Sunteți pe pagina 1din 10

J. Phys. Chem.

A 2010, 114, 75617570

7561

Hydrolysis of TiCl4: Initial Steps in the Production of TiO2


Tsang-Hsiu Wang, Alejandra M. Navarrete-Lopez, Shenggang Li, and David A. Dixon*
Department of Chemistry, The UniVersity of Alabama, Shelby Hall, Box 870336, Tuscaloosa, Alabama 35487-0336

James L. Gole
Schools of Physics and Mechanical Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332-0430 ReceiVed: March 5, 2010; ReVised Manuscript ReceiVed: May 26, 2010

The hydrolysis of titanium tetrachloride (TiCl4) to produce titanium dioxide (TiO2) nanoparticles has been studied to provide insight into the mechanism for forming these nanoparticles. We provide calculations of the potential energy surfaces, the thermochemistry of the intermediates, and the reaction paths for the initial steps in the hydrolysis of TiCl4. We assess the role of the titanium oxychlorides (TixOyClz; x ) 2-4, y ) 1, 3-6, and z ) 2, 4, 6) and their viable reaction paths. Using transition-state theory and RRKM theory, we predicted rate constants including the effect of tunneling. Heats of formation at 0 and 298 K are predicted for TiCl4, TiCl3OH, TiOCl2, TiOClOH, TiCl2(OH)2, TiCl(OH)3, Ti(OH)4, and TiO2 using the CCSD(T) method with correlation consistent basis sets extrapolated to the complete basis set limit and compared with the available experimental data. Clustering energies and heats of formation are calculated for neutral clusters. The calculated heats of formation were used to study condensation reactions that eliminate HCl or H2O. The reaction energy is substantially endothermic if more than two HCl molecules are eliminated. The results show that the mechanisms leading to formation of TiO2 nanoparticles and larger ones are complicated and will have a strong dependence on the experimental conditions.
Introduction Titanium dioxide (TiO2) is technologically one of the most important compounds formed by the group IVB transition-metal elements. It is widely used as a white pigment, catalyst support, and photocatalyst.1-4 At room temperature, bulk TiO2 exists in three phases: rutile, anatase, and brookite.1 TiO2 as a photocatalyst has been used for solar energy conversion and for the removal of organic pollutants from wastewater.5-8 It is wellestablished that anatase TiO2 has a higher photocatalytic activity than the rutile or brookite phases. For example, one of the most active commercial TiO2 photocatalysts, Degussa P25, is 60-80% anatase phase.9 The pure anatase phase is thermodynamically less stable than rutile at room temperature, and it can undergo thermal conversion into the rutile phase in the temperature range of 700-800 C.10 The most important commercial route for the production of TiO2 nanoparticles and larger particles is based on the chloridebased process where puried TiCl4 is oxidized at high temperature (1200-1700 C) and modest pressure (300 kPa), in an oxygen plasma or ame, as given by reaction 1a.11,12 Reaction 1b corresponds to the same reaction for the formation of (TiO2)n nanoparticles. rutile phase as otherwise the anatase phase is formed.13,14 Introduction of a small amount of water is important in the industrial process to initiate the above reaction. The aqueous hydrolysis of TiCl4, as indicated in reaction 2a, represents another process to produce nanostructured polycrystalline TiO2.15 This process can be run at room temperature. TiO2 nanoparticles have also been generated by the vapor-phase hydrolysis of TiCl4 at temperatures in the range of 360-550 C with the TiCl4 present at 1% or lower by volume (reaction 2b). This process can yield anatase-phase particles because of the lower reaction temperature.16 Reaction 2c represents the formation of TiO2 nanoclusters in the gas-phase hydrolysis.

TiCl4(g) + 2H2O(aq) 9 TiO2(s,rutile) + 4HCl(aq) 8

298 K

(2a)
TiCl4(g) + 2H2O(g) 9 TiO2(s,rutile) + 4HCl(g) 8
630-830 K

(2b)
nTiCl4(g) + 2nH2O(g) f (TiO2)n(g) + 4nHCl(g)

TiCl4 + O2 f TiO2(s) + 2Cl2 nTiCl4 + nO2 f (TiO2)n + 2nCl2

(1a) (1b)

(2c)
TiO2 colloidal particles have been synthesized from titanium isopropoxide in aqueous acid where the acid leads to charging of the particles, which can enable the control of their growth.17 Nanocrystalline TiO2 lms have also been synthesized by a sol-gel method using reverse micelles formed by Triton X-100 and water in cyclohexane with titanium isopropoxide as the

In the industrial combustion process, AlCl3 is added in small quantities to the TiCl4 reactor feed to promote formation of the
* To whom correspondence should be addressed. E-mail: dadixon@ bama.ua.edu.

10.1021/jp102020h 2010 American Chemical Society Published on Web 06/24/2010

7562

J. Phys. Chem. A, Vol. 114, No. 28, 2010

Wang et al. step in the combustion system and is relevant to the hydrolysis mechanism as well. The formation of the species described herein can be coupled with those based on the thermal decomposition of TiCl4 in further studies of the formation of TiO2 nanoparticles. For a number of species, the thermodynamics have been obtained at the coupled cluster CCSD(T) level at the complete basis set (CBS) limit using approaches developed in our group in collaboration with Washington State University.32 We have previously used such an approach to study the thermodynamic properties of TiO2 and other transition-metal oxide clusters.19 The focus of the current study is on the species formed by the reaction of H2O with TiCl4 and their subsequent clustering reactions. Computational Methods The potential energy surfaces were initially calculated at the density functional theory level with the B3LYP exchangecorrelation functional33,34 and the aug-cc-pVDz/aug-cc-pVDZPP basis set described below. Equilibrium geometries and harmonic vibrational frequencies were calculated at the secondorder Mller-Plesset perturbation theory (MP2) level for various structures on the potential energy surface using the Gaussian 03 program.35 These calculations were done with the aug-cc-pVnZ basis sets36 for H and O, the aug-cc-pV(n+d)Z basis sets for Cl,37 and the small core effective core potential (ECP) based aug-cc-pVnZ-PP basis sets19,38 for Ti with n ) D and T. (We denote this combination of basis sets as aVnZ.) For the hydrates and transition states, the geometries calculated at the MP2/aVTZ level were subsequently used in single-point CCSD(T) calculations with the aVDZ, aVTZ, and aVQZ basis sets. The geometries of TiCl4, TiCl3OH, TiOCl2, TiOClOH, and TiO2 were also optimized at the CCSD(T) level with the aVDZ and aVTZ basis sets. The geometry calculated at the CCSD(T)/ aVTZ level was then used in single-point CCSD(T)/aVQZ calculation. All of the CCSD(T) calculations were performed with the MOLPRO 2006.1 program.39 The CCSD(T) energies were then extrapolated to the CBS limit using a mixed Gaussian/ exponential formula (eq 3).40 n ) 2, 3, and 4 for aVDZ, aVTZ, and aVQZ, respectively.

TABLE 1: Reaction Enthalpy (H, kcal/mol) at 298 K for Reactions 1a and 1ba
H298K reaction reaction reaction reaction reaction
a

H1500K 110.8 103.0 85.5 66.7 -41.5

G298K 103.7 96.6 77.8 65.1 -39.0

G1500K 64.9 64.3 44.4 59.8 -23.0

1b, 1b, 1b, 1b, 1a

n n n n

) ) ) )

1 2 3 4

114.2 105.4 86.6 66.7 -43.4

Heats of formation for (TiO2)n from ref 19 and for TiCl4 (-182.4 kcal/mol) and TiO2(s,rutile) from ref 21.

TABLE 2: Reaction Enthalpy (H, kcal/mol) at 298 K for Reactions 2a and 2ba
H298K H700K H1500K G298K G700K G1500K
reaction reaction reaction reaction reaction reaction 2c, 2c, 2c, 2c, 2b 2ab

n n n n

) ) ) )

1 2 3 4

141.4 159.8 171.7 175.5 -16.2 -66.6

140.8 159.9 172.8 177.4 -15.2

139.4 160.3 174.8 181.2 -12.8

121.8 132.9 135.8 137.7 -20.9 -51.0

95.7 96.6 86.8 85.8 -27.8

44.8 24.1 -12.4 -20.6 -43.1

a Heats of formation for (TiO2)n from ref 19 and for TiCl4 (-182.4 kcal/mol) and TiO2(s,rutile) from ref 21 unless noted otherwise. b Heats of formation from ref 22. If aqueous TiCl4 is used, then H ) 1.3 kcal/mol and the reaction is essentially thermoneutral.

reagent.18 An important issue is that both reactions 1b and 2c are endothermic to produce small (TiO2)n nanoparticles in the gas phase. Therefore, one must partially oxidize or hydrolyze the TiCl4 to build up an oxychloride/hydroxyoxychloride particle large enough so that the reaction thermodynamics begins to resemble that of the solid as evidenced in Table 1 for TiO2 clusters starting from Cl2 (reactions 1a and 1b) and in Table 2 starting from H2O.19-22 The enthalpy and free energy in Table 1 at 298 K show that, as the particle gets larger, the energy to form it decreases. At 1500 K, the structure of the individual particle begins to play a more important role as the presence of low-lying vibrational modes in the cluster can signicantly contribute to the entropy contribution to the free energy. In Table 2, the addition of H2O leads to more exothermic (in terms of the free energy) reactions as the temperature increases, especially as the nanoparticle size gets larger, due to the release of more HCl particles in the gas phase. The hydrolysis reaction to form rutile is overall exothermic, independent of whether the process takes place with gas-phase reactants or in the aqueous phase. The gas-phase processes for the generation of TiO2 have been modeled in terms of the oxidation of TiClx radicals which then cluster to form (TiO2)n nanoparticles for large n.23 In another case, TiOCl2 has been suggested to be a key intermediate.24,25 West et al.26 have predicted the thermochemical parameters for a number of TiOxCly intermediates using density functional theory (DFT)27 with different functionals and with coupled cluster28 CCSD(T)29 calculations using modest basis sets. These data were then used in kinetic models to predict the formation of TiO2 nanoparticles.14,30,31 The major reactions they studied were based on the thermal decomposition of TiCl4 to from the TiCl3 and Cl radicals. The compounds that we studied are similar to those given by Kraft and co-workers14,30,31 with the addition of the presence of hydroxyl groups, and their compounds could also play a role in the hydrolysis process. Although the combustion approach to the synthesis of TiO2 particles has been used in industry for many years, the overall mechanism is still poorly understood. The goal of this work is to study the potential energy surface for the initial steps in the gas-phase hydrolysis of TiCl4, which may serve as an initiation

E(n) ) ECBS + Be-(n-1) + Ce-(n-1)

(3)

To predict thermochemical properties to high accuracy, it is necessary to include additional corrections. For the complexes with one Ti atom, the core-valence correlation corrections (ECV) were obtained at the CCSD(T)-DK/aug-cc-pwCVTZDK level.38 For the calculation of transition-metal compound atomization energies, the core-valence calculations should be calculated at the CCSD(T)-DK level to achieve the best accuracy.19 In addition, we also account for relativistic effects in atoms and molecules. The rst is the spin-orbit correction (ESO), which lowers the sum of the atomic energies (decreasing D0) by replacing energies that correspond to an average over the available spin multiplets with energies for the lowest multiplets. This correction is required as most electronic structure codes produce only spin multiplet averaged wave functions. The experimental ground-state atomic spin-orbit corrections are ESO(Ti) ) 0.64 kcal/mol, ESO(Cl) ) 0.84 kcal/mol, and ESO(O) ) 0.22 kcal/mol.41 The second correction is the scalar relativistic correction (ESR). As we already have such a correction for Ti through the use of the ECP, we evaluated ESR for O and Cl as expectation values of the two dominant terms in the Breit-Pauli Hamiltonian, the mass-

Hydrolysis of TiCl4 velocity and one-electron Darwin (MVD) corrections, at the conguration interaction singles and doubles (CISD) level with the aVTZ basis set at the MP2/aVTZ geometry. Our computed total atomization energy (D0) values were obtained with eq 4.

J. Phys. Chem. A, Vol. 114, No. 28, 2010 7563


TABLE 3: Calculated and Experimental Ti-X (X ) Cl, O, OH) Bond Lengths () at the CCSD(T)/aVTZ or MP2/aVTZ Levels
molecule TiCl4
1

state A1/Td A1/C2V A/Cs A1/C2V A/Cs A/C1 A/C3 A/S4

method CCSD(T)/aVTZ exptla CCSD(T)/aVTZ B97-1/6-311+G(d,p)b CCSD(T)/aVTZ CCSD(T)/aVTZc exptld MP2/aVTZ MP2/aVTZ MP2/aVTZ MP2/aVTZ

Ti-Cl Ti-O Ti-OH 2.186 2.170 2.250 1.619 2.234 1.587 2.270 1.627 1.666 1.651 2.192 2.191 2.213 2.238

D0 ) ECBS + EZPE + ECV + ESR + ESO

(4)
Given the known heats of formation at 0 K for the elements, Hf,0K(Ti) ) 112.4 ( 0.7 kcal/mol,42 Hf,0K(Cl) ) 28.59 kcal/ mol,21 Hf,0K(O) ) 58.98 kcal/mol,21 and Hf,0K(H) ) 51.63 kcal/mol,21 we can derive Hf,0K values for the molecules under study. The heats of formation at 298 K can be obtained by following the procedures outlined by Curtiss et al.43 The temperature dependence of the enthalpy and the entropy calculations were done in the rigid-rotor, harmonic oscillator approximation with hindered rotors treated as a vibration.44 At higher temperatures, this may lead to modest errors in the thermodynamic properties due to the transition of a hindered rotor to a free rotor. The CCSD(T) method scales approximately as N7 with N basis functions, and large basis sets are required to reach the CBS limit. The approach described above for the CCSD(T)/ CBS calculations has been used to predict a wide range of thermodynamic properties to chemical accuracy.19,20,32,45 DFT has a much better scaling, scaling as N3 to N4 depending on the exchange-correlation functional, but is not as accurate as the CCSD(T)/CBS approach for the calculation of a broad range of thermodynamic properties including atomization energies. For small clusters with one Ti atom, accurate thermodynamic properties can be calculated using the CCSD(T)/CBS approach, which is more reliable than similar DFT-based approaches.19,20,46 In addition, the CCSD(T)/CBS values can be used to benchmark different DFT functionals in the future. We avoided the use of isodesmic and similar types of reaction energies as used previously because of issues with the experimental heats of formation of similar Ti compounds and because there are so few experimental values.47 For the larger clusters, we used the normalized clustering energy method that we have previously developed19,20 for the prediction of the heats of formation of transition-metal oxide clusters. All calculations were performed on a Xeon- and Opteronbased Penguin Computing Linux cluster in our group, the Itanium 2-based SGI Altix and the Opteron-based dense memory cluster Linux cluster at the Alabama Supercomputer Center, the Xeon-based Dell Linux cluster at The University of Alabama, and the Opteron-based Linux cluster at the Pacic Northwest National Laboratory. Results and Discussion Geometries and Frequencies. The calculated Ti-X (X ) Cl, O, OH) bond lengths at the CCSD(T)/aVTZ level for selected reactants, intermediates, and products are given in Table 3. They are compared with available experimental values. The electronic states and symmetry labels for these molecules are given in Table 3 and will not be repeated hereafter. The optimized molecular structures for these molecules together with those of the transition states are given in the potential energy surface plots (Figure 1). The total CCSD(T) energies, calculated harmonic frequencies compared with the available experimental values, and T1 diagnostics are given as Supporting Information. The calculated Ti-Cl bond distance of 2.186 at the CCSD(T)/aVTZ level in TiCl4 is slightly longer (<0.02 ) than

TiOCl2 TiOClOH TiO2 TiCl3OH TiCl2(OH)2 TiCl(OH)3 Ti(OH)4


a

1 1

1.843 1.767 1.784 1.775 1.797 1.813

1 1

Reference 48. b Reference 26. c Reference 19. d Reference 50.

the gas-phase electron diffraction48 value of 2.170 . Lower level calculations are also in agreement.49 The Ti-Cl bond distances in TiCl3OH, TiOCl2, TiOClOH, TiCl2(OH)2, and TiCl(OH)3 are predicted to lengthen from that in TiCl4 with an increase in the number of O atoms and with changes in the number of ligands bonded to the Ti atom. The Ti-Cl bond distance in TiOCl2 of 2.234 calculated26 at the B97-1/6311+G(d,p) level is shorter than our CCSD(T)/aVTZ value by 0.02 . The calculated Ti-O bond distance in the series from TiO2 to TiOClOH to TiOCl2 decreases as the number of ligands increases from two to three and Cl is substituted for OH. The Ti-O bond distance in TiOCl2 calculated at the B97-1/6311+G(d,p) level26 is shorter than our CCSD(T)/aVTZ value by 0.03 . The Ti-O bond distance in TiO2 has been measured by Fourier transform microwave spectroscopy to be 1.651 ;50 the CCSD(T)/aVTZ calculations overestimate this distance by 0.015 .19 The Ti-OH bond distance increases by 0.067 from TiCl3OH to TiOClOH. The role of multireference character in the wave function can be estimated from the T1 diagnostic51 for the CCSD calculation. The values for the T1 diagnostics are modest (<0.03), showing that the wave function is dominated by a single electron conguration. TiO2 has a slightly larger T1 diagnostic but is well-behaved on the basis of our recent work.19 Atomization Energies and Heats of Formation. The different energetic components for the total atomization energies (TAEs) are given in Table 4. The core-valence corrections for the Ti compounds are found to increase the TAEs by 1.3-4.8 kcal/mol. The scalar relativistic corrections for the Ti compounds are usually smaller than the core-valence corrections and are found to decrease the TAEs by 0.7-2.0 kcal/mol. The calculated heats of formation are also given in Table 4, and we use the values at 298 K in our discussion below unless specied otherwise. The estimated error bars for the calculated heats of formation are (1.5 kcal/mol. Here we include our estimates of the errors in the energy extrapolation, vibrational frequencies, and other electronic energy components, as well as the error associated with the heat of formation of the Ti atom based on our recent work.20 The calculated heat of formation of TiCl4 is in good agreement with the experimental value20 considering the experimental error limits of (0.7 kcal/mol for both Hf(TiCl4) and Hf(Ti). The calculated heat of formation of TiOCl2 is 11.4 kcal/mol more negative than the estimated value of -130.4 kcal/mol.21 For TiOCl2, the calculated heat of formation by Green and co-workers26 at the B97-1/6-311+G(d,p) level of -142.9 ( 4.8 kcal/mol is in good agreement with our

7564

J. Phys. Chem. A, Vol. 114, No. 28, 2010

Wang et al. TiCl4 can be considered in two stages (Figures 1a and 2). The rst stage can be described by the overall reaction given as follows, as shown in Figure 1a:

TiCl4 + H2O f TiOCl2 + 2HCl

(5)

The second stage is the reaction of TiOCl2 with a second water molecule shown in Figure 2:

TiOCl2 + H2O f TiO2 + 2HCl

(6)

Both overall reactions are substantially endothermic. The rst reaction step of the rst stage

TiCl4 + H2O f TiCl3OH + HCl

(5a)

is only slightly endothermic by 2.5 kcal/mol at 0 K. The overall reaction free energy becomes more favorable as the temperature increases mostly due to the difference in the entropies of TiCl4 and TiCl3OH. At 700 K, the reaction will proceed to the right with Keq 5. This step starts with an exothermic Lewis acid base addition of H2O to TiCl4, forming a stable complex. Complex formation is followed by a hydrogen transfer from water to a chlorine, leading to HCl elimination. The reaction barrier for the H transfer from the complex is 18.8 kcal/mol with 8.2 kcal/mol due to the endothermicity (0 K) of the process relative to the starting reactant complex and the formation of a hydrogen-bonded complex between HCl and TiCl3OH. The overall reaction is only weakly endothermic, so the reaction will proceed to a reasonable extent at the high processing temperatures. Transfer of the second H atom from the OH group to a second Cl atom to form the TiOCl2 HCl Lewis acid-base complex is endothermic by 33.4 kcal/mol, and the total barrier for H transfer of 35.7 kcal/mol is predominantly due to the endothermicity of this reaction. HCl forms a strong Lewis acid-base complex with TiOCl2 with a (H)Cl-Ti dative binding energy of 17.0 kcal/mol. The overall endothermicity of the unimolecular reaction of TiCl3OH to form TiOCl2 + HCl shows that the reaction will not likely proceed past the rst proton transfer step. The effects of temperature on this step are not large in terms of the enthalpy or free energy. A more likely process is that TiCl3OH reacts further with another water molecule (Figure 1b):

TiCl3OH + H2O f TiCl2(OH)2 + HCl

(7)

Figure 1. Potential energy surfaces (kcal/mol) for reactions (a) TiCl4 + H2O f TiOCl2 + 2HCl, (b) TiCl3OH + H2O, (c) TiCl2(OH)2 + H2O, and (d) TiCl(OH)3 + H2O. Key: black, E from heats of formation at 0 K; crimson, ECBS + ECV + EZPE; green, MP2/aVTZ + EZPE; blue, B3LYP/aVDZ + EZPE; yellow, Ti; red, O; green, Cl; white, H.

CCSD(T)/CBS value. These results suggest that the tabulated heat of formation for TiOCl2 needs to be reevaluated.21 The calculated thermodynamic properties as a function of temperature for use in other modeling schemes are given in the Supporting Information. Hydrolysis Potential Energy Surface. The calculated potential energy surfaces at 0 K of the hydrolysis of TiCl4 are shown in Figures 1 and 2. The initial reaction of water with

This reaction is endothermic by only 4.2 kcal/mol at 0 K. The entropy effects for this reaction are small, so there is only a slight decrease in the reaction free energy as the temperature increases. At equilibrium at 700 K, Keq 0.05, so the product TiCl2(OH)2 will be formed to some extent. As the product further reacts, Le Chateliers principle suggests that the product will continue to be formed. Again H2O forms a Lewis acid-base complex with the TiCl3OH, and there is a barrier of 18.4 kcal/mol for transfer of the proton, leading to a hydrogenbonded complex of HCl with TiCl2(OH)2. The hydrogen bond energy is 2.5 kcal/mol. TiCl2(OH)2 can further react with H2O to give a hydrogenbonded complex as part of the following overall reaction:

TiCl2(OH)2 + H2O f TiCl(OH)3 + HCl

(8)

Hydrolysis of TiCl4

J. Phys. Chem. A, Vol. 114, No. 28, 2010 7565

TABLE 4: Total Atomization Energies (D0,0K, kcal/mol) at 0 K and Heats of Formation at 0 and 298 K (Hf,0K and Hf,298K, kcal/mol) Calculated at the CCSD(T) Level
molecule TiCl4 TiCl3OH TiOCl2 TiOClOH TiO2i TiCl2(OH)2 TiCl(OH)3 Ti(OH)4 H2 O TiCl4 H2O TS1 TiCl3OH HCl HCl TS2 TiOCl2 HCl TiOCl2 H2O ECBSa 413.56 533.91 375.57 490.09 299.77 653.35 770.77 885.14 232.95 655.90 632.94 644.14 107.40 495.63 499.92 640.95 EZPEb -3.65 -10.27 -3.23 -10.41 -3.26 -17.32 -24.69 -31.81 -13.44 -19.22 -15.24 -15.93 -4.36 -7.80 -8.95 -19.01 ECVc 3.32 3.88 1.35 1.87 2.92 4.04 4.74 4.34 0.34 3.89 4.05 4.20 0.19 3.80 2.98 2.99 ESRd -1.46 -1.67 -1.17 -1.32 -0.78 -1.86 -2.00 -2.14 -0.29 -1.87 -1.88 -1.90 -0.23 -1.50 -1.48 -1.64 ESOe -4.00 -3.38 -2.54 -1.92 -1.08 -2.76 -2.14 -1.52 -0.22 -4.22 -4.22 -4.22 -0.84 -3.38 -3.38 -2.76 D0,0Kf 407.78 522.47 369.98 478.31 297.57 635.45 746.69 855.01 219.34 634.48 611.60 626.30 102.17 486.75 489.09 620.54 Hf,0Kg -181.0 -213.7 -141.4 -167.7 -67.2 -244.7 -273.9 -300.2 -57.1 -245.5 -226.6 -237.3 -21.9 -178.0 -180.3 -229.7 Hf,298Kg -181.5 -214.5 -141.8 -168.8 -67.8 -246.1 -276.2 -303.2 -57.8 -247.2 -228.2 -238.3 -22.0 -179.5 -181.0 -231.4 Hf,298K (exptl) -182.4 ( 0.7h -130.4h -73.0 ( 3i

-57.7978 ( 0.0096h

-22.06 ( 0.024i

a CCSD(T)/CBS energies obtained using eq 3. b MP2/aVTZ. c CCSD(T)/aug-cc-pwCVTZ-DK. d Expectation value of the MVD operators on the CISD/aVTZ wave function. e Experimental atomic spin orbit corrections from ref 41. f Equation 4. g See the text. h Reference 17. i Reference 15.

and HCl due to the large endothermicities inherent to these processes. The overall reaction energies for each of the reactions in the initial hydrolysis of TiCl4 are given in Table 5. Of course, any of the intermediates that are formed can react with other species in the plasma, for example, any O or Cl atoms that are formed, or with themselves as discussed below. For the following reactions, we used conventional transitionstate theory (TST)52 and Rice-Ramsperger-Kassel-Marcus (RRKM) theory53 to predict the rate coefcients (k):

TiCl4 H2O f TiCl3OH HCl


Figure 2. Potential energy surface (kcal/mol) for TiOCl2 + H2O. See the Figure 1 caption for details.

(5b) (7a) (8a) (9a)

TiCl3OH H2O f TiCl2(OH)2 HCl TiCl2(OH)2 H2O f TiCl(OH)3 HCl

Reaction 8 is less favorable energetically than reactions 5 and 7, and for this reaction at 700 K, Keq 8 10-3, so enough product will be formed for the reaction to continue. This complex undergoes a hydrogen transfer with a barrier of 17.9 kcal/mol to form a complex with HCl bonded to TiCl(OH)3 (Figure 1c). The H-bond energy is 4.6 kcal/mol, consistent with the HCl being involved in two H bonds, one as an acceptor and one as a donor. The overall reaction is endothermic by 6.0 kcal/mol at 0 K. The nal step is the reaction of TiCl(OH)3 with H2O:

TiCl(OH)3 H2O f Ti(OH)4 HCl

TiCl(OH)3 + H2O f Ti(OH)4 + HCl

(9)

It follows the same path of Lewis acid-base addition followed by H transfer with essentially the same energetics (Figure 1d). The nal H-bond energy for the HCl with Ti(OH)4 is 5.2 kcal/ mol. This reaction is the most endothermic for the hydrolysis reactions. We can briey summarize the reaction energetics for all steps as (1) initial complex formation with a complexation energy of about 9 kcal/mol and (2) a hydrogen transfer step with a barrier of about 18 kcal/mol to form a complex with a HCl and the Ti species with complexation energies of 2-4 kcal/mol and an overall reaction endothermicity of 2-9 kcal/mol at 0 K. Thus, the hydrolysis of TiCl4 will not form TiOCl2 and HCl or TiO2

The predicted rates as a function of temperature are given in the Supporting Information. The limiting rate constants from RRKM theory in the temperature range (T) of 273-373 K and pressure range (P) of 380-1520 Torr and in the temperature range of 200-2000 K and pressure range of 760-2250 Torr are given in Table 6. The rate constant for the lower temperature range is essentially independent of pressure. The effect of tunneling52,54,55 on the rate constants at room temperature leads to an increase of up to a factor of 2. We used the Skodje and Truhlar55 expression for the tunneling coefcient because it provides a better approximation. The rate constants at a temperature of 1500 K more like that in the combustion reactor are quite fast and exhibit a dependence on pressure up to about 3 atm. The reaction rates at 700 K representative of the hightemperature hydrolysis reaction process are also quite fast. At the higher temperatures used in the combustion or gas-phase hydrolysis processes, tunneling will not be important. The reaction rates are comparable within an order of magnitude for the displacement of HCl from tetracoordinate Ti at all temperatures. Unimolecular reaction 6a is predicted to be substantially slower at 298 and 700 K but is predicted to be within an order of magnitude of the other reactions at 1500 K. We also include

7566

J. Phys. Chem. A, Vol. 114, No. 28, 2010

Wang et al.

TABLE 5: Calculated Reaction Energies (kcal/mol) and Rate Constants k (cm3/(molecule s)) from TST at 298, 700, and 1500 K for Bimolecular Reactions
reaction TiCl4 + H2O f TS1 f TiCl3OH + HCl (5a) TiCl3OH + H2O f TS3 f TiCl2(OH)2 + HClc (7) TiCl2(OH)2 + H2O f TS4 f TiCl(OH)3 + HCld (8) TiCl(OH)3 + H2O f TS5 f Ti(OH)4 + HCle (9) TiOCl2 + H2O f TiOClOH + HCl (6a)
b a e

H298K H700K H1500K G298K G700K G1500K k(298 K)a 2.7 4.1 5.6 8.7 8.7 3.5 4.8 6.3 9.4 9.3 4.9 6.3 7.7 10.8 10.8 0.7 4.4 6.2 9.4 8.8 -2.2 4.3 6.7 10.0 7.6 -9.7 3.3 6.3 10.1 7.5 2.0 10 5.1 10-4 4.8 10-3 3.5 10-3 f
-3

k(700 K) k(1500 K) 3.0 102 3.4 101 8.4 101 1.3 102 f 5.0 104 4.3 103 6.1 103 1.3 104 f

The tunneling factors are not included in the rate constants. b Qtunnel(298 K) ) 2.38. c Qtunnel(298 K) ) 1.94. d Qtunnel(298 K) ) 1.65. Qtunnel(298 K) ) 1.53. f There is no reaction barrier for the bimolecular reaction other than the reaction endothermicity.

TABLE 6: Calculated Rate Constants from TST (k, s-1) and the RRKM Rate Constant (s-1) at 298, 700, and 1500 K for the Unimolecular Reactions from the Precomplex to the Postcomplexa
reaction
TiCl4 H2O f TS1 f TiCl3OH HCl TiCl3OH H2O f TS3 f TiCl2(OH)2 HClc TiCl2(OH)2 H2O f TS4 f TiCl(OH)3 HCld TiCl(OH)3 H2O f TS5 f Ti(OH)4 HCle TiOCl2 H2O f TS6 f TiOClOH HCl
b

k(298 K)
5.1 10 3.0 10-2 1.9 10-2 3.5 10-1 3.1 10-6
-2

k(700 K)
4.7 10 2.0 106 9.9 105 2.1 107 4.0 104
6

k(1500 K)
8.9 108 6.2 108 4.3 108 1.8 109 1.4 108

reaction TiCl4 H2O f TS1 f TiCl3OH HCl TiCl3OH H2O f TS3 f TiCl2(OH)2 HCl TiCl2(OH)2 H2O f TS4 f TiCl(OH)3 HCl TiCl(OH)3 H2O f TS5 f Ti(OH)4 HCl TiOCl2 H2O f TS5 f TiOClOH HCl

k(RRKM) (T ) 273-373 K, P ) 380-1520 Torr) k(T) k(T) k(T) k(T) k(T) ) ) ) ) ) (3.98 (1.26 (5.01 (1.58 (1.26 1012) 1012) 1011) 1013) 1012) exp(-19.0/RT) exp(-18.6/RT) exp(-18.3/RT) exp(-18.6/RT) exp(-24.2/RT)

k(RRKM) (T ) 200-2000 K, P ) 760-2250 Torr) k(T,P) k(T,P) k(T,P) k(T,P) k(T,P)


b

) ) ) ) )

(5.0 (2.0 (2.0 (2.5 (4.0

1010)P0.36 1010)P0.34 1010)P0.28 1010)P0.45 1010)P0.31

exp(-17.8/RT) exp(-17.5/RT) exp(-17.6/RT) exp(-16.6/RT) exp(-23.4/RT)

a The tunneling factors are not included in the rate constants. Activation energies in kcal/mol. 1.95. d Qtunnel(298 K) ) 1.65. e Qtunnel(298 K) ) 1.53.

Qtunnel(298 K) ) 2.38. c Qtunnel(298 K) )

the TST rate constants assuming a bimolecular reaction (reactions 5a, 7, 8, and 9) at 298, 700, and 1500 K in Table 5 because the precomplexes may not have long lifetimes at the higher temperatures. Both the unimolecular and bimolecular reaction rates are substantial at T g 700 K. TiOCl2 has been observed experimentally in some of the systems,24,25 so we also studied the reaction of TiOCl2 with water (reaction 6). The rst step as shown in the following reaction begins with the formation of an initial Lewis acid-base complex with a much larger complexation energy of 37.8 kcal/mol (Figure 2):

TiOCl2 + H2O f TiClOH + HCl

(6a)

The hydrogen transfer reaction from the complex has a higher barrier of 24.2 kcal/mol, and the resulting complex is a Lewis acid-base adduct of HCl with TiO(OH)Cl with a Lewis acid-base adduct bond energy of 16.1 kcal/mol. The overall reaction to form HCl + TiO(OH)Cl is endothermic by 9.2 kcal/ mol at 0 K. The value of Keq at 700 K for reaction 6a is 4 10-3, which is large enough for some product to be formed which can be involved in further clustering reactions. There is no reaction barrier on the bimolecular energy surface other than the endothermicity of the reaction as the transition state for hydrogen transfer is below the energy of the reactants. At higher temperatures, the reaction will occur at the collision frequency reduced by the appropriate exponential expression involving the endothermicity (a factor of 10-20 at 1500 K). The loss of the second HCl is highly endothermic as expected. Titanium Oxychloride Hydroxide Species. We have noted above that the formation of TiO2 nanoclusters must proceed by way of TixOyClz (or TixOyClz(OH)w) clusters as discussed in the Introduction and have shown that it is possible to generate the intermediates needed for clustering reactions at the experimental

temperatures. Figure 3 shows the calculated molecular structures with key geometry parameters for Ti2O3Cl2, Ti2O2Cl4, Ti2OCl6, Ti3O5Cl2, Ti3O4Cl4, Ti3O3Cl6, Ti4O6Cl4 and for a number of TixOyClz(OH)w clusters. These molecules were optimized at the DFT level with the B3LYP, BP86,56,57 PBE,58,59 and PW9160,61 exchange-correlation functionals and the aVDZ basis set. The geometries obtained with the aVDZ basis set were then used in single-point DFT energy calculations with the aVTZ basis set. The geometries obtained at the B3LYP/aVDZ level were also used in single-point energy calculations at the CCSD(T)/aVDZ level. In our previous work on TiO2 clusters,19 we studied the clustering energies of (TiO2)n (n ) 2-4) clusters, and we found that the normalized clustering energies at 0 K range from 60 to 100 kcal/mol, increasing from the dimer to the tetramer. The normalized clustering energy is the average binding energy of the monomers in a cluster and for a (TiO2)n cluster with n Ti atoms, Enorm,n is dened by the following equation:

Enorm,n ) [nE(monomer) - E(cluster)]/n

(10)

For the TiO2 clusters previously studied, the CBS extrapolation effect from the aVDZ basis set is small in contrast to the basis set extrapolation in the total atomization energies. We can then use the following equation to obtain the heat of formation of a cluster where we are summing over the heats of formation of the monomers:

Hf,0K(cluster) ) Hf,0K(monomer) - nEnorm,n

(11)
The advantage of this approach is that the normalized clustering energies (Enorm,n) are not strongly dependent on the basis set.19

Hydrolysis of TiCl4

J. Phys. Chem. A, Vol. 114, No. 28, 2010 7567 predicted the same trends observed in our study. The calculated Ti-Cl bond distances in Ti3O5Cl2 are longer than those in the clusters with three and four Ti atoms. The Ti-O bond distance in the trimers and tetramer are longer than the Ti-O bond distance in TiO2. The Ti-O bond distance in the trimers and the tetramer is similar to that of the dimers. Our calculations showed no signicant dependence in the calculated geometries on the choice of the DFT exchange-correlation functional. The calculated clustering energies and heats of formation are given in Table 7. The calculated normalized clustering energy of Ti2O2Cl4 at the CCSD(T)/aD level is less exothermic than that of Ti3O3Cl6 by 8.8 kcal/mol. The calculated heat of formation at 298 K of Ti2O2Cl4, -370.6 kcal/mol, is in excellent agreement with the B97-1/6-311G(d,p) value26 of -370.9 kcal/ mol. Our calculated heat of formation at 298 K of Ti2O3Cl2, -316.0 kcal/mol, is similar to the previously calculated value26 of -318.1 kcal/mol. For Ti3O4Cl4, our calculated value is lower than that of Green and co-workers26 by 3.3 kcal/mol. A second set of dimers that involve OH as a substituent as well can be generated as shown in Figure 3 and Table 7. Substitution of a OH for a Cl does not lead to much geometry change in the core involving the Ti atoms. The Ti-O bridge bond distance increases by up to 0.03 . In most cases, the Ti-O(H) bond distances are comparable to or slightly shorter than a Ti-O(Ti) bond distance. Cluster Reaction Energies. The clustering energies given in Table 7 are in general highly exothermic,so condesation reactions of species such as Ti(O)Cl2 and Ti(O)Cl(OH) will readily occur at the temperatures of interest. This leads to important oxygenated clusters that can serve as precursors to TiO2 nanoparticles. We can use the calculated heats of formation to explore a range of additional condensation reactions coupled with elimination of H2O or HCl that may be relevant to the formation of TiO2 nanoparticles as demonstrated in Table 8. The reaction of TiCl3OH with itself or with TiCl4 forms the dimer Ti2OCl6 in an exothermic reaction and releases either H2O or HCl. The dimer Ti2O2Cl4 can be formed either by reaction of TiCl3OH with itself or by the reaction of TiCl2(OH)2 with TiCl4. Both reactions are slightly endothermic on the enthalpy scale but will be nearly thermoneutral or exothermic on the free energy scale depending on the temperature due to the formation of an additional molecule of HCl. The reaction of TiCl2(OH)2 with TiCl4 can also form a single bridged cluster with elimination of HCl, and this reaction is slightly exothermic. The reactions of TiCl(OH)3 are mostly exothermic if H2O is eliminated but are somewhat endothermic if two HCl molecules are eliminated. The reactions of Ti(OH)4 are exothermic, and the reaction of Ti(OH)4 with Ti(O)(OH)2 is highly exothermic. As expected, reactions that eliminate more than two HCl molecules to enhance oxygen atom incorporation in the dimer are substantially endothermic. Basis Set Dependence and DFT Performance. The dependence of the relative energies on the basis set is given in the Supporting Information. There is only a small basis set dependence at the CCSD(T) level, so a reasonable potential energy surface can be obtained with a modest basis set. Figures 1 and 2 compare the Etotal results and ECBS + ECV results with the B3LYP/aVDZ results. The complexation energies of H2O to a Ti calculated at the B3LYP/aVDZ level are in general too small as compared to those calculated at the CCSD(T) level. This arises because of the lack of formal dispersion energy treatments in the DFT functionals used in this study and the difculty that DFT can have in predicting

Figure 3. Optimized molecular structures for Ti2O3Cl2, Ti2O2Cl4, Ti2OCl6, Ti3O4Cl4, Ti3O5Cl2, Ti3O3Cl6, Ti4O6Cl4, TiO(OH)2, cisTi2O2Cl2(OH)2, trans-Ti2O2Cl2(OH)2, Ti2O2(OH)4, Ti2OCl5OH, Ti2OCl4(OH)2, Ti2OCl2(OH)4, Ti2O(OH)6, Ti2O2Cl(OH)3, and isoTi2O2Cl2(OH)2: yellow, Ti; red, O; green, Cl; white, H.

Thus, the heat of formation of the cluster can be calculated from the CCSD(T)/CBS heats of formation of the monomers and the CCSD(T)/aVDZ normalized clustering energies of the cluster, as given in eq 10. Ti2O3Cl2, Ti2O2Cl4, and Ti2OCl6 are clusters with two Ti atoms that can be potentially formed in the oxidation of TiCl4 and involve only Cl and O as substituents. The calculated Ti-Cl bond distance in Ti2O3Cl2 is longer than that of Ti2O2Cl4 and Ti2OCl6. Most of the calculated Ti-O bond distances in Ti2O3Cl2, Ti2O2Cl4, and Ti2OCl6 are longer than that of TiO2, and the Ti-O bond distance in these dimers increases as the number of oxygens increases. Previous work26 on the structures of Ti2O3Cl2 and Ti2O2Cl4 at the B97-1/6-311+G(d,p) level

7568

J. Phys. Chem. A, Vol. 114, No. 28, 2010

Wang et al.

TABLE 7: Calculated Heats of Formation (kcal/mol) of the Clusters and Reaction Energies (kcal/mol) at the CCSD(T)/aVDZ// B3LYP/aVDZ Level
reaction 2TiOCl2 f Ti2O2Cl4 3TiOCl2 f Ti3O3Cl6 TiOCl2 + TiO2 f Ti2O3Cl2 TiOCl2 + 2TiO2 f Ti3O5Cl2 2TiOCl2 + TiO2 f Ti3O4Cl4 TiOCl2 + TiCl4 f Ti2OCl6 2TiOCl2 + 2TiO2 f Ti4O6Cl4 TiCl4 + TiOClOH f Ti2OCl5OH 2TiOClOH f Ti2O2Cl2(OH)2 (cis) 2TiOClOH f Ti2O2Cl2(OH)2 (trans) 2TiOClOH f Ti2O2Cl2(OH)2 (iso) Ti(OH)4 + TiO(OH)2 f Ti2O(OH)6 2TiO(OH)2 f Ti2O2(OH)4 TiOClOH + TiO(OH)2 f Ti2O2Cl(OH)3 H0K -86.3 -155.9 -105.7 -203.7 -200.9 -54.9 -296.9 -58.6 -92.2 -93.2 -91.1 -55.1 -88.8 -91.5 H298K -87.0 -156.4 -106.4 -204.2 -201.9 -55.2 -298.4 -59.1 -92.5 -93.5 -92.4 -54.4 -88.7 -91.7 H700K -85.7 -153.5 -105.1 -201.4 -199.2 -54.5 -294.1 -57.0 -90.9 -92.1 -90.9 -52.6 -86.9 -93.1 H1500K -82.5 -147.1 -102.0 -195.1 -192.9 -52.9 -284.5 -53.8 -87.7 -88.9 -87.8 -49.3 -83.5 -120.2 Hf,0K -361.9 -580.1 -314.3 -479.5 -550.9 -377.3 -714.1 -407.8 -427.6 -428.6 -426.5 -552.7 -483.6 -456.6 Hf,298K -370.6 -581.8 -316.0 -481.6 -553.3 -378.5 -717.6 -409.4 -430.1 -431.1 -430.0 -557.8 -487.8 -460.0

TABLE 8: Calculated Reaction Energies (kcal/mol)


reaction TiCl3OH + TiCl4 f Ti2OCl6 + HCl 2TiCl3OH f Ti2OCl6 + H2O 2TiCl3OH + TiCl4 f Ti2O2Cl4 + 2HCl TiCl2(OH)2 + TiCl4 f Ti2O2Cl4 + 2HCl TiCl2(OH)2 + TiCl4 f Ti2OCl5OH + HCl 2TiCl2(OH)2 f Ti2O4 + 4HCl 2TiCl2(OH)2 f Ti2O2Cl4 + 2H2O TiCl3OH + TiCl2(OH)2 f Ti2 O3Cl2 + 3HCl 2TiCl(OH)3 f cis-Ti2O2Cl2(OH)2 + 2H2O 2TiCl(OH)3 f trans-Ti2O2Cl2(OH)2 + 2H2O 2TiCl(OH)3 f iso-Ti2O2Cl2(OH)2 + 2H2O 2TiCl(OH)3 f Ti2O2(OH)4 + 2HCl 2TiCl(OH)3 f Ti2OCl2(OH)4 + H2O TiCl(OH)3 + TiCl4 f Ti2OCl4(OH)2 + HCl 2Ti(OH)4 f Ti2O(OH)6 + H2O 2Ti(OH)4 f Ti2O2(OH)4 + 2H2O Ti(OH)4 + TiO(OH)2 f Ti2O2(OH)4 + H2O H0K -4.7 -7.3 14.0 12.4 -4.7 145.1 5.6 77.7 6.5 5.0 6.2 19.9 107.5 49.1 -10.3 2.3 -43.3 H298K -4.6 -7.3 14.3 12.9 -3.9 146.3 6.0 78.4 6.7 5.7 6.8 20.5 108.0 50.2 -9.2 3.0 -42.9 H700K -4.9 -6.2 13.5 12.1 -3.4 143.4 3.8 76.6 3.5 3.6 4.7 19.8 107.1 50.8 -9.3 1.0 -43.0 H1500K -5.4 -4.3 12.4 11.0 -2.3 138.1 -0.1 73.4 -4.1 -0.2 0.8 18.8 105.2 51.9 -9.6 -2.8 -43.2

dative bond energies.45,62 These interactions are not dominated by hydrogen bonding. The DFT functionals used may underestimate the reaction barriers as found in some of the hydrogen transfer reactions. The relative energies of TiCl4 H2O, TiCl3OH H2O, TiCl2(OH)2 H2O, TiCl3OH H2O, and TiOCl2 H2O calculated at the B3LYP/aVDZ level are greater than those at the CCSD(T) level for ECBS + ECV by 2.8, 5.8, 4.8, 4.9, and 3.1 kcal/mol, respectively. Thus, DFT/ B3LYP provides a qualitatively correct potential energy surface but not a quantitative one. Conclusions Coupled cluster [CCSD(T)] theory, Mller-Plesset perturbation theory (MP2), and density functional theory (DFT) have been used to study the structural and energetic properties of the intermediates in the hydrolysis of TiCl4 and the formation of titanium oxychloride species (TixOyClz; x ) 2-4, y ) 1, 3-6, and z ) 2, 4, 6). The normalized clustering energies at 0 K for the ground state of (TiOCl2)n (n ) 2, 3) clusters range from 40 to 55 kcal/mol, increasing from the dimer to the trimer. The reaction energy of the condensation reactions is substantially endothermic if more than two HCl molecules are eliminated. In addition, we also predict the reaction paths for the initial steps of hydrolysis of TiCl4. We predict that the hydrolysis of TiCl4 will not form TiOCl2 and HCl or TiO2 and HCl due to the substantial endothermicities associated with the formation of gas-phase TiO2. Using transition-state theory and RRKM theory, we predicted rate constants including the effect of tunneling. The low-temper-

ature rate constant is independent of pressure, and the effect of tunneling on the rate constant is an increase of about a factor of 2 at 298 K. At higher temperatures, the predicted rate constants are large enough for the reactions to occur readily and show that key oxygenated intermediates can be formed. The calculations show, at the temperatures relevant to the experimental conditions for the gas-phase hydrolysis of TiCl4 or the combustion of TiCl4 leading to TiO2 particles, that partially oxygenated Ti molecules can be generated. These molecules can condense to form partially oxygenated Ti clusters which are precursors to TiO2 nanoparticles. Our calculated data will enable the development of more detailed gas-phase mechanisms for the formation of TiO2 nanoparticles and larger particles which incorporate the hydrolysis of TiCl4. Within this framework, our calculations also show that the reported values for the heat of formation of TiOCl2 should be remeasured. Acknowledgment. This work was supported by the National Science Foundation (Grant CTS-0608896), through the NIRT program, and by the Chemical Sciences, Geosciences and Biosciences Division, Ofce of Basic Energy Sciences, U.S. Department of Energy (DOE), under Grant DE-FG0203ER15481 (Catalysis Center Program). D.A.D. also thanks the Robert Ramsay Chair Fund of The University of Alabama for support. Part of this work was performed in the Molecular Sciences Computing Facility at the W. R. Wiley Environmental Molecular Sciences Laboratory, a national scientic user facility sponsored by the DOEs Ofce of Biological

Hydrolysis of TiCl4 and Environmental Research and located at Pacic Northwest National Laboratory, operated for the DOE by Battelle. We thank Prof. K. A. Peterson of Washington State University for providing the Ti basis set. Supporting Information Available: Calculated geometry parameters for the molecules, CCSD(T)/aVnZ total energies (Eh) as a function of the basis set, calculated harmonic frequencies for molecules at the MP2/aVTZ level, calculated harmonic frequencies for Ti2O3Cl2, Ti2O2Cl4, Ti2OCl6, Ti3O4Cl4, Ti3O5Cl2, Ti3O3Cl6, and Ti4O6Cl4 at the B3LYP/ aVDZ, BP86/aVDZ, PBE/aVDZ, and PW91/aVDZ levels, T1 diagnostics calculated at the CCSD(T)/aVQZ level, contributions to the relative energies on the different potential energy surfaces, CCSD(T) relative energies for the potential energy surfaces as a function of the basis set, x, y, and z coordinates of molecules at the MP2/aVTZ level (), complete potential energy surface for the hydrolysis of TiCl4 to produce TiO2 with different calculation methods (kcal/mol), and plots of the rate constant (k) vs temperature. This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes
(1) Greenwood, N. N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Pergamon Press: Oxford, U.K., 1984. (2) Cotton, F. A.; Wilkinson, G.; Murillo, C. A.; Bochmann, M. AdVanced Inorganic Chemistry, 6th ed.; Wiley: New York, 1999. (3) Thompson, T. L.; Yates, J. T. Chem. ReV. 2006, 106, 4428. (4) Anpo, M.; Dohshi, S.; Kitano, M.; Hu, Y. In Metal Oxides: Chemistry and Applications; Fierro, J. L. G., Ed.; CRC Press: Boca Raton, FL, 2006; pp 595-622. (5) Gratzel, M. Nature 2001, 414, 338. (6) Fujishima, A.; Hashimoto, K.; Watanabe, H. TiO2 Photocatalysis: Fundamentals and Applications; BKC, Inc.: Tokyo, 1997. (7) Hattori, A.; Yamamoto, M.; Tada, H.; Ito, S. Chem. Phys. Lett. 1998, 8, 707. (8) Sopyan, I.; Murasawa, S.; Hasimoto, K.; Fujishima, A. Chem. Phys. Lett. 1998, 10, 723. (9) Dayte, A. K.; Riegel, G.; Bolton, J. R.; Huang, M.; Prairie, M. R. J. Solid State Chem. 1995, 115, 236. (10) Cervais, C.; Simth, M. E.; Pottier, A.; Jolivet, J. P.; Babonneau, F. Chem. Mater. 2001, 13, 462. (11) Gonzalez, R. A.; Musick, C. D.; Tilton, J. N. Process for controlling agglomeration in the manufacture of TiO2. U.S. Patent 5,508,015, April 16, 1996. (12) Deberry, J. C.; Robinson, M.; Pomponi, M. D.; Beach, A. J.; Xiong, Y.; Akhtar, K. Controlled vapor phase oxidation of titanium tetrachloride to manufacture titanium dioxide. U.S. Patent 6,387,347, May 14, 2002. (13) Akhtar, M. K.; Pratsinis, S. E.; Mastrangelo, S. V. R. J. Mater. Res. 1994, 9, 1241. (14) Shirley, R.; Liu, Y.; Totton, T. S.; West, R. H.; Kraft, M. J. Phys. Chem. A 2009, 113, 13790. (15) Addamo, M.; Augugliaro, V.; Di Paola, A.; Garca-Lopez, E.; Loddo, V.; Marc, G.; Palmisano, L. Colloids Surf., A 2005, 265, 23. (16) Xia, B.; Huang, H; Xie, Y. Mater. Sci. Eng., B 1999, 57, 150. (17) ORegan, B.; Moser, J.; Anderson, M.; Gratzel, M. J. Phys. Chem. 1990, 94, 8720. Liu, D.; Kamat, P. J. Phys. Chem. 1993, 97, 10769. (18) Stathatos, E.; Lianos, P.; del Monte, F.; Levy, D.; Tsiourvas, D. J. Sol-Gel Sci. Technol. 1997, 10, 83. (19) Li, S.; Dixon, D. A. J. Phys. Chem. A 2008, 112, 6646. (20) Li, S.; Hennigan, J. M.; Dixon, D. A. J. Phys. Chem. A 2009, 113, 7861. (21) Chase, M. W. NIST-JANAF Themochemical Tables, 4th ed.; J. Phys. Chem. Ref. Data, Monogr. 1998, No. 9, 1-1951. (22) Wagman, D. D.; Evans, W. H.; Parker, V. B.; Schumm, R. H.; Halow, I.; Bailey, S. M.; Churney, K. L.; Nuttall, R. L. J. Phys. Chem. Ref. Data 1982, 11, Suppl. 2. (23) Pratsinis, S. E.; Bai, H.; Biswas, P.; Frenklach, M.; Mastrangelo, S. V. R. J. Am. Ceram. Soc. 1990, 73, 2158. (24) Karlemo, B.; Koukkari, P.; Paloniemi, J. Plasma Chem. Plasma Process. 1996, 16, 59. (25) Hildenbrand, D. L.; Lau, K. H.; Mastrangelo, S. V. R. J. Phys. Chem. 1991, 95, 3435. (26) West, R. H.; Beran, G. J. O.; Green, W. H.; Kraft, M. J. Phys. Chem. A 2007, 111, 3560.

J. Phys. Chem. A, Vol. 114, No. 28, 2010 7569


(27) Parr, R. G.; Yang, W. Density-Functional Theory of Atoms and Molecules; Oxford University Press: New York, 1989. (28) Bartlett, R. J.; Musial, M. ReV. Mod. Phys. 2007, 79, 291. (29) (a) Purvis, G. D., III; Bartlett, R. J. J. Chem. Phys. 1982, 76, 1910. (b) Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. Chem. Phys. Lett. 1989, 157, 479. (c) Watts, J. D.; Gauss, J.; Bartlett, R. J. J. Chem. Phys. 1993, 98, 8718. (30) West, R. H.; Celnik, M. S.; Inderwildi, R.; Kraft, M.; Beran, G. J. O.; Green, W. H. Ind. Eng. Chem. 2007, 46, 6147. (31) West, R. H.; Shirley, R. A.; Kraft, M.; Goldsmith, C. F.; Green, W. H. Combust. Flame 2009, 156, 1764. (32) (a) Feller, D.; Dixon, D. A. J. Phys. Chem. A 2000, 104, 3048. (b) Feller, D.; Dixon, D. A. J. Chem. Phys. 2001, 115, 3484. (c) Dixon, D. A.; Feller, D.; Peterson, K. A. J. Chem. Phys. 2001, 115, 2576. (d) Ruscic, B.; Wagner, A. F.; Harding, L. B.; Asher, R. L.; Feller, D.; Dixon, D. A.; Peterson, K. A.; Song, Y.; Qian, X.; Ng, C.; Liu, J.; Chen, W.; Schwenke, D. W. J. Phys. Chem. A 2002, 106, 2727. (e) Feller, D.; Dixon, D. A. J. Phys. Chem. A 2003, 107, 9641. (f) Pollack, L.; Windus, T. L.; de Jong, W. A.; Dixon, D. A. J. Phys. Chem. A 2005, 109, 6934. (g) Feller, D.; Peterson, K. A.; Dixon, D. A. J. Chem. Phys. 2008, 129, 204015. (33) Lee, C.; Yang, W.; Parr, R. G. Phys. ReV. B 1988, 37, 785. (34) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (35) Frisch, M. J. Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, revision E.01; Gaussian, Inc.: Wallingford, CT, 2004. (36) Kendall, R. A.; Dunning, T. H., Jr.; Harrison, R. J. J. Chem. Phys. 1992, 96, 6796. (37) Dunning, T. H., Jr.; Peterson, K. A.; Wilson, A. K. J. Chem. Phys. 2001, 114, 9244. (38) Peterson, K. A.; Figgen, D.; Dolg, M; Stoll, H. J. Chem. Phys. 2007, 126, 124101. Balabanov, N.; Peterson, K. A. J. Chem. Phys. 2005, 123, 064107. (39) Amos, R. D.; Bernhardsson, A.; Berning, A.; Celani, P.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.; Hampel, C.; Hetzer, G.; Knowles, P. J.; Korona, T.; Lindh, R.; Lloyd, A. W.; McNicholas, S. J.; Manby, F. R.; Meyer, W.; Mura, M. E.; Nicklass, A.; Palmieri, P.; Pitzer, R.; Rauhut, G., Schutz, M.; Schumann, U.; Stoll, H., Stone, A. J.; Tarroni, R.; Thorsteinsson, T.; Werner, H.-J. MOLPRO, a package of ab initio programs designed by H.-J. Werner and P. J. Knowles, version 2006.1, Universitat Stuttgart, Stuttgart, Germany, and University of Birmingham, Birmingham, U.K. (40) Peterson, K. A.; Woon, D. E.; Dunning, T. H., Jr. J. Chem. Phys. 1994, 100, 7410. (41) Moore, C. E. Atomic Energy LeVels As DeriVed from the Analysis of Optical Spectra, Vol. 1, H to V; U.S. National Bureau of Standards Circular 467; National Technical Information Service, COM-72-50282; U.S. Department of Commerce: Washington, DC, 1949. (42) Cox, J. D.; Wagman, D. D.; Medvedev, V. A. CODATA Key Values for Thermodynamics; Hemisphere Publishing Corp.: New York, 1989. The heat of formation for Ti is given in this reference at 298 K. The heat of formation of Ti at 0 K is obtained from this value with the correction from 0 to 298 K given in ref 21. (43) Curtiss, L. A.; Raghavacchari, K.; Redfern, P. C.; Pople, J. A. J. Chem. Phys. 1997, 106, 1063. (44) McQuarrie, D. A. Statistical Mechanics; University Science Books: Sausalito, CA, 2000. (45) (a) Dixon, D. A.; Gutowski, M. J. Phys. Chem. A 2005, 109, 5129. (b) Matus, M. H.; Anderson, K. A.; Camaioni, D. M.; Autrey, S. T.; Dixon, D. A. J. Phys. Chem. A 2007, 111, 4411. (46) Li, S.; Dixon, D. A. J. Phys. Chem. A 2010, 114, 2665. (47) Hildenbrand, D. L. J. Phys. Chem. A 2009, 113, 1472; High Temp. Mater. Sci. 1996, 35, 151. (48) Mrino, Y.; Uehara, U. J. Chem. Phys. 1966, 45, 4543. (49) Webb, S. P.; Gordon, M. S. J. Am. Chem. Soc. 1999, 121, 2552. (50) Brunken, S.; Muller, H. S. P.; Menten, K. M.; McCarthy, M. C.; Thaddeus, P. Astrophys. J. 2008, 676, 1367. (51) Lee, T. J.; Taylor, P. R. Int. J. Quantum Chem. Symp. 1989, 23, 199.

7570

J. Phys. Chem. A, Vol. 114, No. 28, 2010

Wang et al.
(59) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. ReV. Lett. 1997, 78, 1396. (60) Burke, K.; Perdew, J. P. Wang, Y. In Electronic Density Functional Theory: Recent Progress and New Directions; Dobson, J. F., Vignale, G., Das, M. P., Eds.; Plenum: New York, 1998. (61) Perdew, J. P.; Wang, Y. Phys. ReV. B 1991, 45, 13244. (62) Gilbert, T. M. J. Phys. Chem. A 2004, 108, 2550.

(52) Steinfeld, J. I.; Francisco, J. S.; Hase, W. L. Chemical Kinetics and Dynamics, 2nd ed.; Prentice Hall: Englewood Cliffs, NJ, 1999. (53) Holbrook, K. A.; Pilling, M. J.; Robertson, S. H. Unimolecular Reaction, 2nd ed.; Wiley: Chichester, U.K., 1996. (54) Wigner, E. Z. Z. Phys. Chem. B 1932, 19, 203. (55) Skodje, R. T.; Truhlar, D. J. J. Chem. Phys. 1981, 85, 624. (56) Becke, A. D. Phys. ReV. A 1988, 38, 3098. (57) Perdew, J. P. Phys. ReV. B 1986, 33, 8822. (58) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. ReV. Lett. 1996, 77, 3865.

JP102020H

S-ar putea să vă placă și