1 s2.0 S0021951713000614 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Journal of Catalysis 302 (2013) 3136

Contents lists available at SciVerse ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Structureactivity relationships for propane oxidative dehydrogenation by anatase-supported vanadium oxide monomers and dimers
Lei Cheng a, Glen Allen Ferguson a, Stan A. Zygmunt c, Larry A. Curtiss a,b,
a

Materials Science Division, Argonne National Laboratory, Argonne, IL 60439, United States Center for Nanoscale Materials, Argonne National Laboratory, Argonne, IL 60439, United States c Department of Physics and Astronomy, Valparaiso University, Valparaiso, IN 46383, United States
b

a r t i c l e

i n f o

a b s t r a c t
To understand the importance of the effect of molecular structure on reactivity, we have studied the activity of anatase TiO2 (0 0 1) supported VOx catalytic sites for propane oxidative dehydrogenation (ODH). First, possible structures of monomeric and dimeric VOx species on anatase (0 0 1) after VO4H3 grafting and water elimination were determined. We then studied the conversion reaction of propane to propanol by the supported VOx to elucidate the structurereactivity relationship. The coordination number of the vanadium atom was the key structural parameter in predicting the catalytic activity. This key structural difference alone resulted in an increase of up to 800 times in the reaction rate of CAH bond activation (rate-determining for propane ODH) for the various vanadium oxide species at 600 K. These results demonstrate the remarkable sensitivity of the catalytic site activity to its geometric structure and its implications for achieving optimal catalyst performance. 2013 Elsevier Inc. All rights reserved.

Article history: Received 5 October 2012 Revised 6 December 2012 Accepted 13 February 2013 Available online 1 April 2013 Keywords: Propane oxidative dehydrogenation Supported vanadium oxide DFT CAH activation

1. Introduction Supported vanadium oxide catalysts have been used industrially for many years to promote selective oxidation reactions [1] and have recently attracted special attention due to their exceptional activity and selectivity for the oxidative dehydrogenation (ODH) of light alkanes to alkenes. Dispersed vanadium oxide (VOx) can be prepared by anchoring VOx species atop a metal oxide support such as anatase (TiO2) at low vanadium oxide loading, and these dispersed catalysts exhibit higher alkene selectivity than crystalline V2O5 for ethane and propane ODH. However, due to the difculty of characterizing supported VOx structures in the submonolayer regime, the elucidation of the catalyst structureactivity relationship, a long-standing goal of catalysis science, has not yet been achieved. Many early attempts to characterize the surface structure of supported VOx species using techniques designed for bulk systems such as EXAFS, XANES, and 51V MAS-NMR led to conicting results [27]. In contrast, Raman spectroscopy provides outstanding structural information for metal oxide catalytic sites at a molecular level [816]. With decades of experimental studies [216] and more recently computational efforts [13,1720], it is generally accepted that a tetrahedral structure with one oxo group and three oxygen
Corresponding author at: Materials Science Division and Center for Nanoscale Materials, Argonne National Laboratory, 9700 Cass Ave., Argonne, IL 60439, United States. E-mail address: curtiss@anl.gov (L.A. Curtiss).
0021-9517/$ - see front matter 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.jcat.2013.02.012

atoms bridging to the support is the most likely structure for the VOx surface site. However, the catalytic surface is inhomogeneous. This distribution of surface structures results from variations in catalyst preparation conditions that produce different thermodynamically stable surface species. Since the species are stable by either having a similar thermodynamic stability or by being kinetically trapped, they can coexist on the substrate. In a recent study [13], we reported three distinct vanadium oxide structures on the h-alumina surface using a deep-UV Raman spectroscopy supported by DFT calculations. These structures result from different degrees of dehydration during the preparation process. Similar phenomena have been observed on silica-supported vanadium oxide [8]. More importantly, these different species were found to have differences in reducibility [21]. Therefore, when we evaluate the catalytic activities of these materials, it is important to distinguish different reactivities of each individual structure and identify the most active species. This information is critical for the rational design and synthesis of catalysts with high activity and selectivity. The CAH bond activation through hydrogen abstraction by the oxo group has been identied as the key step for alkane ODH to alkene on V2O5 and supported vanadium oxide in several mechanistic studies [2226]. Cheng et al. examined propane ODH with a V4O10 cluster and found that the rate-determining step is hydrogen abstraction by the vanadium oxo group forming an isopropyl radical [23]. A similar conclusion was reached by Rozanska et al. for propane ODH on SiO2-supported vanadium oxide [22]. These studies provided valuable information on the reaction mechanism, but

32

L. Cheng et al. / Journal of Catalysis 302 (2013) 3136

did not consider that some hydroxyl groups may remain on the support surface or on the vanadium oxide [8,13] and play an important role in changing the reactivity of the catalyst. In this work, we present the activities of various possible monomeric and dimeric VOx structures supported on an anatase (0 0 1) surface. These structures were constructed by grafting a VO4H3 precursor molecule(s) on the hydroxylated anatase surface and eliminating one or more water molecules to nd the possible stable structures on the surface. The reaction energies and barriers from propane up to propanol formation on selected monomer and dimer structures were calculated to show the structure reactivity relationship of the catalyst toward propane ODH. The exact minimum energy crossing point (MECP) between two electronic states on the reaction path has also been located and, to our knowledge, is reported here for the rst time. More importantly, the calculations of reaction energies and barriers using more realistic monomer and dimer models provide insight into the effect of the structure on catalyst reactivity.

vanadia active sites with vanadium in tetrahedral and square pyramidal coordination spheres, respectively. In structure calculations involving these two clusters, all atoms were allowed to relax. We also used a cluster, namely VO5H5_bulk, as the smallest model to represent the crystalline V2O5 (0 1 0) surface. The structure of VO5H5_bulk is very similar to VO5H5, except that its oxygen atoms were frozen at experimentally derived oxygen positions for V2O5 bulk in all calculations, and all OAH bond lengths in the terminating hydroxyl and aqua groups were also xed at 0.96 , in accordance with our previous work [27]. All other atoms were allowed to relax in calculations with VO5H5_bulk. Key structural parameters of these three clusters are also shown in Fig. 1 to illustrate the differences.

2. Theoretical models and methods 2.1. Models 2.1.1. Gas-phase vanadium oxide models The structure of the smallest models of active sites, VO4H3 and VO5H5, is shown in Fig. 1. VO4H3 and VO5H5 represent gas-phase

2.1.2. Anatase-supported vanadium oxide models Experimentally, vanadium oxide species are anchored atop support surfaces using a variety of preparation methods [1]. When this process is carried out in an aqueous environment, we expect the anatase surface to be hydroxylated. Computational studies have shown [17,28], and our own work conrms, that water readily dissociates upon adsorption on the (0 0 1) anatase surface by breaking one TiAO bond and forming two TiAOH bonds with enhanced stability due to hydrogen bonding. A previous experimental study [29] also indicated more hydroxyl groups on the anatase support led to the formation of more segregated VOx clusters instead of polymeric chains. We used a hydrogen-terminated cluster H20O18Ti4, shown

1.56 1.82 1.77 1.77 1.77 1.78

1.57 1.82 (1.91) (1.75) 2.27

1.57 (1.91)

(2.10)

VO4H3

VO5H5

VO5H5_bulk

Fig. 1. Structures of gas-phase vanadia species VO4H3, VO5H5 and VO5H5_bulk. Bond lengths in gure are in and numbers shown in parenthesis are xed during calculations.

VO4H3

+
+3H2O M-bidentate_2

H20O18Ti4 +H2O M-molecular

+3H2O M-dioxo +H2O M-bidentate M-tridentate +2H2O +3H2O

M-monodentate

Fig. 2. Illustration of the grafting of a VO4H3 precursor on hydroxylated anatase (0 0 1) forming various supported structures.

L. Cheng et al. / Journal of Catalysis 302 (2013) 3136

33

in Fig. 2, to represent the hydroxylated (0 0 1) surface of anatase. Either during the VO4H3 precursor grafting process or during subsequent calcination of the catalyst, the hydroxyl groups on the support are likely to condense as water and be driven off. It is possible that some surface hydroxyl groups remain under reaction conditions. To explore this possibility, VOx/TiO2 structures were considered with and without hydroxyl groups. Based on the work of Vittadini and Selloni [17], the grafting of VOx units on the surface was envisioned as proceeding by the interaction of VO4H3 molecular precursors with the hydroxyl groups on the TiO2 surface. This step was followed by a series of water eliminations illustrated in Fig. 2. In calculations of the supported clusters, constrained geometry optimizations were performed in which the terminal hydroxyl and aqua groups were xed. The terminal oxygen atoms were frozen at experimentally observed positions [30], and the H atoms were frozen at an hydroxyl distance of 0.96 along the direction of the titanium atom in the experimental structure. 2.2. Theoretical methods Structures reported in this paper were optimized using the B3LYP [31] hybrid density functional method with the 6-31G(d) basis set. For small cluster models VO4H3 and VO5H5, single point calculations using CCSD(T) [3235] and 6-31G(d) basis set combinations were also performed in order to assess and correct the B3LYP/6-31G(d) results. The GAUSSIAN 09 suite of programs [36] was used for these calculations. We employed a fragment guess technique to nd transition states on the open-shell singlet potential energy surface. The reaction path involves crossing of two states of different spin multiplicities. We used the method of Bearpark et al. [37,38] to locate the minimum energy crossing points (MECPs). 3. Results and discussion 3.1. Equilibrium structures and relative energetics of supported monomers and dimers The rst step in modeling propane ODH by supported VOx catalysts was to determine the geometries and relative energies of monomeric (M) and dimeric (D) VOx structures on the TiO2 support. The structures of six monomeric species (M-molecular, M-monodentate, M-bidentate, M-bidentate_2, M-dioxo and M-tridentate) and two dimeric species (D-molecular and D-bidentate) that can be formed during grafting of VO4H3 are shown in Figs. 2 and 3, respectively. The electronic energies and Gibbs free energies relative to gas-phase VO4H3 and hydroxylated anatase (H20O18Ti4) at a water partial pressure of 0.001 atm (typical catalytic condition) and temperatures of 298, 600 and 800 K, respectively, are reported in Table 1. These energies are also plotted in a graph shown in Fig. S2. The relative stabilities of the species are temperature dependent. As the temperature increases, the formation of Mmolecular and M-mono species becomes less favorable (less exothermic), while the formation of other species becomes more favorable (more exothermic or less endothermic). At 600 K, the formations of all structures are thermodynamically favorable (exothermic reaction) except for the M-dioxo species. Therefore, the M-dioxo structure is not likely to be formed in signicant quantities on the support. The energies reported in Table 1 are based solely on calculations of the stoichiometric formation reaction. The relative population of these species can also be adjusted by other factors such as nature of precursors or water concentration. The vanadium atoms in the M-monodentate, M-bidentate, M-bidentate_2, and D-bidentate structures have tetrahedral

2VO4H3

+
H20O18Ti4

+3H2O

+5H2O

D-molecular

D-bidentate

Fig. 3. Grafting of two VO4H3 precursors to form VOx dimers on anatase (0 0 1).

congurations with an oxo group and are also bound to three other oxygen atoms, in agreement with most supported V2O5 cluster results in the literature [8,13,1720]. The coordination of the vanadium atoms in the M-molecular, M-tridentate, and D-molecular is somewhat unique: apart from the oxo group, vanadium is also coordinated with four, rather than three, other oxygen atoms, resulting in a distorted square pyramidal geometry. A square pyramidal geometry is not uncommon for vanadium. In coordination complexes, the square pyramidal structure is especially preferred when the metal has one ligand with a multiple bond, like in VV and VIV coordination complexes such as VOCl3acetonitrile adduct and VO(acac)2 (vanadyl acetylacetonate). The square pyramidal geometry was also observed recently in ab initio molecular dynamics simulations of anatase (0 0 1)-supported vanadia under hydration conditions [28] and was also found to be the most stable vanadium oxide monomeric conguration on anatase (0 0 1) support in a periodic DFT calculation [39]. The V@O bond lengths in all structures reported in Figs. 2 and 3 are very similar, being in the range of 1.561.59 . 3.2. Reaction energies of propane to propanol by ODH on selected monomeric and dimeric structures Propanol formation from propane by supported vanadia involves two steps: propane CAH bond activation (or hydrogen abstraction) by the vanadyl group and OAC bond formation, as illustrated in Fig. 4. The relative reaction energies and barriers for the two steps are reported in Table 2. The rst hydrogen abstraction by a vanadyl group starts with a singlet ground state I: C3H8 + O@V5+ and ends with a diradical intermediate II: C3 H7 AHOAV4 . For all monomer structures reported in Table 2, the diradical intermediates have two nearly degenerate electronic states: the triplet (T) and the open-shell singlet (OS). For the M-molecular, M-tridentate, and M-bidentate clusters, the OS state of II is slightly lower in energy than the T, while for the M-bidentate_2, the OS state is slightly less stable than the T. The potential energy surfaces of the two states are highly parallel at this point of the reaction coordinate, and this is the point of the electronic curve crossing from the initial singlet ground state potential energy surface (PES) to the triplet PES. After the hydrogen abstraction, propanol C3H7OHAV3+ formation proceeds on the triplet PES because the triplet propanol has signicantly lower energy than the closedshell singlet for all the calculated structures, consistent with other theoretical work [22,23]. Fig. 4a illustrates the reaction energy prole of propane conversion to propanol on the M-molecular cluster, with all transition states and intermediates, as well as the MECPs. The crossing prior to diradical (II) formation, noted as MECP1 in

34

L. Cheng et al. / Journal of Catalysis 302 (2013) 3136 Table 1 Electronic energies (Ee) and Gibbs free energies (G) in kcal/mol of monomeric structures in Fig. 2 relative to gas-phase VO4H3 and hydroxylated anatase (H20O18Ti4) and dimeric structures in Fig. 3 relative to two gas-phase VO4H3 and hydroxylated anatase (H20O18Ti4). Energies in parenthesis are zero point corrected. Ee M-molecular + H2O M-mono + H2O M-bidentate + 2H2O M-dioxo + 3H2O M-bidentate_2 + 3H2O M-tridentate + 3H2O D-molecular + 3H2O D-bidentate + 5H2O 19.1 (18.7) 22.3 (21.8) 11.8 (13.5) 61.87 (57.5) 43.0 (38.6) 15.1 (11.7) 20.6 (22.1) 3.6 (9.7) G (298.15 K, 0.001 atm) 13.8 17.8 22.8 32.8 15.0 10.9 55.8 12.4 G (600 K, 0.001 atm) 9.4 14.1 33.4 5.1 11.0 35.7 30.8 79.6 G (800 K, 0.001 atm) 6.4 11.6 40.0 12.7 27.6 51.5 33.9 103.3

47.6 Triplet

MECP1 57.0 MECP2 35.7 39.8(34.6) TS1OS 35.6 35.1, OS

TS2OS 52.6 47.4 TS2T 38.6, CS

0.0, CS Singlet C3H8 +O=V5+ C3H7-HO-V4+

8.9 C3H7OH-V3+

(a)
61.1 Triplet MECP1 53.5 MECP2 44.6 41.8 41.2, OS TS2OS 57.7 52.2 TS2T 43.0, CS

43.9(39.0) TS1OS

18.8 0.0, CS Singlet C3H8 +O=V5+ C3H7-HO-V4+ C3H7OH-V3+

(b)
Fig. 4. The potential energy (kcal/mol) diagram of the conversion of propane to propanol on (a) the M-molecular cluster, and (b) the D-bidentate cluster. Numbers in parenthesis include the zero point energy (ZPE) correction. On the singlet potential energy curve the open-shell (OS) and closed-shell (CS) electronic conguration are specied.

Fig. 4, occurs at a much higher energy than both the TS1OS and the MECP2. At the crossing point geometry MECP1, the OS state energy is 55.3 kcal/mol, lower than the singlet/triplet energy at the same

crossing point geometry. This conrms that the open-shell singlet potential curve is lower in energy than the triplet and closed-shell singlet at this point on the reaction coordinate, and the reaction proceeds through the TS1OS state instead of crossing to the triplet PES, which occurs later in the reaction coordinate. Mulliken spin density analysis shows that in all the calculated TS1OS, II, and TS2 structures, spin is localized on the C and V atoms; while for the intermediate III structures, the spin is localized on the vanadium. The reduction in the support, a phenomenon that has been reported for reducible CeO2 support [40], was not observed on TiO2 in the current study. An alternative pathway for hydrogen abstraction by a VAOATi bridging site is less favorable than the abstraction by the vanadyl group for the systems in this study. One example is the reaction at the bridging oxygen of M-bidentate_2 cluster with activation barrier and reaction energy of 55.2 and 47.5 kcal/mol, respectively. These energies are much higher than the vanadyl pathway with an activation energy of 45.6 (TS1os) and a reaction energy of 42.2 kcal/ mol (II). The reaction mechanism on the two dimeric clusters is the same as on the monomeric structures reported above. The reaction energies and barriers of the two dimeric clusters are also both reported in Table 2, and the reaction energy prole on the D-bidentate cluster is illustrated in Fig. 4b. In all the calculated TS1OS, II, and TS2 structures, the spin is localized at the C and V atoms, while in the triplet intermediate III structures, the spin is localized on the reacting vanadium atom of the D-bidentate cluster. On the D-molecular structure, the spin is distributed between the two vanadium atoms. Comparing reaction energies and barriers on the two dimeric structures with the monomeric structures, we notice that these values of the D-bidentate are very similar to those of the monomeric structures M-bidentate and M-bidentate_2. For D-molecular, the TS1os energy is similar to that of the M-molecular (37.1 vs. 39.8 kcal/mol), while the II, TS2, and III energies are noticeably lower than those of the M-molecular. In order to evaluate the accuracy of these B3LYP/6-31G(d) results, we calculated the propanol formation energies and hydrogen abstraction barriers on VO5H5 and VO4H3 clusters using accurate wavefunction-based model chemistries. As presented in the Sup-

Table 2 Relative energies (in kcal/mol) of intermediates and transition states on reaction pathway of propanol formation on monomeric and dimeric supported VOx structures. Numbers in parenthesis include the zero point energy (ZPE) correction. Electronic conguration of each structure is also indicated: CS for closed-shell singlet, OS for open-shell singlet and T for triplet. Structure M-molecular M-tridentate M-bidentate M-bidentate_2 D-molecular D-bidentate TS1OS 39.8 40.1 44.4 45.6 37.1 43.9 (34.6) (35.3) (39.8) (40.5) (32.3) (39.0)
IV II C3 H 7 AHOAV

TS2 47.4(T), 45.7(T), 53.3(T), 55.3(T), 38.0(T), 52.2(T), 52.6(OS) 50.5(OS) 58.7(OS) 60.5(OS) 42.7(OS) 57.7(OS)

III (C3H7OHAVIII) 8.9(T), 1.7(T), 22.2(T), 23.0(T), 5.7(T), 18.8(T), 38.6(CS) 24.1(CS) 48.5(CS) 48.9(CS) 29.8(CS) 43.0(CS)

35.1(OS), 35.6(T) 37.4(OS), 39.1(T) 41.8(OS), 42.2(T) 41.9(T), 42.2(OS) 29.8(T), 30.0(OS) 41.2(OS), 41.8(T)

L. Cheng et al. / Journal of Catalysis 302 (2013) 3136

35

porting information, we found that the B3LYP/6-31G(d) model chemistry predicts energy differences between hydrogen abstraction barriers of different species in close agreement with more accurate methods. The difference between the VO5H5 and VO4H3 CAH activation barriers calculated by B3LYP/6-31G(d) is 0.2 kcal/ mol lower than that calculated by CCSD(T)/6-31G(d). 3.3. Structurereactivity relationship of supported VOx clusters To relate structure with reactivity of the monomeric clusters, we note that the active site VOx in both the M-bidentate and Mbidentate_2 structures has a tetrahedral coordination environment, one oxo group, two bridging O atoms bonded to Ti atoms, and one terminal hydroxyl group. The only structural difference between the two is that the support of the M-bidentate structure is hydroxylated, while that of the M-bidentate_2 is not. The fact that the reaction energies and barriers of the two structures are very similar, as shown in Table 2, indicates that the hydroxyl groups on the support surface have minimum effect on the reactivity of the catalytic site. The reactivity of the catalyst does not simply depend on the size of the active site either. The monomeric Mbidentate_2 shares a very similar reaction energy prole with the dimeric D-bidentate cluster, but differs signicantly from some other monomeric structures such as M-molecular and Mtridentate. The reactivity of the cluster does seem to be correlated to the coordination environment of the active V site: the vanadium atoms in the M-molecular and M-tridentate structures have square pyramidal coordination environment, and they both correspond to slightly lower TS1OS and intermediate II energies, as well as significantly lower TS2 and intermediate III energies, in comparison with the tetrahedral M-bidentate and M-bidentate_2. Similar trends were observed for the dimeric clusters as well: the V atom of the D-molecular cluster has a square pyramidal-like coordination, and it corresponds to a slightly lower TS1OS energy, as well as signicantly lower intermediate II, TS2, and intermediate III energies than the D-bidentate, in which the V has tetrahedral-like coordination. This is because when a V5+ is reduced to V4+ or V3+, the electron or electrons in d orbitals are more stabilized in a square pyramidal structure. If we consider ligand eld theory, when a V4+ or V3+ ion is placed in the surrounding oxygen ligand eld, the degeneracy of the metal d orbitals is broken. Due to the symmetry of the ligand eld, the overlap between a square pyramidal eld and the V d orbitals (dxz and dyz) is more effective than that with the tetrahedral eld. This leads to a larger decrease in orbital energies and thus more stabilization of the d electron(s) in the square pyramidal structure. This causes the M-molecular, M-tridentate and D-molecular clusters with the square pyramidal V to be more facile to oxidation state change and more active toward reduction and dehydrogenation reactions. 3.4. Discussion Despite the fact that TS2 structures are higher in energy than TS1OS structures for all VOx structures studied, the initial CAH activation is rate-limiting for the overall reaction of propane ODH by supported vanadia, as evidenced by experimental studies [2426] and explained with a kinetic scheme based on computational results for SiO2-supported VOx species by Rozanska et al. [22] The hydrogen abstraction barriers calculated for various monomeric and dimeric supported VOx structures are in the range of 32 40 kcal/mol. These values are comparable with the previously reported experimental (27.5 5 kcal/mol) [26] and theoretical values (34.4 kcal/mol) [22]. However, the TS1OS energies of the most and least active species reported in Table 2 differ by 8 kcal/mol, and this difference (as well as the differences between TS1OS energies

Table 3 Reaction barriers (kcal/mol) of hydrogen abstraction reaction on VO5H5 and VO4H3 clusters. Energies were calculated using B3LYP/ 6-31G(d) optimized structures with all atoms allowed to relax for the VO5H5 and VO4H3 clusters, and with the constraint that the terminating hydroxyl and aqua groups were xed at lattice V2O5 bulk distances for the VO5H5_bulk. Numbers in parenthesis include the zero point energy (ZPE) correction. Structure VO4H3 VO5H5 VO5H5_bulk TS1OS 47.8 (43.1) 38.5 (34.5) 47.5 (43.0)

of all other species) is the same when the entropic contributions are included as shown in Fig. S3. We also expect the collision frequencies between propane and all sites are the same. Since the prefactors are almost the same between all species, the 8 kcal/ mol difference in barriers leads to a difference of more than two orders of magnitude (150800 times) in reaction rate for these two species in the temperature range 600800 K. Therefore, structural differences could actually lead to signicant differences in the turnover frequency of the catalytic site. We believe that effective stabilization of reduced V species (V4+ and V3+) is the key to the catalytic ODH reaction activity of the vanadium oxide species. Previous studies of CeO2-supported VOx have shown that the origin of enhanced reactivity of the vanadia/ ceria system in Marsvan Krevelen-type oxidation reactions is the ability of ceria to stabilize reduced states by accommodating electrons in localized f-states [40,41]. While the ceria support showed a strong electronic effect that stabilizes reduced V, we nd that the square pyramidal structure stabilizes reduced V better than the tetrahedral structure, indicating a strong geometric/structural effect. This geometric effect on reactivity is reected in the simple small gas-phase cluster as well: the propane hydrogen abstraction barrier on the square pyramidal cluster VO5H5 is 9 kcal/mol lower than that of the tetrahedral cluster VO4H3 as reported in Table 3. This is consistent with the trend we have observed for the anatase-supported clusters, that is, the pyramidal coordinated vanadyl group is more active for hydrogen abstraction than the tetrahedral. In addition, the relaxation of the ligands surrounding the vanadium atom also plays a key factor in lowering the reaction barrier. For example, the barrier of the pyramidal cluster with terminating hydroxyl and aqua ligands xed at bulk V2O5 lattice distance (VO5H5_bulk in Table 3) corresponds to a barrier 9 kcal/mol higher than fully relaxed VO5H5 cluster or supported square pyramidal clusters. In this sense, segregated VOx clusters on support surfaces are less restricted geometrically and can thus undergo greater structural relaxation, in comparison with crystalline V2O5. The low barrier of 19.5 kcal/mol, as reported by Cheng et al. [23] for propane CAH activation by V2O5, may be partially due to the relaxation of V4O10 clusters used in their calculation.

4. Conclusions Density functional methods have been used to investigate structures of anatase-supported monomeric and dimeric VOx structures and their activity for propane oxidative dehydrogenation. We found that a variety of thermodynamically stable structures can coexist on the support surface including ones with a square pyramidal coordination environment and ones with tetrahedral environment. The relative populations will depend on temperature, hydration condition, etc. These different structures show signicantly different reactivities for propane ODH: the supported VOx structures with a square pyramidal coordination environment

36

L. Cheng et al. / Journal of Catalysis 302 (2013) 3136 [13] H.S. Kim, S.A. Zygmunt, P.C. Stair, P. Zapol, L.A. Curtiss, J. Phys. Chem. C 113 (2009) 88368843. [14] D.E. Keller, F.M.F. de Groot, D.C. Koningsberger, B.M. Weckhuysen, J. Phys. Chem. B 109 (2005) 1022310233. [15] C. Cristiani, P. Forzatti, G. Busca, J. Catal. 116 (1989) 586589. [16] J.L. Bronkema, A.T. Bell, J. Phys. Chem. C 111 (2007) 420430. [17] A. Vittadini, A. Selloni, J. Phys. Chem. B 108 (2004) 73377343. [18] T.K. Todorova, M.V. Ganduglia-Pirovano, J. Sauer, J. Phys. Chem. C 111 (2007) 51415153. [19] T.K. Todorova, M.V. Ganduglia-Pirovano, J. Sauer, J. Phys. Chem. B 109 (2005) 2352323531. [20] V. Brazdova, M.V. Ganduglia-Pirovano, J. Sauer, J. Phys. Chem. B 109 (2005) 2353223542. [21] H. Kim, G.A. Ferguson, L. Cheng, S.A. Zygmunt, P.C. Stair, L.A. Curtiss, J. Phys. Chem. C 116 (2012) 29272932. [22] X. Rozanska, R. Fortrie, J. Sauer, J. Phys. Chem. C 111 (2007) 60416050. [23] M.-J. Cheng, K. Chenoweth, J. Oxgaard, A. van Duin, W.A. Goddard III, J. Phys. Chem. C 111 (2007) 51155127. [24] K.D. Chen, A. Khodakov, J. Yang, A.T. Bell, E. Iglesia, J. Catal. 186 (1999) 325 333. [25] K.D. Chen, E. Iglesia, A.T. Bell, J. Catal. 192 (2000) 197203. [26] M.D. Argyle, K.D. Chen, A.T. Bell, E. Iglesia, J. Catal. 208 (2002) 139149. [27] P.C. Redfern, P. Zapol, M. Sternberg, S.P. Adiga, S.A. Zygmunt, L.A. Curtiss, J. Phys. Chem. B 110 (2006) 83638371. [28] A.E. Lewandowska, M. Calatayud, F. Tielens, M.A. Banares, J. Phys. Chem. C 115 (2011) 2413324142. [29] S.T. Choo, Y.G. Lee, I.S. Nam, S.W. Ham, J.B. Lee, Appl. Catal. A: Gen. 200 (2000) 177188. [30] M. Horn, C.F. Schwerdtfeger, E.P.Z. Meagher, Kristallogr. New Cryst. Struct. 136 (1972) 273281. [31] A.D. Becke, J. Chem. Phys. 98 (1993) 56485652. [32] G.E. Scuseria, H.F. Schaefer, J. Chem. Phys. 90 (1989) 37003703. [33] G.E. Scuseria, C.L. Janssen, H.F. Schaefer, J. Chem. Phys. 89 (1988) 73827387. [34] G.D. Purvis, R.J. Bartlett, J. Chem. Phys. 76 (1982) 19101918. [35] J.A. Pople, M. Headgordon, K. Raghavachari, J. Chem. Phys. 87 (1987) 5968 5975. [36] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N.J. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, . Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian, Inc., Wallingford, CT, 2009. [37] J.N. Harvey, M. Aschi, H. Schwarz, W. Koch, Theor. Chem. Acc. 99 (1998) 9599. [38] M.J. Bearpark, M.A. Robb, H.B. Schlegel, Chem. Phys. Lett. 223 (1994) 269274. [39] Y.J. Du, Z.H. Li, K.N. Fan, Surf. Sci. 606 (2012) 956964. [40] M.V. Ganduglia-Pirovano, C. Popa, J. Sauer, H. Abbott, A. Uhl, M. Baron, D. Stacchiola, O. Bondarchuk, S. Shaikhutdinov, H.J. Freund, J. Am. Chem. Soc. 132 (2010) 23452349. [41] C. Popa, M.V. Ganduglia-Pirovano, J. Sauer, J. Phys. Chem. C 115 (2011) 7399 7410.

being much more active for CAH bond activation. The origin of this difference in activity is that the square pyramidal coordination provides more effective stabilization of reduced V species in the reaction intermediate structures than the tetrahedral one. These results show that the coordination environment of the active vanadium site is a key structural parameter for its activity. This structural difference can result in an increase of 800 times in reaction rate of CAH activation at 600 K. Furthermore, the varied activities of catalytic structure coexisting on the support suggest that advanced catalyst synthesis technique to control site specicity could provide greatly improved catalytic efciency. Acknowledgments This research was supported by the U.S. Department of Energy, Ofce of Basic Energy Sciences, Division of Materials Sciences, Contract No. DE-AC-02-06CH11357. We also thank the Argonne Center for Nanoscale Materials for computing resources. We are grateful for discussion with Dr. Paul Redfern, Argonne National Laboratory. Appendix A. Supplementary material Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.jcat.2013.02.012. References
[1] B.M. Weckhuysen, D.E. Keller, Catal. Today 78 (2003) 2546. [2] B.M. Weckhuysen, J.M. Jehng, I.E.J. Wachs, Phys. Chem. B 104 (2000) 7382 7387. [3] T. Tanaka, H. Yamashita, R. Tsuchitani, T. Funabiki, S.J. Yoshida, Chem. Soc. Faraday Trans. I 84 (1988) 29872999. [4] M. Ruitenbeek, A.J. van Dillen, F.M.F. de Groot, I.E. Wachs, J.W. Geus, D.C. Koningsberger, Top. Catal. 10 (2000) 241254. [5] U.G. Nielsen, N.Y. Topsoe, M. Brorson, J. Skibsted, H.J. Jakobsen, J. Am. Chem. Soc. 126 (2004) 49264933. [6] R. Kozlowski, R.F. Pettifer, J.M. Thomas, J. Phys. Chem. 87 (1983) 51765181. [7] Y. Izumi, F. Kiyotaki, H. Yoshitake, K. Aika, T. Sugihara, T. Tatsumi, Y. Tanizawa, T. Shido, Y. Iwasawa, Chem. Commun. (2002) 24022403. [8] Z.L. Wu, S. Dai, S.H. Overbury, J. Phys. Chem. C 114 (2010) 412422. [9] G.T. Went, S.T. Oyama, A.T. Bell, J. Phys. Chem. 94 (1990) 42404246. [10] M.A. Vuurman, I.E. Wachs, J. Phys. Chem. 96 (1992) 50085016. [11] N. Magg, B. Immaraporn, J.B. Giorgi, T. Schroeder, M. Baumer, J. Dobler, Z.L. Wu, E. Kondratenko, M. Cherian, M. Baerns, P.C. Stair, J. Sauer, H.J. Freund, J. Catal. 226 (2004) 88. [12] E.L. Lee, I.E. Wachs, J. Phys. Chem. C 111 (2007) 1441014425.

You might also like