, indicating some exchanged O1 atoms (Figure 7a), while the catalyst produced propylene and CO2 at the steady state long before these changes were observed in the Raman spectra (Figure S14a). However, due to the limited sensitivity of Raman spectroscopy, further systematic operando investigations in combination with mass spectrometry are needed here to obtain more clarity.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcc.2c00174.

  • XRD pattern of V2O5, TG and DTA analysis of V2O5, HAADF-STEM images of V2O5, description of multiwavelength Raman spectroscopy, mass spectra of oxygen exchange of V2O5, uncorrected Raman spectra of V2O5, the atomic displacement patterns of Raman active phonon modes, Raman spectra of V2O5 after treatment in the presence of C3H8 and 16O2 at 573 °C, table of calculated lattice parameters and bond lengths, table of atomic coordinates of the optimized structure, description of the deconvolution of Raman spectra, discussions of surface reaction and bulk diffusion, and operando Raman spectroscopy of V2O5 during propane oxidation (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Yuanqing Wang - Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4−6, D-14195 Berlin, GermanyBasCat, UniCat BASF Jointlab, Technische Universität Berlin, D-10623 Berlin, GermanyPresent Address: Materials Genome Institute, Shanghai University, 200444, Shanghai, ChinaOrcidhttps://orcid.org/0000-0002-1332-7956
    • Frank Rosowski - BasCat, UniCat BASF Jointlab, Technische Universität Berlin, D-10623 Berlin, GermanyHeterogeneous Catalysis, BASF SE, Process Research and Chemical Engineering, Carl-Bosch-Straße 38, D-67065 Ludwigshafen, Germany
    • Robert Schlögl - Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg 4−6, D-14195 Berlin, GermanyMax-Planck-Institut für Chemische Energiekonversion, Stiftstrase 34−36, D-45470 Mülheim, Germany
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

  • Funding

    Open access funded by Max Planck Society.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank Dr. Andrey Tarasov for TG and DTA measurements, Dr. Girgsdies and Jasmin Allan for XRD measurements, Maike Hashagen for BET measurements, Dr. Olaf Timpe for XRF measurements, Dr. Xing Huang for STEM measurements, and Jutta Kröhnert for UV/vis spectrum measurements. We also express thanks for a helpful discussion with Dr. Anna Wernbacher. The DFT calculations were performed using the computational resource of the Computer Support Group (PP&B) at the Fritz-Haber-Institut. This work was conducted in the framework of the BasCat collaboration among BASF SE, the Technical University Berlin, the Fritz-Haber-Institut der Max-Planck-Gesellschaft, and the cluster of excellence “Unified Concepts in Catalysis” (UniCat http://www.unicat.tu-berlin.de).

References

Click to copy section linkSection link copied!

This article references 81 other publications.

  1. 1
    Panov, G. I.; Dubkov, K. A.; Starokon, E. V. Active Oxygen in Selective Oxidation Catalysis. Catal. Today 2006, 117, 148155,  DOI: 10.1016/j.cattod.2006.05.019
  2. 2
    Freund, H.-J. Oxygen Activation on Oxide Surfaces: A Perspective at the Atomic Level. Catal. Today 2014, 238, 29,  DOI: 10.1016/j.cattod.2014.05.037
  3. 3
    Hwang, J.; Rao, R. R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in Catalysis and Electrocatalysis. Science 2017, 358, 751756,  DOI: 10.1126/science.aam7092
  4. 4
    Adler, S. B. Factors Governing Oxygen Reduction in Solid Oxide Fuel Cell Cathodes. Chem. Rev. 2004, 104, 47914844,  DOI: 10.1021/cr020724o
  5. 5
    Gurlo, A. Interplay between O2 and SnO2: Oxygen Ionosorption and Spectroscopic Evidence for Adsorbed Oxygen. ChemPhysChem 2006, 7, 20412052,  DOI: 10.1002/cphc.200600292
  6. 6
    Schlögl, R. Heterogeneous Catalysis. Angew. Chem., Int. Ed. 2015, 54, 34653520,  DOI: 10.1002/anie.201410738
  7. 7
    Haber, J.; Turek, W. Kinetic Studies as a Method to Differentiate between Oxygen Species Involved in the Oxidation of Propene. J. Catal. 2000, 190, 320326,  DOI: 10.1006/jcat.1999.2764
  8. 8
    Rasmussen, M. D.; Molina, L. M.; Hammer, B. Adsorption, Diffusion, and Dissociation of Molecular Oxygen at Defected TiO2(110): A Density Functional Theory Study. J. Chem. Phys. 2004, 120, 988997,  DOI: 10.1063/1.1631922
  9. 9
    Ganduglia-Pirovano, M. V.; Hofmann, A.; Sauer, J. Oxygen Vacancies in Transition Metal and Rare Earth Oxides: Current State of Understanding and Remaining Challenges. Surf. Sci. Rep. 2007, 62, 219270,  DOI: 10.1016/j.surfrep.2007.03.002
  10. 10
    Cui, Y.; Shao, X.; Baldofski, M.; Sauer, J.; Nilius, N.; Freund, H.-J. Adsorption, Activation, and Dissociation of Oxygen on Doped Oxides. Angew. Chem., Int. Ed. 2013, 52, 1138511387,  DOI: 10.1002/anie.201305119
  11. 11
    McFarland, E. W.; Metiu, H. Catalysis by Doped Oxides. Chem. Rev. 2013, 113, 43914427,  DOI: 10.1021/cr300418s
  12. 12
    Schwach, P.; Hamilton, N.; Eichelbaum, M.; Thum, L.; Lunkenbein, T.; Schlögl, R.; Trunschke, A. Structure Sensitivity of the Oxidative Activation of Methane over MgO Model Catalysts: II. Nature of Active Sites and Reaction Mechanism. J. Catal. 2015, 329, 574587,  DOI: 10.1016/j.jcat.2015.05.008
  13. 13
    Zasada, F.; Piskorz, W.; Janas, J.; Gryboś, J.; Indyka, P.; Sojka, Z. Reactive Oxygen Species on the (100) Facet of Cobalt Spinel Nanocatalyst and Their Relevance in 16O2/18O2 Isotopic Exchange, deN2O, and deCH4 Processes─a Theoretical and Experimental Account. ACS Catal. 2015, 5, 68796892,  DOI: 10.1021/acscatal.5b01900
  14. 14
    Hävecker, M.; Wrabetz, S.; Kröhnert, J.; Csepei, L.-I.; Naumann d’Alnoncourt, R.; Kolen’ko, Y. V.; Girgsdies, F.; Schlögl, R.; Trunschke, A. Surface Chemistry of Phase-Pure M1Movtenb Oxide During Operation in Selective Oxidation of Propane to Acrylic Acid. J. Catal. 2012, 285, 4860,  DOI: 10.1016/j.jcat.2011.09.012
  15. 15
    Amakawa, K. How Strain Affects the Reactivity of Surface Metal Oxide Catalysts. Angew. Chem., Int. Ed. 2013, 52, 1355313557,  DOI: 10.1002/anie.201306620
  16. 16
    Davydov, A. A. Molecular Spectroscopy of Oxide Catalyst Surfaces; John Wiley & Sons Ltd.: Chichester, U.K., 2003.
  17. 17
    Wernbacher, A. M.; Kube, P.; Hävecker, M.; Schlögl, R.; Trunschke, A. Electronic and Dielectric Properties of MoV-Oxide (M1 Phase) under Alkane Oxidation Conditions. J. Phys. Chem. C 2019, 123, 1326913282,  DOI: 10.1021/acs.jpcc.9b01273
  18. 18
    Wernbacher, A. M.; Eichelbaum, M.; Risse, T.; Cap, S.; Trunschke, A.; Schlögl, R. Operando Electrical Conductivity and Complex Permittivity Study on Vanadia Oxidation Catalysts. J. Phys. Chem. C 2019, 123, 80058017,  DOI: 10.1021/acs.jpcc.8b07417
  19. 19
    Koch, G. Surface Conditions That Constrain Alkane Oxidation on Perovskites. ACS Catal. 2020, 10, 70077020,  DOI: 10.1021/acscatal.0c01289
  20. 20
    Che, M.; Tench, A. J. Characterization and Reactivity of Mononuclear Oxygen Species on Oxide Surfaces. Adv. Catal. 1982, 31, 77133,  DOI: 10.1016/S0360-0564(08)60453-8
  21. 21
    Che, M.; Tench, A. J. Characterization and Reactivity of Molecular Oxygen Species on Oxide Surfaces. Adv. Catal. 1983, 32, 1148,  DOI: 10.1016/S0360-0564(08)60439-3
  22. 22
    Nováková, J. Isotopic Exchange of Oxygen 18O between the Gaseous Phase and Oxide Catalysts. Catal. Rev. 1971, 4, 77113,  DOI: 10.1080/01614947108075486
  23. 23
    Choi, S. O.; Penninger, M.; Kim, C. H.; Schneider, W. F.; Thompson, L. T. Experimental and Computational Investigation of Effect of Sr on NO Oxidation and Oxygen Exchange for La1–xSrxCoO3 Perovskite Catalysts. ACS Catal. 2013, 3, 27192728,  DOI: 10.1021/cs400522r
  24. 24
    Winter, E. R. S. The Reactivity of Oxide Surfaces. Adv. Catal. 1958, 10, 196241,  DOI: 10.1016/S0360-0564(08)60408-3
  25. 25
    Winter, E. R. S. Exchange Reactions of Oxides. Part IX. J. Chem. Soc. A: Inorg., Phys., Theor. 1968, 28892902,  DOI: 10.1039/j19680002889
  26. 26
    Klier, K.; Nováková, J.; Jíru, P. Exchange Reactions of Oxygen between Oxygen Molecules and Solid Oxides. J. Catal. 1963, 2, 479484,  DOI: 10.1016/0021-9517(63)90003-4
  27. 27
    Boreskov, G. K. The Catalysis of Isotopic Exchange in Molecular Oxygen. Adv. Catal. 1965, 15, 285339,  DOI: 10.1016/S0360-0564(08)60556-8
  28. 28
    Doornkamp, C.; Clement, M.; Gao, X.; Deo, G.; Wachs, I. E.; Ponec, V. The Oxygen Isotopic Exchange Reaction on Vanadium Oxide Catalysts. J. Catal. 1999, 185, 415422,  DOI: 10.1006/jcat.1999.2490
  29. 29
    Doornkamp, C.; Clement, M.; Ponec, V. The Isotopic Exchange Reaction of Oxygen on Metal Oxides. J. Catal. 1999, 182, 390399,  DOI: 10.1006/jcat.1998.2377
  30. 30
    Wachs, I. E.; Jehng, J.-M.; Deo, G.; Weckhuysen, B. M.; Guliants, V. V.; Benziger, J. B.; Sundaresan, S. Fundamental Studies of Butane Oxidation over Model-Supported Vanadium Oxide Catalysts: Molecular Structure-Reactivity Relationships. J. Catal. 1997, 170, 7588,  DOI: 10.1006/jcat.1997.1742
  31. 31
    Klug, C. A.; Kroeker, S.; Aguiar, P. M.; Zhou, M.; Stec, D. F.; Wachs, I. E. Insights into Oxygen Exchange between Gaseous O2 and Supported Vanadium Oxide Catalysts Via 17O NMR. Chem. Mater. 2009, 21, 41274134,  DOI: 10.1021/cm802680x
  32. 32
    Avdeev, V. I.; Bedilo, A. F. Molecular Mechanism of Oxygen Isotopic Exchange over Supported Vanadium Oxide Catalyst VOx/TiO2. J. Phys. Chem. C 2013, 117, 28792887,  DOI: 10.1021/jp311322b
  33. 33
    Avdeev, V. I.; Bedilo, A. F. Electronic Structure of Oxygen Radicals on the Surface of VOx/TiO2 Catalysts and Their Role in Oxygen Isotopic Exchange. J. Phys. Chem. C 2013, 117, 1470114709,  DOI: 10.1021/jp404921d
  34. 34
    Koranne, M. M.; Goodwin, J. G.; Marcelin, G. Oxygen Involvement in the Partial Oxidation of Methane on Supported and Unsupported V2O5. J. Catal. 1994, 148, 378387,  DOI: 10.1006/jcat.1994.1218
  35. 35
    Kera, Y.; Teratani, S.; Hirota, K. Infrared Spectra of Surface V═O Bond of Vanadium Pentoxide. Bull. Chem. Soc. Jpn. 1967, 40, 24582458,  DOI: 10.1246/bcsj.40.2458
  36. 36
    Hirota, K.; Kera, Y.; Teratani, S. Carbon Monoxide Oxidation with an Oxygen Tracer over a Vanadium Pentoxide Catalyst. J. Phys. Chem. 1968, 72, 31333141,  DOI: 10.1021/j100855a010
  37. 37
    Kera, Y.; Hirota, K. Infrared Spectroscopic Study of Oxygen Species in Vanadium Pentoxide with Reference to Its Activity in Catalytic Oxidation. J. Phys. Chem. 1969, 73, 39733981,  DOI: 10.1021/j100845a070
  38. 38
    Fletcher, W. H.; Rayside, J. S. High Resolution Vibrational Raman Spectrum of Oxygen. J. Raman Spectrosc. 1974, 2, 314,  DOI: 10.1002/jrs.1250020102
  39. 39
    Pushkarev, V. V.; Kovalchuk, V. I.; d’Itri, J. L. Probing Defect Sites on the CeO2 Surface with Dioxygen. J. Phys. Chem. B 2004, 108, 53415348,  DOI: 10.1021/jp0311254
  40. 40
    Choi, Y. M.; Abernathy, H.; Chen, H.-T.; Lin, M. C.; Liu, M. Characterization of O2–CeO2 Interactions Using in Situ Raman Spectroscopy and First-Principle Calculations. ChemPhysChem 2006, 7, 19571963,  DOI: 10.1002/cphc.200600190
  41. 41
    Wu, Z.; Dai, S.; Overbury, S. H. Multiwavelength Raman Spectroscopic Study of Silica-Supported Vanadium Oxide Catalysts. J. Phys. Chem. C 2010, 114, 412422,  DOI: 10.1021/jp9084876
  42. 42
    Weckhuysen, B. M.; Jehng, J.-M.; Wachs, I. E. In Situ Raman Spectroscopy of Supported Transition Metal Oxide Catalysts: 18O216O2 Isotopic Labeling Studies. J. Phys. Chem. B 2000, 104, 73827387,  DOI: 10.1021/jp000055n
  43. 43
    Lee, E. L.; Wachs, I. E. In Situ Raman Spectroscopy of SiO2-Supported Transition Metal Oxide Catalysts: An Isotopic 18O-16O Exchange Study. J. Phys. Chem. C 2008, 112, 64876498,  DOI: 10.1021/jp076485w
  44. 44
    Moisii, C.; van de Burgt, L. J.; Stiegman, A. E. Resonance Raman Spectroscopy of Discrete Silica-Supported Vanadium Oxide. Chem. Mater. 2008, 20, 39273935,  DOI: 10.1021/cm800095g
  45. 45
    Magg, N. Vibrational Spectra of Alumina- and Silica-Supported Vanadia Revisited: An Experimental and Theoretical Model Catalyst Study. J. Catal. 2004, 226, 88100,  DOI: 10.1016/j.jcat.2004.04.021
  46. 46
    Döbler, J.; Pritzsche, M.; Sauer, J. Vibrations of Silica Supported Vanadia: Variation with Particle Size and Local Surface Structure. J. Phys. Chem. C 2009, 113, 1245412464,  DOI: 10.1021/jp901774t
  47. 47
    Nitsche, D.; Hess, C. Normal Mode Analysis of Silica-Supported Vanadium Oxide Catalysts: Comparison of Theory with Experiment. Catal. Commun. 2014, 52, 4044,  DOI: 10.1016/j.catcom.2014.04.008
  48. 48
    Oyama, S. T.; Went, G. T.; Lewis, K. B.; Bell, A. T.; Somorjai, G. A. Oxygen Chemisorption and Laser Raman Spectroscopy of Unsupported and Silica-Supported Vanadium Oxide Catalysts. J. Phys. Chem. 1989, 93, 67866790,  DOI: 10.1021/j100355a041
  49. 49
    Ono, T.; Tanaka, Y.; Takeuchi, T.; Yamamoto, K. Characterization of K-Mixed V2O5 Catalyst and Oxidative Dehydrogenation of Propane on It. J. Mol. Catal. A Chem. 2000, 159, 293300,  DOI: 10.1016/S1381-1169(00)00218-1
  50. 50
    Ono, T.; Numata, H. Characteristic Features of Raman Band Shifts of Vanadium Oxide Catalysts Exchanged with the 18O Tracer and Active Sites for Reoxidation. J. Mol. Catal. A Chem. 1997, 116, 421429,  DOI: 10.1016/S1381-1169(96)00425-6
  51. 51
    Browne, M. P.; Sofer, Z.; Pumera, M. Layered and Two Dimensional Metal Oxides for Electrochemical Energy Conversion. Energy Environ. Sci. 2019, 12, 4158,  DOI: 10.1039/C8EE02495B
  52. 52
    Wachs, I. E. Recent Conceptual Advances in the Catalysis Science of Mixed Metal Oxide Catalytic Materials. Catal. Today 2005, 100, 7994,  DOI: 10.1016/j.cattod.2004.12.019
  53. 53
    Grant, J. T.; Venegas, J. M.; McDermott, W. P.; Hermans, I. Aerobic Oxidations of Light Alkanes over Solid Metal Oxide Catalysts. Chem. Rev. 2018, 118, 27692815,  DOI: 10.1021/acs.chemrev.7b00236
  54. 54
    Kuba, S.; Knözinger, H. Time-Resolved in Situ Raman Spectroscopy of Working Catalysts: Sulfated and Tungstated Zirconia. J. Raman Spectrosc. 2002, 33, 325332,  DOI: 10.1002/jrs.815
  55. 55
    Giannozzi, P. Quantum Espresso: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter 2009, 21, 395502,  DOI: 10.1088/0953-8984/21/39/395502
  56. 56
    Shklover, V.; Haibach, T.; Ried, F.; Nesper, R.; Novák, P. Crystal Structure of the Product of Mg2+ Insertion into V2O5 single Crystals. J. Solid State Chem. 1996, 123, 317323,  DOI: 10.1006/jssc.1996.0186
  57. 57
    Grimme, S. Semiempirical Gga-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 17871799,  DOI: 10.1002/jcc.20495
  58. 58
    Barone, V.; Casarin, M.; Forrer, D.; Pavone, M.; Sambi, M.; Vittadini, A. Role and Effective Treatment of Dispersive Forces in Materials: Polyethylene and Graphite Crystals as Test Cases. J. Comput. Chem. 2009, 30, 934939,  DOI: 10.1002/jcc.21112
  59. 59
    Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 51885192,  DOI: 10.1103/PhysRevB.13.5188
  60. 60
    Kim, H.; Kosuda, K. M.; Van Duyne, R. P.; Stair, P. C. Resonance Raman and Surface- and Tip-Enhanced Raman Spectroscopy Methods to Study Solid Catalysts and Heterogeneous Catalytic Reactions. Chem. Soc. Rev. 2010, 39, 48204844,  DOI: 10.1039/c0cs00044b
  61. 61
    Harima, H. Raman Scattering Characterization on SiC. Microelectron. Eng. 2006, 83, 126129,  DOI: 10.1016/j.mee.2005.10.037
  62. 62
    Gilson, T. R.; Bizri, O. F.; Cheetham, N. Single-Crystal Raman and Infrared Spectra of Vanadium(V) Oxide. J. Chem. Soc., Dalton Trans. 1973, 291294,  DOI: 10.1039/dt9730000291
  63. 63
    Abello, L.; Husson, E.; Repelin, Y.; Lucazeau, G. Vibrational Spectra and Valence Force Field of Crystalline V2O5. Spectrochim. Acta, Part A 1983, 39, 641651,  DOI: 10.1016/0584-8539(83)80040-3
  64. 64
    Clauws, P.; Broeckx, J.; Vennik, J. Lattice Vibrations of V2O5. Calculation of Normal Vibrations in a Urey-Bradley Force Field. Phys. Status Solidi B 1985, 131, 459473,  DOI: 10.1002/pssb.2221310207
  65. 65
    Brázdová, V.; Ganduglia-Pirovano, M. V.; Sauer, J. Periodic Density Functional Study on Structural and Vibrational Properties of Vanadium Oxide Aggregates. Phys. Rev. B 2004, 69, 165420,  DOI: 10.1103/PhysRevB.69.165420
  66. 66
    Zhou, B.; He, D. Raman Spectrum of Vanadium Pentoxide from Density-Functional Perturbation Theory. J. Raman Spectrosc. 2008, 39, 14751481,  DOI: 10.1002/jrs.2025
  67. 67
    Wu, Z.; Kim, H.-S.; Stair, P. C.; Rugmini, S.; Jackson, S. D. On the Structure of Vanadium Oxide Supported on Aluminas: UV and Visible Raman Spectroscopy, UV–Visible Diffuse Reflectance Spectroscopy, and Temperature-Programmed Reduction Studies. J. Phys. Chem. B 2005, 109, 27932800,  DOI: 10.1021/jp046011m
  68. 68
    Hermann, K.; Witko, M.; Druzinic, R.; Tokarz, R. Hydrogen Assisted Oxygen Desorption from the V2O5(010) Surface. Top. Catal. 2000, 11/12, 6775,  DOI: 10.1023/A:1027206705195
  69. 69
    Huang, Y.-L.; Pellegrinelli, C.; Wachsman, E. D. Reaction Kinetics of Gas–Solid Exchange Using Gas Phase Isotopic Oxygen Exchange. ACS Catal. 2016, 6, 60256032,  DOI: 10.1021/acscatal.6b01462
  70. 70
    Huang, Y.-L.; Pellegrinelli, C.; Sakbodin, M.; Wachsman, E. D. Molecular Reactions of O2 and CO2 on Ionically Conducting Catalyst. ACS Catal. 2018, 8, 12311237,  DOI: 10.1021/acscatal.7b03467
  71. 71
    Huang, Y.-L.; Pellegrinelli, C.; Wachsman, E. D. Oxygen Dissociation Kinetics of Concurrent Heterogeneous Reactions on Metal Oxides. ACS Catal. 2017, 7, 57665772,  DOI: 10.1021/acscatal.7b01096
  72. 72
    Tilley, R. J. Understanding Solids: The Science of Materials; Wiley: 2013.
  73. 73
    Cameron, W. C.; Parkas, A.; Litz, L. M. Exchange of Isotopic Oxygen between Vanadium Pentoxide, Gaseous Oxygen and Water. J. Phys. Chem. 1953, 57, 229238,  DOI: 10.1021/j150503a022
  74. 74
    Milan, E. F. The Dissociation Pressure of Vanadium Pentoxide. J. Phys. Chem. 1929, 33, 498508,  DOI: 10.1021/j150298a002
  75. 75
    Nasu, N. The Dissociation Pressure of Vanadium Pentoxide. J. Chem. Soc. Jpn. 1935, 56, 666669,  DOI: 10.1246/nikkashi1921.56.6_666
  76. 76
    Spitsyn, B. N.; Maidanovskaya, L. L. The Thermal Dissociation of Vanadium Pentoxide. Zh. Fiz. Khim. 1959, 33, 180183
  77. 77
    Chen, K.; Khodakov, A.; Yang, J.; Bell, A. T.; Iglesia, E. Isotopic Tracer and Kinetic Studies of Oxidative Dehydrogenation Pathways on Vanadium Oxide Catalysts. J. Catal. 1999, 186, 325333,  DOI: 10.1006/jcat.1999.2510
  78. 78
    Ma, W. Y.; Zhou, B.; Wang, J. F.; Zhang, X. D.; Jiang, Z. Y. Effect of Oxygen Vacancy on Li-Ion Diffusion in a V2O5 Cathode: A First-Principles Study. J. Phys. D Appl. Phys. 2013, 46, 105306,  DOI: 10.1088/0022-3727/46/10/105306
  79. 79
    Mars, P.; van Krevelen, D. W. Oxidations Carried out by Means of Vanadium Oxide Catalysts. Chem. Eng. Sci. 1954, 3, 4159,  DOI: 10.1016/S0009-2509(54)80005-4
  80. 80
    Routray, K.; Reddy, K. R. S. K.; Deo, G. Oxidative Dehydrogenation of Propane on V2O5/Al2O3 and V2O5/TiO2 Catalysts: Understanding the Effect of Support by Parameter Estimation. Appl. Catal. A Gen. 2004, 265, 103113,  DOI: 10.1016/j.apcata.2004.01.006
  81. 81
    Alexopoulos, K.; Reyniers, M.-F.; Marin, G. B. Reaction Path Analysis of Propane Selective Oxidation over V2O5 and V2O5/TiO2. J. Catal. 2012, 289, 127139,  DOI: 10.1016/j.jcat.2012.01.019

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 11 publications.

  1. Sudha Arumugam, Ranjithvel Mohanasundaram, Parasuraman Swaminathan. Salt Electrolytes for Vanadium Pentaoxide-Based Electrochromic Devices. ACS Applied Engineering Materials 2025, 3 (4) , 901-913. https://doi.org/10.1021/acsaenm.5c00040
  2. Daniyal Kiani, Israel E. Wachs. The Conundrum of “Pair Sites” in Langmuir–Hinshelwood Reaction Kinetics in Heterogeneous Catalysis. ACS Catalysis 2024, 14 (13) , 10260-10270. https://doi.org/10.1021/acscatal.4c02813
  3. Gregor Koch, Giulia Bellini, Frank Girgsdies, Michael Hävecker, Spencer J. Carey, Olaf Timpe, Thomas Götsch, Thomas Lunkenbein, Gudrun Auffermann, Andrey Tarasov, Robert Schlögl, Annette Trunschke. Structural Analysis and Redox Properties of Oxygen-“Breathing” A(Mn1–xCux)O3 (A = La, Pr) Perovskites. Chemistry of Materials 2024, 36 (11) , 5388-5404. https://doi.org/10.1021/acs.chemmater.4c00136
  4. Bapuso M. Babar, Umesh D. Babar, Abhijeet J. Kale, Tushar T. Bhosale, Udayraj T. Pawar, Prakash M. Kadam, Sandip V. Nipane, Nishad G. Deshpande, Pramod S. Patil, Laxman D. Kadam. Polyvinylpyrrolidone influenced reduced graphene oxide modified vanadium pentoxide nano rods for NO2 detection and supercapacitor application. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2025, 710 , 136239. https://doi.org/10.1016/j.colsurfa.2025.136239
  5. Valentina Paolucci, Maria Basso, Vittorio Ricci, Elena Colusso, Mattia Cattelan, Enrico Napolitani, Alessandro Martucci, Carlo Cantalini. Pulsed-laser annealed amorphous/crystalline a-V2O5/VO2 heterostructure for selective NO2 and H2 sensing. Sensors and Actuators B: Chemical 2025, 428 , 137235. https://doi.org/10.1016/j.snb.2025.137235
  6. Peng Zhang, Jingyuan Fan, Yuanqing Wang, Yuying Dang, Saskia Heumann, Yuxiao Ding. Insights into the role of defects on the Raman spectroscopy of carbon nanotube and biomass-derived carbon. Carbon 2024, 222 , 118998. https://doi.org/10.1016/j.carbon.2024.118998
  7. Jose Alirio Mendoza Mesa, Sven Robijns, Iqtidar Ali Khan, Marco G. Rigamonti, Max L. Bols, Michiel Dusselier. Support effects in vanadium incipient wetness impregnation for oxidative and non-oxidative propane dehydrogenation catalysis. Catalysis Today 2024, 430 , 114546. https://doi.org/10.1016/j.cattod.2024.114546
  8. V.M. Dzhagan, M. Ya Valakh, O.F. Isaieva, V.O. Yukhymchuk, O.A. Stadnik, O. Yo Gudymenko, P.M. Lytvyn, O.A. Kulbachynskyi, V.S. Yefanov, B.M. Romanyuk, V.P. Melnik. Raman fingerprints of different vanadium oxides as impurity phases in VO2 films. Optical Materials 2024, 148 , 114894. https://doi.org/10.1016/j.optmat.2024.114894
  9. O. Chukova, S. G. Nedilko, T. Voitenko, R. Minikayev, W. Paszkowicz, V. Stasiv, Y. Zhydachevskyy, A. Suchocki. Luminescence mechanisms in the 2V2O5–xLi2O–(98 − x)B2O3 glass matrices developed for creation of glass–ceramic materials. Journal of Materials Science: Materials in Electronics 2023, 34 (7) https://doi.org/10.1007/s10854-023-10026-4
  10. Sanja J. Armaković, Aleksandra Jovanoski Kostić, Andrijana Bilić, Maria M. Savanović, Nataša Tomić, Aleksandar Kremenović, Maja Šćepanović, Mirjana Grujić-Brojčin, Jovana Ćirković, Stevan Armaković. Photocatalytic Activity of the V2O5 Catalyst toward Selected Pharmaceuticals and Their Mixture: Influence of the Molecular Structure on the Efficiency of the Process. Molecules 2023, 28 (2) , 655. https://doi.org/10.3390/molecules28020655
  11. Yuan Gao, Mingxin Jiang, Liuqingqing Yang, Zhuo Li, Fei-Xiang Tian, Yulian He. Recent progress of catalytic methane combustion over transition metal oxide catalysts. Frontiers in Chemistry 2022, 10 https://doi.org/10.3389/fchem.2022.959422

The Journal of Physical Chemistry C

Cite this: J. Phys. Chem. C 2022, 126, 7, 3443–3456
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcc.2c00174
Published February 9, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

3130

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Unit cell of V2O5: O1, terminal oxygen; O2, bridging oxygen; O3, chain oxygen.

    Figure 2

    Figure 2. UV/vis spectrum (a) and Raman spectra (b) of V2O5 collected at room temperature (ex situ). The laser wavelength used for the Raman measurements is shown next to the Raman spectra. The Raman spectra were corrected with respect to instrumental effects, taking into account the known response curve of a white lamp and the absorption spectrum of V2O5 (see Figure S8 for details of the correction procedure and Figure S10 for uncorrected spectra); All spectra were normalized to the corresponding maximum band intensity ([0,1]), which differs for the different excitation energies; The band positions indicated were determined for the spectrum measured with the excitation wavelength 532 nm; The intensity ratio of the band at 994 cm–1 to the band at 144 cm–1 as a function of excitation wavelength is plotted in part a.

    Figure 3

    Figure 3. Raman spectra of V2O5 measured at room temperature before and after isotopic oxygen exchange; Gray and red spectra denote the cases before and after exchange at 322, 431, and 573 °C, respectively, using 18O2 (20% in He) or a mixture of propane and oxygen (C3H8 (1%) + 18O2 (19%) in He) for 2 h. All the spectra were normalized to [0,1]. The positions of all spectra were aligned with respect to the band at 994 cm–1. Laser: 532 nm. Heating rate: 5 °C/min. Total flow rate: 10 mL/min.

    Figure 4

    Figure 4. Deconvolution of Raman spectra collected at room temperature after treatment at different temperatures and in different gas atmospheres (the same Raman spectra as shown in Figure 3) in the V–O3 (a) as well as V═O1 and O1–V–O2 stretching vibration region (b). Temperatures and gas atmospheres are specified in the right top corner of each section. (c) Corresponding representations of vibrational motions of phonon modes.

    Figure 5

    Figure 5. In situ Raman spectra of V2O5 at 573 °C under 20% 18O2 in He (a), 1% C3H8 + 19% 18O2 in He (b), and 1% C3D8 + 19% 18O2 in He (c) as a function of time. Conditions of exchange experiments are described in the caption of Figure 3. The numbers show the time that had passed since the start of the experiment at the moment of recording. Laser: 532 nm.

    Figure 6

    Figure 6. Extent of oxygen exchange as a function of time; The lines are the fitted results using eqs 3 (dashed) and 4 (solid). The experimental data points were taken from the experiments shown in Figure 5. Note that the time used in the graph is the average time taken to record each spectrum.

    Figure 7

    Figure 7. Effect of the nature of the alkane on isotope oxygen exchange. (a) Raman spectra of V2O5 measured at room temperature before and after isotopic oxygen exchange experiments under 10% C3H8 + 5% 18O2 and 10% C2H6 + 5% 18O2, respectively, at 400 °C. (b) In situ Raman spectra of V2O5 at 400 °C under 16O2 and 10% C4H10 + 5% 18O2 as a function of time. Numbers on the spectra indicate the starting time of measurement. Asterisks represent cut cosmic ray signals. Raman spectra were normalized to [0,1]. Laser: 532 nm. Exchange experiments were conducted at 400 °C for 2 h in the Harrick Raman chamber. Heating rate: 5 °C/min. Total flow rate: 10 mL/min.

  • References


    This article references 81 other publications.

    1. 1
      Panov, G. I.; Dubkov, K. A.; Starokon, E. V. Active Oxygen in Selective Oxidation Catalysis. Catal. Today 2006, 117, 148155,  DOI: 10.1016/j.cattod.2006.05.019
    2. 2
      Freund, H.-J. Oxygen Activation on Oxide Surfaces: A Perspective at the Atomic Level. Catal. Today 2014, 238, 29,  DOI: 10.1016/j.cattod.2014.05.037
    3. 3
      Hwang, J.; Rao, R. R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in Catalysis and Electrocatalysis. Science 2017, 358, 751756,  DOI: 10.1126/science.aam7092
    4. 4
      Adler, S. B. Factors Governing Oxygen Reduction in Solid Oxide Fuel Cell Cathodes. Chem. Rev. 2004, 104, 47914844,  DOI: 10.1021/cr020724o
    5. 5
      Gurlo, A. Interplay between O2 and SnO2: Oxygen Ionosorption and Spectroscopic Evidence for Adsorbed Oxygen. ChemPhysChem 2006, 7, 20412052,  DOI: 10.1002/cphc.200600292
    6. 6
      Schlögl, R. Heterogeneous Catalysis. Angew. Chem., Int. Ed. 2015, 54, 34653520,  DOI: 10.1002/anie.201410738
    7. 7
      Haber, J.; Turek, W. Kinetic Studies as a Method to Differentiate between Oxygen Species Involved in the Oxidation of Propene. J. Catal. 2000, 190, 320326,  DOI: 10.1006/jcat.1999.2764
    8. 8
      Rasmussen, M. D.; Molina, L. M.; Hammer, B. Adsorption, Diffusion, and Dissociation of Molecular Oxygen at Defected TiO2(110): A Density Functional Theory Study. J. Chem. Phys. 2004, 120, 988997,  DOI: 10.1063/1.1631922
    9. 9
      Ganduglia-Pirovano, M. V.; Hofmann, A.; Sauer, J. Oxygen Vacancies in Transition Metal and Rare Earth Oxides: Current State of Understanding and Remaining Challenges. Surf. Sci. Rep. 2007, 62, 219270,  DOI: 10.1016/j.surfrep.2007.03.002
    10. 10
      Cui, Y.; Shao, X.; Baldofski, M.; Sauer, J.; Nilius, N.; Freund, H.-J. Adsorption, Activation, and Dissociation of Oxygen on Doped Oxides. Angew. Chem., Int. Ed. 2013, 52, 1138511387,  DOI: 10.1002/anie.201305119
    11. 11
      McFarland, E. W.; Metiu, H. Catalysis by Doped Oxides. Chem. Rev. 2013, 113, 43914427,  DOI: 10.1021/cr300418s
    12. 12
      Schwach, P.; Hamilton, N.; Eichelbaum, M.; Thum, L.; Lunkenbein, T.; Schlögl, R.; Trunschke, A. Structure Sensitivity of the Oxidative Activation of Methane over MgO Model Catalysts: II. Nature of Active Sites and Reaction Mechanism. J. Catal. 2015, 329, 574587,  DOI: 10.1016/j.jcat.2015.05.008
    13. 13
      Zasada, F.; Piskorz, W.; Janas, J.; Gryboś, J.; Indyka, P.; Sojka, Z. Reactive Oxygen Species on the (100) Facet of Cobalt Spinel Nanocatalyst and Their Relevance in 16O2/18O2 Isotopic Exchange, deN2O, and deCH4 Processes─a Theoretical and Experimental Account. ACS Catal. 2015, 5, 68796892,  DOI: 10.1021/acscatal.5b01900
    14. 14
      Hävecker, M.; Wrabetz, S.; Kröhnert, J.; Csepei, L.-I.; Naumann d’Alnoncourt, R.; Kolen’ko, Y. V.; Girgsdies, F.; Schlögl, R.; Trunschke, A. Surface Chemistry of Phase-Pure M1Movtenb Oxide During Operation in Selective Oxidation of Propane to Acrylic Acid. J. Catal. 2012, 285, 4860,  DOI: 10.1016/j.jcat.2011.09.012
    15. 15
      Amakawa, K. How Strain Affects the Reactivity of Surface Metal Oxide Catalysts. Angew. Chem., Int. Ed. 2013, 52, 1355313557,  DOI: 10.1002/anie.201306620
    16. 16
      Davydov, A. A. Molecular Spectroscopy of Oxide Catalyst Surfaces; John Wiley & Sons Ltd.: Chichester, U.K., 2003.
    17. 17
      Wernbacher, A. M.; Kube, P.; Hävecker, M.; Schlögl, R.; Trunschke, A. Electronic and Dielectric Properties of MoV-Oxide (M1 Phase) under Alkane Oxidation Conditions. J. Phys. Chem. C 2019, 123, 1326913282,  DOI: 10.1021/acs.jpcc.9b01273
    18. 18
      Wernbacher, A. M.; Eichelbaum, M.; Risse, T.; Cap, S.; Trunschke, A.; Schlögl, R. Operando Electrical Conductivity and Complex Permittivity Study on Vanadia Oxidation Catalysts. J. Phys. Chem. C 2019, 123, 80058017,  DOI: 10.1021/acs.jpcc.8b07417
    19. 19
      Koch, G. Surface Conditions That Constrain Alkane Oxidation on Perovskites. ACS Catal. 2020, 10, 70077020,  DOI: 10.1021/acscatal.0c01289
    20. 20
      Che, M.; Tench, A. J. Characterization and Reactivity of Mononuclear Oxygen Species on Oxide Surfaces. Adv. Catal. 1982, 31, 77133,  DOI: 10.1016/S0360-0564(08)60453-8
    21. 21
      Che, M.; Tench, A. J. Characterization and Reactivity of Molecular Oxygen Species on Oxide Surfaces. Adv. Catal. 1983, 32, 1148,  DOI: 10.1016/S0360-0564(08)60439-3
    22. 22
      Nováková, J. Isotopic Exchange of Oxygen 18O between the Gaseous Phase and Oxide Catalysts. Catal. Rev. 1971, 4, 77113,  DOI: 10.1080/01614947108075486
    23. 23
      Choi, S. O.; Penninger, M.; Kim, C. H.; Schneider, W. F.; Thompson, L. T. Experimental and Computational Investigation of Effect of Sr on NO Oxidation and Oxygen Exchange for La1–xSrxCoO3 Perovskite Catalysts. ACS Catal. 2013, 3, 27192728,  DOI: 10.1021/cs400522r
    24. 24
      Winter, E. R. S. The Reactivity of Oxide Surfaces. Adv. Catal. 1958, 10, 196241,  DOI: 10.1016/S0360-0564(08)60408-3
    25. 25
      Winter, E. R. S. Exchange Reactions of Oxides. Part IX. J. Chem. Soc. A: Inorg., Phys., Theor. 1968, 28892902,  DOI: 10.1039/j19680002889
    26. 26
      Klier, K.; Nováková, J.; Jíru, P. Exchange Reactions of Oxygen between Oxygen Molecules and Solid Oxides. J. Catal. 1963, 2, 479484,  DOI: 10.1016/0021-9517(63)90003-4
    27. 27
      Boreskov, G. K. The Catalysis of Isotopic Exchange in Molecular Oxygen. Adv. Catal. 1965, 15, 285339,  DOI: 10.1016/S0360-0564(08)60556-8
    28. 28
      Doornkamp, C.; Clement, M.; Gao, X.; Deo, G.; Wachs, I. E.; Ponec, V. The Oxygen Isotopic Exchange Reaction on Vanadium Oxide Catalysts. J. Catal. 1999, 185, 415422,  DOI: 10.1006/jcat.1999.2490
    29. 29
      Doornkamp, C.; Clement, M.; Ponec, V. The Isotopic Exchange Reaction of Oxygen on Metal Oxides. J. Catal. 1999, 182, 390399,  DOI: 10.1006/jcat.1998.2377
    30. 30
      Wachs, I. E.; Jehng, J.-M.; Deo, G.; Weckhuysen, B. M.; Guliants, V. V.; Benziger, J. B.; Sundaresan, S. Fundamental Studies of Butane Oxidation over Model-Supported Vanadium Oxide Catalysts: Molecular Structure-Reactivity Relationships. J. Catal. 1997, 170, 7588,  DOI: 10.1006/jcat.1997.1742
    31. 31
      Klug, C. A.; Kroeker, S.; Aguiar, P. M.; Zhou, M.; Stec, D. F.; Wachs, I. E. Insights into Oxygen Exchange between Gaseous O2 and Supported Vanadium Oxide Catalysts Via 17O NMR. Chem. Mater. 2009, 21, 41274134,  DOI: 10.1021/cm802680x
    32. 32
      Avdeev, V. I.; Bedilo, A. F. Molecular Mechanism of Oxygen Isotopic Exchange over Supported Vanadium Oxide Catalyst VOx/TiO2. J. Phys. Chem. C 2013, 117, 28792887,  DOI: 10.1021/jp311322b
    33. 33
      Avdeev, V. I.; Bedilo, A. F. Electronic Structure of Oxygen Radicals on the Surface of VOx/TiO2 Catalysts and Their Role in Oxygen Isotopic Exchange. J. Phys. Chem. C 2013, 117, 1470114709,  DOI: 10.1021/jp404921d
    34. 34
      Koranne, M. M.; Goodwin, J. G.; Marcelin, G. Oxygen Involvement in the Partial Oxidation of Methane on Supported and Unsupported V2O5. J. Catal. 1994, 148, 378387,  DOI: 10.1006/jcat.1994.1218
    35. 35
      Kera, Y.; Teratani, S.; Hirota, K. Infrared Spectra of Surface V═O Bond of Vanadium Pentoxide. Bull. Chem. Soc. Jpn. 1967, 40, 24582458,  DOI: 10.1246/bcsj.40.2458
    36. 36
      Hirota, K.; Kera, Y.; Teratani, S. Carbon Monoxide Oxidation with an Oxygen Tracer over a Vanadium Pentoxide Catalyst. J. Phys. Chem. 1968, 72, 31333141,  DOI: 10.1021/j100855a010
    37. 37
      Kera, Y.; Hirota, K. Infrared Spectroscopic Study of Oxygen Species in Vanadium Pentoxide with Reference to Its Activity in Catalytic Oxidation. J. Phys. Chem. 1969, 73, 39733981,  DOI: 10.1021/j100845a070
    38. 38
      Fletcher, W. H.; Rayside, J. S. High Resolution Vibrational Raman Spectrum of Oxygen. J. Raman Spectrosc. 1974, 2, 314,  DOI: 10.1002/jrs.1250020102
    39. 39
      Pushkarev, V. V.; Kovalchuk, V. I.; d’Itri, J. L. Probing Defect Sites on the CeO2 Surface with Dioxygen. J. Phys. Chem. B 2004, 108, 53415348,  DOI: 10.1021/jp0311254
    40. 40
      Choi, Y. M.; Abernathy, H.; Chen, H.-T.; Lin, M. C.; Liu, M. Characterization of O2–CeO2 Interactions Using in Situ Raman Spectroscopy and First-Principle Calculations. ChemPhysChem 2006, 7, 19571963,  DOI: 10.1002/cphc.200600190
    41. 41
      Wu, Z.; Dai, S.; Overbury, S. H. Multiwavelength Raman Spectroscopic Study of Silica-Supported Vanadium Oxide Catalysts. J. Phys. Chem. C 2010, 114, 412422,  DOI: 10.1021/jp9084876
    42. 42
      Weckhuysen, B. M.; Jehng, J.-M.; Wachs, I. E. In Situ Raman Spectroscopy of Supported Transition Metal Oxide Catalysts: 18O216O2 Isotopic Labeling Studies. J. Phys. Chem. B 2000, 104, 73827387,  DOI: 10.1021/jp000055n
    43. 43
      Lee, E. L.; Wachs, I. E. In Situ Raman Spectroscopy of SiO2-Supported Transition Metal Oxide Catalysts: An Isotopic 18O-16O Exchange Study. J. Phys. Chem. C 2008, 112, 64876498,  DOI: 10.1021/jp076485w
    44. 44
      Moisii, C.; van de Burgt, L. J.; Stiegman, A. E. Resonance Raman Spectroscopy of Discrete Silica-Supported Vanadium Oxide. Chem. Mater. 2008, 20, 39273935,  DOI: 10.1021/cm800095g
    45. 45
      Magg, N. Vibrational Spectra of Alumina- and Silica-Supported Vanadia Revisited: An Experimental and Theoretical Model Catalyst Study. J. Catal. 2004, 226, 88100,  DOI: 10.1016/j.jcat.2004.04.021
    46. 46
      Döbler, J.; Pritzsche, M.; Sauer, J. Vibrations of Silica Supported Vanadia: Variation with Particle Size and Local Surface Structure. J. Phys. Chem. C 2009, 113, 1245412464,  DOI: 10.1021/jp901774t
    47. 47
      Nitsche, D.; Hess, C. Normal Mode Analysis of Silica-Supported Vanadium Oxide Catalysts: Comparison of Theory with Experiment. Catal. Commun. 2014, 52, 4044,  DOI: 10.1016/j.catcom.2014.04.008
    48. 48
      Oyama, S. T.; Went, G. T.; Lewis, K. B.; Bell, A. T.; Somorjai, G. A. Oxygen Chemisorption and Laser Raman Spectroscopy of Unsupported and Silica-Supported Vanadium Oxide Catalysts. J. Phys. Chem. 1989, 93, 67866790,  DOI: 10.1021/j100355a041
    49. 49
      Ono, T.; Tanaka, Y.; Takeuchi, T.; Yamamoto, K. Characterization of K-Mixed V2O5 Catalyst and Oxidative Dehydrogenation of Propane on It. J. Mol. Catal. A Chem. 2000, 159, 293300,  DOI: 10.1016/S1381-1169(00)00218-1
    50. 50
      Ono, T.; Numata, H. Characteristic Features of Raman Band Shifts of Vanadium Oxide Catalysts Exchanged with the 18O Tracer and Active Sites for Reoxidation. J. Mol. Catal. A Chem. 1997, 116, 421429,  DOI: 10.1016/S1381-1169(96)00425-6
    51. 51
      Browne, M. P.; Sofer, Z.; Pumera, M. Layered and Two Dimensional Metal Oxides for Electrochemical Energy Conversion. Energy Environ. Sci. 2019, 12, 4158,  DOI: 10.1039/C8EE02495B
    52. 52
      Wachs, I. E. Recent Conceptual Advances in the Catalysis Science of Mixed Metal Oxide Catalytic Materials. Catal. Today 2005, 100, 7994,  DOI: 10.1016/j.cattod.2004.12.019
    53. 53
      Grant, J. T.; Venegas, J. M.; McDermott, W. P.; Hermans, I. Aerobic Oxidations of Light Alkanes over Solid Metal Oxide Catalysts. Chem. Rev. 2018, 118, 27692815,  DOI: 10.1021/acs.chemrev.7b00236
    54. 54
      Kuba, S.; Knözinger, H. Time-Resolved in Situ Raman Spectroscopy of Working Catalysts: Sulfated and Tungstated Zirconia. J. Raman Spectrosc. 2002, 33, 325332,  DOI: 10.1002/jrs.815
    55. 55
      Giannozzi, P. Quantum Espresso: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter 2009, 21, 395502,  DOI: 10.1088/0953-8984/21/39/395502
    56. 56
      Shklover, V.; Haibach, T.; Ried, F.; Nesper, R.; Novák, P. Crystal Structure of the Product of Mg2+ Insertion into V2O5 single Crystals. J. Solid State Chem. 1996, 123, 317323,  DOI: 10.1006/jssc.1996.0186
    57. 57
      Grimme, S. Semiempirical Gga-Type Density Functional Constructed with a Long-Range Dispersion Correction. J. Comput. Chem. 2006, 27, 17871799,  DOI: 10.1002/jcc.20495
    58. 58
      Barone, V.; Casarin, M.; Forrer, D.; Pavone, M.; Sambi, M.; Vittadini, A. Role and Effective Treatment of Dispersive Forces in Materials: Polyethylene and Graphite Crystals as Test Cases. J. Comput. Chem. 2009, 30, 934939,  DOI: 10.1002/jcc.21112
    59. 59
      Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 51885192,  DOI: 10.1103/PhysRevB.13.5188
    60. 60
      Kim, H.; Kosuda, K. M.; Van Duyne, R. P.; Stair, P. C. Resonance Raman and Surface- and Tip-Enhanced Raman Spectroscopy Methods to Study Solid Catalysts and Heterogeneous Catalytic Reactions. Chem. Soc. Rev. 2010, 39, 48204844,  DOI: 10.1039/c0cs00044b
    61. 61
      Harima, H. Raman Scattering Characterization on SiC. Microelectron. Eng. 2006, 83, 126129,  DOI: 10.1016/j.mee.2005.10.037
    62. 62
      Gilson, T. R.; Bizri, O. F.; Cheetham, N. Single-Crystal Raman and Infrared Spectra of Vanadium(V) Oxide. J. Chem. Soc., Dalton Trans. 1973, 291294,  DOI: 10.1039/dt9730000291
    63. 63
      Abello, L.; Husson, E.; Repelin, Y.; Lucazeau, G. Vibrational Spectra and Valence Force Field of Crystalline V2O5. Spectrochim. Acta, Part A 1983, 39, 641651,  DOI: 10.1016/0584-8539(83)80040-3
    64. 64
      Clauws, P.; Broeckx, J.; Vennik, J. Lattice Vibrations of V2O5. Calculation of Normal Vibrations in a Urey-Bradley Force Field. Phys. Status Solidi B 1985, 131, 459473,  DOI: 10.1002/pssb.2221310207
    65. 65
      Brázdová, V.; Ganduglia-Pirovano, M. V.; Sauer, J. Periodic Density Functional Study on Structural and Vibrational Properties of Vanadium Oxide Aggregates. Phys. Rev. B 2004, 69, 165420,  DOI: 10.1103/PhysRevB.69.165420
    66. 66
      Zhou, B.; He, D. Raman Spectrum of Vanadium Pentoxide from Density-Functional Perturbation Theory. J. Raman Spectrosc. 2008, 39, 14751481,  DOI: 10.1002/jrs.2025
    67. 67
      Wu, Z.; Kim, H.-S.; Stair, P. C.; Rugmini, S.; Jackson, S. D. On the Structure of Vanadium Oxide Supported on Aluminas: UV and Visible Raman Spectroscopy, UV–Visible Diffuse Reflectance Spectroscopy, and Temperature-Programmed Reduction Studies. J. Phys. Chem. B 2005, 109, 27932800,  DOI: 10.1021/jp046011m
    68. 68
      Hermann, K.; Witko, M.; Druzinic, R.; Tokarz, R. Hydrogen Assisted Oxygen Desorption from the V2O5(010) Surface. Top. Catal. 2000, 11/12, 6775,  DOI: 10.1023/A:1027206705195
    69. 69
      Huang, Y.-L.; Pellegrinelli, C.; Wachsman, E. D. Reaction Kinetics of Gas–Solid Exchange Using Gas Phase Isotopic Oxygen Exchange. ACS Catal. 2016, 6, 60256032,  DOI: 10.1021/acscatal.6b01462
    70. 70
      Huang, Y.-L.; Pellegrinelli, C.; Sakbodin, M.; Wachsman, E. D. Molecular Reactions of O2 and CO2 on Ionically Conducting Catalyst. ACS Catal. 2018, 8, 12311237,  DOI: 10.1021/acscatal.7b03467
    71. 71
      Huang, Y.-L.; Pellegrinelli, C.; Wachsman, E. D. Oxygen Dissociation Kinetics of Concurrent Heterogeneous Reactions on Metal Oxides. ACS Catal. 2017, 7, 57665772,  DOI: 10.1021/acscatal.7b01096
    72. 72
      Tilley, R. J. Understanding Solids: The Science of Materials; Wiley: 2013.
    73. 73
      Cameron, W. C.; Parkas, A.; Litz, L. M. Exchange of Isotopic Oxygen between Vanadium Pentoxide, Gaseous Oxygen and Water. J. Phys. Chem. 1953, 57, 229238,  DOI: 10.1021/j150503a022
    74. 74
      Milan, E. F. The Dissociation Pressure of Vanadium Pentoxide. J. Phys. Chem. 1929, 33, 498508,  DOI: 10.1021/j150298a002
    75. 75
      Nasu, N. The Dissociation Pressure of Vanadium Pentoxide. J. Chem. Soc. Jpn. 1935, 56, 666669,  DOI: 10.1246/nikkashi1921.56.6_666
    76. 76
      Spitsyn, B. N.; Maidanovskaya, L. L. The Thermal Dissociation of Vanadium Pentoxide. Zh. Fiz. Khim. 1959, 33, 180183
    77. 77
      Chen, K.; Khodakov, A.; Yang, J.; Bell, A. T.; Iglesia, E. Isotopic Tracer and Kinetic Studies of Oxidative Dehydrogenation Pathways on Vanadium Oxide Catalysts. J. Catal. 1999, 186, 325333,  DOI: 10.1006/jcat.1999.2510
    78. 78
      Ma, W. Y.; Zhou, B.; Wang, J. F.; Zhang, X. D.; Jiang, Z. Y. Effect of Oxygen Vacancy on Li-Ion Diffusion in a V2O5 Cathode: A First-Principles Study. J. Phys. D Appl. Phys. 2013, 46, 105306,  DOI: 10.1088/0022-3727/46/10/105306
    79. 79
      Mars, P.; van Krevelen, D. W. Oxidations Carried out by Means of Vanadium Oxide Catalysts. Chem. Eng. Sci. 1954, 3, 4159,  DOI: 10.1016/S0009-2509(54)80005-4
    80. 80
      Routray, K.; Reddy, K. R. S. K.; Deo, G. Oxidative Dehydrogenation of Propane on V2O5/Al2O3 and V2O5/TiO2 Catalysts: Understanding the Effect of Support by Parameter Estimation. Appl. Catal. A Gen. 2004, 265, 103113,  DOI: 10.1016/j.apcata.2004.01.006
    81. 81
      Alexopoulos, K.; Reyniers, M.-F.; Marin, G. B. Reaction Path Analysis of Propane Selective Oxidation over V2O5 and V2O5/TiO2. J. Catal. 2012, 289, 127139,  DOI: 10.1016/j.jcat.2012.01.019
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpcc.2c00174.

    • XRD pattern of V2O5, TG and DTA analysis of V2O5, HAADF-STEM images of V2O5, description of multiwavelength Raman spectroscopy, mass spectra of oxygen exchange of V2O5, uncorrected Raman spectra of V2O5, the atomic displacement patterns of Raman active phonon modes, Raman spectra of V2O5 after treatment in the presence of C3H8 and 16O2 at 573 °C, table of calculated lattice parameters and bond lengths, table of atomic coordinates of the optimized structure, description of the deconvolution of Raman spectra, discussions of surface reaction and bulk diffusion, and operando Raman spectroscopy of V2O5 during propane oxidation (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.