2+ and

May 23, 2011 - European Commission, Joint Research Centre, Institute for Transuranium ... The joint analysis of literature experimental and our comput...
1 downloads 0 Views 1MB Size
ARTICLE pubs.acs.org/JPCA

Computed Vibrational Frequencies of Actinide Oxides AnO0/þ/2þ and AnO20/þ/2þ (An = Th, Pa, U, Np, Pu, Am, Cm) Attila Kovacs*,†,‡ and Rudy J.M. Konings† † ‡

European Commission, Joint Research Centre, Institute for Transuranium Elements, P.O. Box 2340, 76125 Karlsruhe, Germany Materials Structure and Modeling Research Group of the Hungarian Academy of Sciences, Budapest University of Technology and Economics, H-1111 Budapest, Szt. Gellert ter 4

bS Supporting Information ABSTRACT: The vibrational frequencies of the actinide oxides AnO and AnO2 (An = Th, Pa, U, Np, Pu, Am, Cm) and of their mono- and dications have been calculated using advanced quantum chemical techniques. The stretching fundamental frequencies of the monoxides have been determined by fitting the potential function to single-point energies obtained by relativistic CASPT2 calculations along the stretching coordinate and on this basis solving numerically the ro-vibrational Schr€odinger equation. To obtain reliable fundamental frequencies of the dioxides, we developed an empirical approach. In this approach the harmonic vibrational frequencies of the AnO20/þ/2þ species were calculated using eight different exchange-correlation DFT functionals. On the basis of the good correlation found between the vibrational frequencies and computed bond distances, the final frequency values were derived for the CASPT2 reference bond distances from linear regression equations fitted to the DFT data of each species. As a test, the approach provided excellent agreement with accurate experimental data of ThO, ThOþ, UO, and UOþ. The joint analysis of literature experimental and our computed data facilitated the prediction of reliable gas-phase molecular properties for some oxides. They include the stretching frequencies of PuO, ThO2, UO2, and UO2þ and the bond distance of PuO (1.818 Å, being likely within 0.002 Å of the real value). Also the derived equilibrium bond distances of ThO2, UO2, and UO2þ (1.896, 1.790, and 1.758 Å, respectively) should approximate closely the (yet unknown) experimental values. On the basis of the present results, we suggest that the ground electronic state of PuO2 in Ar and Kr matrices is probably different from that in the gaseous phase, similarly to UO2 observed previously.

1. INTRODUCTION Actinide oxides are of high relevance in nuclear technology: the uranium and plutonium dioxides are important nuclear fuels, while the other (minor) actinides are formed in the fuel during irradiation as a result from neutron capture processes. Because of the long-term perspective of nuclear energy, considerable activities are in progress for the development of safer and more economic nuclear reactors. Several of the promising Generation IV reactor types are fast reactors in which the fuel operates at extreme conditions. The temperature in the middle zone of the fuel can reach 2500 K,13 where the actinide oxides have already a substantial vapor pressure. Hence, for a proper thermodynamic description of the high-temperature fuel restructuring processes in these nuclear fuels, the vapor species have to be taken into account, requiring the knowledge of molecular data of the actinide oxides. Due to the complications handling actinide, particularly other than Th and U, material experimental information is available only for a few actinide oxides. Previous experimental investigations of gaseous oxides include the microwave spectroscopic investigations of ThO,46 the high-resolution photoelectron r 2011 American Chemical Society

spectroscopic studies of ThO/ThOþ,7 UO,8 UOþ,9 UO2,10 and UO2þ,11 and the matrix-isolation IR spectroscopic measurements of ThO/ThO2,1215 UO/UO2,1618 UO2þ,1719 and PuO/PuO220 in various matrices. The microwave and highresolution photoelectron spectroscopic studies provided accurate values for the bond lengths, harmonic vibrational frequencies, and first anharmonicities of the studied species, while the matrix-isolation IR measurements provided anharmonic frequencies of the IR active stretching fundamentals. Note that the latter data suffer from matrix shifts and occurrently from perturbation effects (vide infra), requiring caution at their interpretation. For the rest of the actinide oxides there is no direct experimental information available, only estimates based on the properties of the above actinide or related lanthanide compounds.2125 Attempts to fill the gaps in the data of gaseous actinide oxides came from the field of quantum chemical calculations. Our Received: March 17, 2011 Revised: May 5, 2011 Published: May 23, 2011 6646

dx.doi.org/10.1021/jp202538k | J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A literature search found computed vibrational data for ThO,7,15,2633 ThOþ,7 UO,18,3437 UOþ,35,37 PuO,3840 PuOþ,39,41,42 PuO2þ,39,41 AmO,43 AmOþ,43 CmO,43 ThO2,13,15,44 PaO2þ,44 UO2,18,4549 UO2þ,18,19,47,50 UO22þ,18,44,47,4952 NpO2,49 NpO22þ,49 PuO2,39,40,49,53,54 PuO2þ,39 and PuO22þ,39,49,51,55,56 obtained at various levels of theory. The computed bond distances and vibrational frequencies depend strongly on the computational level (and on the sometimes different electronic state) and in most cases differ considerably from the available experimental results. Very recently we performed a systematic theoretical study for all the title species at the relativistic spinorbit CASPT2 (SO-CASPT2) level and reported the electronic ground states, reliable molecular geometries, and ionization energies.57 The goal of our present work is the evaluation of reliable vibrational frequencies for AnO and AnO2 (An = Th, Pa, U, Np, Pu, Am, Cm) molecules as well as for their mono- and dications by means of sophisticated quantum chemical calculations. The work is based on the electronic ground states reported recently.57 The vibrational data of the monoxides are evaluated by analysis of the potential energy curve along the stretching coordinate. We paid special attention to the accuracy of the curve, and therefore we used 2030 sampling points and a very narrow sampling rate around the minimum. In a few cases, we found a strong dependence of the vibrational frequency and anharmonicity values on the sampling parameters. Due to difficulties to obtain vibrational frequencies of triatomic species at the CASPT2 level, we chose for the dioxide species a different approach. The recently found nearly singledeterminant character of the ground-state electronic structures of the title species57 suggested a promising application of the DFT technique. However, the DFT-computed geometries deviate more or less from the reference CASPT2 data, which leads consequently to errors of the DFT vibrational frequencies. Moreover, the deviations of DFT data are not consistent in the actinide row. Therefore, instead of selecting a single “best performing” DFT method we computed each species with eight different exchange-correlation functionals, and on the basis of the found empirical relationships between the optimized bond distances and vibrational frequencies, we evaluated the frequencies for SO-CASPT2 bond distances available from the literature. The reliability of our evaluated molecular parameters is assessed by comparison with available experimental data of the thorium, uranium, and plutonium oxides.420

2. COMPUTATIONAL DETAILS The calculations on the monoxides were performed using the code MOLCAS 7.4 patch 045.5860 The complete active space (CAS) SCF method61 was used to generate molecular orbitals and reference functions for subsequent multiconfigurational second-order perturbation theory calculations of the dynamic correlation energy (CASPT2).62,63 The active space consisted of the 7s, 6d, and 5f orbitals of the actinide atoms as well as of the unoccupied 2p orbitals of oxygen, altogether 16 orbitals occupied by the appropriate number of valence electrons in the various species. All-electron basis sets of atomic natural orbital type developed for relativistic calculations (ANO-RCC) with the Douglas KrollHess Hamiltonian64,65 were used for all the atoms. For the actinides a primitive set of 26s23p17d13f5g3h was contracted to 9s8p6d5f2g1h,66 whereas for O a primitive set of 14s9p4d3f2g

ARTICLE

was contracted to 4s3p2d1f67 achieving VTZP quality. The DouglasKrollHess Hamiltonian was used in the CASSCF calculations to account for scalar relativistic effects. In the second-order perturbation treatment (CASPT2), the orbitals up to 5d of the actinides and 1s of oxygen were kept frozen, while the remaining valence and semivalence orbitals (including 6s and 6p of the actinides and 2s of oxygen) were correlated. These CASPT2 calculations have been performed on the recently reported electronic ground states of the monoxide species.57 The potential energy curves were obtained by singlepoint calculations using steps of 0.1 Å between 1.5 and 2.3 Å. The sampling distance was decreased to 0.02 Å in a range of 0.2 Å around the equilibrium bond distance. In a few cases we obtained unrealistic values for the anharmonicity ωexe which could be corrected by further refinement of the sampling to steps of 0.002 Å in the 0.02 Å range around the equilibrium bond distance. (This is the reason why we did not consider spinorbit effects: the calculation of such numerous data points using the complete active space interaction (CASSI) method68 would require enormous computational costs.) to avoid the change of electronic state along the potential curve, we used the SUPSym keyword of MOLCAS, allowing orbital rotations in the active space between the defined 16 orbitals only. The equilibrium bond distances, the harmonic vibrational frequencies, and other spectroscopic constants of the monoxides were determined by using the program VIBROT available in the MOLCAS package. In the VIBROT run the potential is fitted to an analytical form using cubic splines, and then the ro-vibrational Schr€odinger equation is solved numerically.59 The frequencies of the dioxide species have been calculated by the code Gaussian 0369 using the following exchange-correlation functionals: B3LYP,70,71 B3P86,70,72 mPW1PW91,73 TPSSTPSS,74 PBE1PBE (alias PBE0),75 BLYP,71,76 BP86,72,76 PBEPBE,77,78 and HCTH.79 A linear equation was fitted to the optimized bond distances and computed harmonic vibrational frequencies (one model for all the computed data of a fundamental), and using this equation we evaluated the frequencies for the SOCASPT2 bond distances. This procedure was validated on the experimental data of ThO, ThOþ, UO, and UOþ. Note that the vibrational frequencies obtained by the HCTH functional have been omitted from our analysis because of their considerable deviation from the frequencybond distance relationships shown by the other results. In the DFT calculations, the small-core relativistic effective core potentials of the CologneStuttgart group (ECP60MWB) were used for the actinides28,80 in conjunction with a 14s13p10d8f6g valence basis set contracted to 10s9p5d4f3g (ECP60MWB_SEG basis).80 For oxygen, the correlation-consistent augcc-pVTZ basis set81 was used. Particular care was taken on the electronic states in the DFT calculations. After the SCF cycle reached convergence, orbital population analyses have been performed for each species, and the results were compared to the ground-state electronic configurations known from the literature.57 In the few cases when the SCF converged to a different (generally higher-lying) electronic state, the DFT orbital populations were altered to the reference (literature) ground states.

3. RESULTS AND DISCUSSION 3.1. Vibrational Frequencies of the Monoxides. The computed spin-free CASPT2 bond distances and selected computed 6647

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Table 1. Equilibrium Bond Distances, Harmonic Vibrational Frequencies, and Anharmonicitiesa of the Monoxide Species Computed at the Spin-Free CASPT2 Level species

ωe

ωexe

Σ

1.861

878.9

2.3

Σ

1.827

930.8

2.6

Σ

1.792

987.6

2.7

Φ

ThO ThOþ

2

ThO2þ

1

PaO

2

1.811

926.7

2.1

PaOþ PaO2þ

3

H 2 Φ

1.805 1.752

932.3 1040.0

2.6 2.6

5

1.837

857.9

1.2

UO

4

1.799

910.9

2.4

UO2þ

3

H

1.728

1047.2

2.8

NpO

6

Δ

1.837

899.8

3.0

NpOþ

5

Γ

1.797

966.8

2.7

NpO2þ

4

I

1.722

1082.0

3.4

PuO PuOþ

7

Π 6 Π

1.818 1.777

858.2 918.5

2.8 2.2

Γ

1.720

1012.5

1.0

Σ

1.800

872.2

3.4

Σ

1.776

964.9

3.6

Π

1.774

810.6

6.0

Σ

1.835

834.9

2.9

Σ

1.795

951.3

2.2

Σ

1.788

819.5

12.3

þ

b

re

1

UO

a

stateb

I I

PuO2þ

5

AmO

8

AmOþ

7

AmO2þ

6

CmO

9

CmOþ

8

CmO2þ

7

The vibrational data are given in cm1 and the bond distances in Å. From ref 57.

vibrational parameters of the monoxides (in the electronic ground state57) are compiled in Table 1. We note that our optimized spin-free bond distances agree well with those from the spinorbit CASPT2 calculations of Infante et al.,57 as minor deviations were found only for PaO2þ, PuOþ, PuO2þ, and AmO2þ attributed to the differences in the theoretical level. Experimental data on spectroscopic constants have been reported previously for ThO, ThOþ, UO, UOþ, and PuO. They are compared with our and selected theoretical results from the literature in Table 2. Dewberry et al. determined the bond distance in ThO with very high accuracy using microwave spectroscopy.6 The bond distance of ThOþ has been derived from pulsed field ionizationzero-kinetic-energy photoelectron spectroscopic (PFIZEKE) measurements.7 As the data in Table 2 show, our calculations (similarly to the previous ones including spinorbit coupling57) overestimate the bond distance in ThO and ThOþ by 0.02 Å. This error is accompanied by an underestimation of the harmonic vibrational frequencies4,5,7 of the two species by ca. 20 cm1. On the other hand, the computed first anharmonicities (2.3 and 2.6 cm1 for ThO and ThOþ, respectively) agree well with the experimental values of 2.39 and 2.45 cm1, respectively.4,5,7 We note that most earlier theoretical results show similar (or larger) errors for the bond distances and vibrational frequencies of ThO (cf. Table 2). Better agreement has been achieved only by the DFT calculations of Goncharov7 and Andrews 15 and by the CCSD(T) calculations of Buchachenko. 33 Additional IR data on ThO include the IR absorptions of the species isolated in neon13,15 and argon12,14 matrices. On the basis of the experimental gas-phase harmonic frequency and

anharmonicity,4,5 we can derive the matrix shifts of 4 and 12 cm1 for the Ne and Ar matrices, respectively. Experimental data for the bond distances in UO and UOþ have been reported by Heaven et al. using high-resolution photoelectron spectroscopy.8,9 In contrast to the above shown deviations for the bond distances in the ThO species, the agreement of our CASPT2 results with the experimental ones of UO and UOþ is excellent (cf. Table 2). We comment briefly on the experimental vibrational spectrum of UO, which showed a significant vibronic perturbation of the first vibrational interval for the ground state due to the interactions between the three lowest energy Ω = 4 states. The observed vibrational interval in the gaseous phase was 882.4 cm1.8 Kaledin et al. derived the deperturbed ground state vibrational interval to be 841.9 cm1 and the deperturbed harmonic vibrational constant to be ωe = 846.5 cm1.8 We used this latter value in our comparison with the computed data because the routine quantum chemical frequency calculations do not account for vibronic effects. The vibronic perturbation appears also in the matrix-isolated species; however, there the guesthost interactions affect the magnitude of perturbations, and hence the vibronic effect is different in the various matrices. This can be seen in the anomalous band at 889.5 cm1 in the IR spectrum of UO isolated in the Ne matrix18 being somewhat larger than the perturbed frequency observed in the gaseous phase. On the other hand, the absorptions in the Ar and Kr matrices (820 and 815 cm1, respectively,16,17 cf. Table 2) seem to be quite reasonable compared to the deperturbed gas-phase value. The derived shift in the Ar matrix, ca. 20 cm1, is somewhat larger than that observed for ThO (12 cm1, vide supra). The matrix shift of Kr is ca. 25 cm1. Comparing our computed data with the previous theoretical results on UO, the SO-CASPT2 results by Paulovic et al.37 match well for both the bond distances and vibrational frequencies. We note, however, that their spin-free CASPT2 bond distances deviate slightly, while these vibrational frequencies deviate considerably (e.g., ωeUOþ by 163 cm1, cf. Table 2) from our and the experimental values. (The difference between our and their theoretical level was that Paulovic et al. used a somewhat smaller active space.) In the other theoretical studies listed in Table 2, the bond distances are overestimated. Good agreement was found with the vibrational frequency computed by Krauss and Stevens.35,36 The only experimental information found on plutonium monoxide is the matrix-isolation IR study of Green and Reedy.20 We note the resemblance of the matrix IR data of UO and PuO in Ar and Kr, in both cases the matrix shift from Ar to Kr being 5 cm1 (cf. Table 2). On this basis we assume analogous matrix shifts for UO and PuO and estimate the harmonic fundamental of PuO to be ca. 849 cm1 in the gaseous phase. The data of UO and PuO compiled in Table 2 suggest that our calculations performed similarly for the two oxides: like obtained for UO, our computed harmonic frequency for the PuO stretching is larger by 36 and 41 cm1 than the experimental values from the Ar and Kr matrices, respectively. With respect to our above estimated “gas-phase frequency” of PuO, our computed value is larger by the same magnitude (9 cm1) as obtained in the case of UO. These resemblances suggest that our computed bond distance for PuO possesses a reliability comparable to that of UO (within 0.002 Å of the experimental equilibrium bond distance). 6648

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Table 2. Available Experimental and Selected Computed Resultsa on the AnO0/þ/2þ Species species ThO

method

phase/state

re

frequency

ref

PFI-ZEKE

gas

1.840

MW

gas

1.84018613(24)

7

electron spectroscopy

gas

ωe = 895.77

IR

Ne matrix

887.1

13,15

IR

Ar matrix

878.8

12,14

6 4,5

ωexe = 2.39

ωexe = 1.9(9) 1 þ

Σ Σ

CCI DiracFockRoothaan

1 þ 1 þ

Σ

MRCIb

1.862

28

1.886

29

882

1.853

15

937

1.874

30

Σ

866

1.877

31

Σ0þ

856

1.866

32

898 898.7

1.846 1.845

7 33

879

1.863

present study

894

(1.840)

present study

1.807

7

960

1.814

7

931 941

1.827 (1.807)

present study present study

Σ

1 þ

Σ

1 þ

CASSCF DK3-SO-CASPT2b

1

1 þ

Σ 1 þ Σ

B3PW91 CCSD(T)

26 27

867

1 þ

MRCI

1.923 1.873

865

CASSCF PW91

852 923

ωexe = 3.19 1 þ

Σ

CASPT2 empiricald ThOþ

PFI-ZEKE

ωe = 954.97(6)

gas

ωexe = 2.45(3) 2 þ

Σ

B3PW91

2 þ

Σ

CASPT2 empiricald UO

photoelectron spectroscopy

ωe = 846.5(6)c

gas

1.8383(6)

8

(ω = 882.4) IR

Ne matrix

889.5

18

IR

Ar matrix

819.8

17,18

IR

Ar matrix

820

16

ωexe = 2.5(8) IR B3LYP

Kr matrix 5 Γ 3 

Σ

MP2

16 18

1035.9

1.833

34

845

1.889

35,36

5

CASPT2

5

920

1.850

37

SO-CASPT2

5

855

1.842

37

CASPT2

5

858

1.837

present study

840

(1.8383)

present study

I I I4 I

empirical

photoelectron spectroscopy

gas

ωe = 911.9(2) ωexe = 2.39(4)

1.801(5)

9

MCSCF

4

935

1.842

35

CASPT2

4

1074

1.796

37

SO-CASPT2

4

912

1.802

37

CASPT2

4

911

1.799

present study

900

(1.801)

present study

I I I4.5 I

d

empirical PuO

1.850

MCSCF

d

UOþ

815.45 846

IR

Ar-matrix

822.3

20

IR estimatedf

Kr-matrix gas

817.27 849

20 present study

5 

Σ

QCISD

781

1.83

38 39

ωexe = 2.77 SO-CASPT2b

856

1.820

Π

820

1.834

39

734

1.89

40

7

B3LYP

B3LYP

Π0

7

b e

6649

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Table 2. Continued species PuOþ

method CASPT2 B3LYP b

AmO

present study

Π

899

1.788

39b

Π0.5

881

1.789

39

Π

919

1.777

present study

Σ

525

1.96

41

Γ

961

1.720

39b

6

5 

41,42

B3LYP

5

SO-CASPT2b

5

Γ2

872

1.731

39

CASPT2 CASPT2

5

Γ 8 þ Σ

1013 917

1.720 1.803

present study 43

872

1.800

present study

924

1.758

43

Σ

965

1.776

present study

Σ

858

1.824

43

Σ

835

1.835

present study

CASPT2 CASPT2

CmO

1.818 1.875

SO-CASPT2

CASPT2 þ

858 652

CASPT2 b

AmO

Π Σ

6

B3LYP

ref

6  6

B3LYP

re

frequency

7

b

PuO2þ

phase/state

8 þ

Σ

7 þ

Σ

7 þ

CASPT2

9

CASPT2

9

a The vibrational data are given in cm1 and the bond distances in Å. The harmonic experimental vibrational frequencies are indicated by the ωe symbol (the computed values are in all the cases the harmonic ones). Experimental uncertainties are given in parentheses. b From the numerous calculated data from the given reference, only selected high-level results are shown here. For the full set of results see the reference. c UO suffers from strong perturbation effects. The ωe value is the derived unperturbed result. d Harmonic frequencies evaluated at the experimental reference bond distances (indicated in parentheses) from the trendlines fitted to data obtained with eight exchange-correlation DFT functionals. For details see text. e Electronic state not given in the original paper. f Estimated by assuming analogous general matrix shifts for UO and PuO on the basis of the same shifts from Ar to Kr.

Previous theoretical data on Pu monoxides include the SOCASPT2 calculations of La Macchia et al.39 and the DFT calculations of Gao et al.38 and Li et al.82 Our computed results are in good agreement with the SO-CASPT2 results of La Macchia et al.39 except for the vibrational frequency of PuO2þ, for which the value of La Macchia et al. seems to be too small. Therefore, the increasing trend of frequencies with the decreasing bond distance from PuO to PuO2þ seen in our computer results (cf. Table 1) does not appear in the computed frequencies of that work. The DFT frequencies of Gao et al.38 and Li et al.82 are smaller than the CASPT2 ones and thus deviate more from the experimental values. Regarding the previous theoretical study of AmO, AmOþ, and CmO43 performed at a level of theory similar to the present one, the differences in the computational details can account for the deviations from the present data (cf. Table 2). Our consistent calculations on the light actinide monoxides allow us to draw some general conclusions. Most obvious is the trend in the bond distances when the neutral molecule is ionized. The AnO bond becomes gradually shorter with the loss of one and two electrons, respectively. However, no clear trend can be observed in the spin-free CASPT2 bond distances from ThO to CmO (neither in the ions) similarly to the recently reported spinorbit distances.57 This is the consequence of the somewhat different electronic structures of these actinide oxides. Hence, the frequently supposed analogy within the actinide row or with the lanthanide series2125 may often fail for actinides. Our most important observation on the computed harmonic vibrational frequencies is that they are generally in agreement with the changes in the bond distances. The species with shorter (=stronger) bonds require generally larger energy for the stretching motion. The only exceptions are AmO2þ and CmO2þ where the ωe values are quite low. Inspecting the potential energy curves of these dications, we found that their curves have a flatter character around the minimum compared to the curves of the other dications. The

case is demonstrated by the potential energy curves of PuO0/þ/2þ and CmO0/þ/2þ in Figure 1. In agreement with the flatter character of their curves, AmO2þ and CmO2þ have larger first anharmonic constants (6.2 and 12.3 cm1, respectively) as compared to the other monoxides having values generally around 23 cm1. To clarify the electronic basis of this feature we checked the populations of the bonding orbitals of all the AnO species under study. We found that around the equilibrium bond distance the bonding orbitals in AmO2þ and CmO2þ are less populated by ca. 0.3 e than the bonding orbitals in the other monoxides. This missing ca. 0.3 e populates in AmO2þ and CmO2þ an antibonding orbital consisting of mainly 2pz of O and 5fσ of An. This should be responsible for the relatively longer equilibrium bond distances and the larger anharmonic character of the potential energy curve in the vicinity of the minimum. 3.3. Vibrational Frequencies of the Dioxides. For actinide dioxides, there is less experimental gas-phase information available than for the monoxides. No experimental bond distance has been reported for these molecules, while the only experimental vibrational data on gaseous species are the bending fundamental of UO2 from resonance-enhanced multiphoton ionization (REMPI)10 and the symmetric stretching and bending fundamentals of UO2þ from PFI-ZEKE measurements.11 In addition, some stretching frequencies of ThO2, UO2, UO2þ, and PuO2 have been reported from matrix-isolation IR experiments.1220 The molecular geometries and vibrational frequencies of the dioxide species obtained by eight DFT exchange-correlation functionals are given in the Supporting Information. We note that most of these bond distances were reported recently,83 where DFT data (including other exchange-correlation functionals like PW91, M05, M06, M06-L, MOHLYP, and MPW3LYP) were compared to bond distances and dissociation energies obtained by sophisticated multiconfigurational ab initio methods. In that comparison, none from the tested DFT levels showed a uniform agreement with the reference SO-CASPT2 6650

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Figure 1. Potential energy curves of PuO0/þ/2þ and CmO0/þ/2þ around the equilibrium bond distances. The scale of the y axis was the same for all six curves, but due to the large differences in the absolute energies of the species the curves of the ions were shifted to fit into one graph with those of the neutral oxides.

Figure 2. DFT frequencybond distance correlation for the three fundamentals of UO2.

bond distances. Thus, to obtain reliable vibrational frequencies for the title dioxides, we cannot pick a single DFT level. We analyzed here jointly the DFT equilibrium bond distances and vibrational frequencies and observed a good correlation between the two parameters. We demonstrate this in Figure 2 with the three fundamentals of UO2. In the bond distance (re) range covered by our computations, a trendline can be fitted to the computed data, on which basis the harmonic vibrational frequencies can be evaluated to any bond distance within (and near) the range. The quality of these empirical frequencies depends on two main factors: (i) the accuracy of the reference bond distance and (ii) the accuracy of the DFT potential energy surfaces around the computed equilibrium bond distances. For example, if the potential energy surface is too flat, we obtain too small frequencies at the accurate bond distance. This empirical approach has been tested on the monoxides with available experimental bond distances: ThO, ThOþ, UO, and UOþ. The empirical harmonic vibrational frequencies are included in Table 2 in the rows indicated by “empirical”. There is an excellent agreement with the gas-phase experimental frequencies of ThO and UO (within 2 cm1) for which very accurate experimental bond distances are available (cf. Table 2). For ThOþ and UOþ the deviations are 14 and 8 cm1, respectively, but here the reference bond distances may be less accurate: the experimental bond distance of UOþ has a larger uncertainty, while that of ThOþ is not given in ref 7. Note that with the exception of UOþ the empirical vibrational frequencies are closer to the experimental results than the frequencies computed on the basis of the spin-free CASPT2 potential energy curve.

In the lack of proper gas-phase experimental data on the dioxides, we cannot be sure of the same excellent performance of the empirical approach on these latter molecules. However, the similar bonding in the actinide monoxides and dioxides suggests a pretty good reliability of the empirical dioxide frequencies as far as the reference bond distance is correct. Hence the selection of the reference bond may be the critical issue here. From the numerous computations13,15,18,19,40,4457,84 we selected the recently reported SO-CASPT2 results of Infante et al.57 for this purpose. The only exception is UO2, for which we used the bond distance of 1.770 Å obtained recently by four-component DCIHFSCCSD calculations85 being in good agreement with an earlier SO-CASPT2 result of 1.766 Å.46 The vibrational frequencies of the dioxides obtained for the reference bond distances are summarized in Table 3. The parameters of the trendlines are given as Supporting Information. We note the different character of the trendlines for the stretching frequencies of ThO2þ and PaO22þ (cf. Supporting Information), which may result in a larger uncertainty of the frequency data of these ions. We calculated also the anharmonic vibrational frequencies at the B3LYP or B3P86 levels by calculation of the quartic force fields using the “Freq=Anharmonic” keyword of Gaussian 03. The empirical harmonic frequencies in Table 3 have been corrected by the obtained anharmonic effects and are presented in separate columns in Table 3. A compilation of literature data on the vibrational frequencies of actinide dioxides (both experimental and theoretical) is given in Table 4. We start our discussion with UO2þ for which most experimental information is available for the gaseous phase.11 The empirical anharmonic frequencies evaluated on the basis of the 6651

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Table 3. Empirical Harmonic and Anharmonic Vibrational Frequencies of the Dioxide Species on the Basis of DFT Calculationsa DFT anharmonicc

DFT harmonic r eb

νas

νs

β

νas

νs

β

1.923

719

772

151

715

767

151

1.896d

766

818

151

762

813

151

ThO2þ

1.832

635

790

63

623

780

51

ThO22þ

1.903

736

701

107

727

695

107

PaO2 PaO2þ

1.816 1.767

855 994

900 935

79 106

852 989

899 931

82 108

PaO22þ

1.726

602

981

138

594

961

144

UO2

1.770

959

907

150

956

904

150

1.790d

923

868

136

920

865

136

1.745

1019

951

155

1013

946

155

species ThO2

UO2þ

1.758d

991

921

144

985

916

144



1.710

1089

993

127

1082

987

127

NpO2 NpO2þ

1.761 1.723

942 1031

882 954

188e 216

939 1023

880 947

185 215

NpO22þ

1.700

1092

983

154

1085

977

154

PuO2

1.744

940

870

188

934

862

184

PuO2þ

1.704

1035

945

262

1027

938

261

PuO22þ

1.675

1110

992

239e

1102

984

237

AmO2

1.807

819

761

102

815

757

105

AmO2þ

1.721

981

862

264

970

851

262

AmO22þ CmO2

1.679 1.832

1057 788

918 731

293 92

1045 778

906 721

291 90

CmO2þ

1.746

895

778

193

870

752

192

CmO22þ

1.674

996

817

313

964

783

309

UO2

The fundamentals (in cm1) νas, νs, and β correspond to asymmetric stretching, symmetric stretching, and bending, respectively. All the species except for ThO2 are linear, in which β is doubly degenerated. The harmonic frequencies have been evaluated at the SO-CASPT2 reference bond distance using the presently developed empirical approach. For details see text. b The reference bond distances (Å) were taken from ref 57 except for UO2. The 1.770 Å bond distance of UO2 has been computed at the four-component DC-IHFSCCSD level in ref 85. c Anharmonic effects were calculated by the Freq=Anharmonic keyword of Gaussian 03, and the corrections were added to the empirical harmonic values given in columns 35. d The values of 1.896, 1.790, and 1.758 Å were estimated in the present study for ThO2, UO2, and UO2þ, respectively, on the basis of the experimental vibrational data. They are likely more reliable for the equilibrium bond distances than the SO-CASPT2 values. e The DFT calculations with Gaussian03 gave erroneously two close frequency values for the degenerate bending fundamentals (difference of 20 cm1 for NpO2 and 9 cm1 for PuO22þ). The data given in the Table are the average values. a

SO-CASPT2 bond distance (946 and 155 cm1) compare reasonably well with the experimental values of 921(4) and 145.5 cm1 (Table 4): the stretching fundamental is overestimated by ca. 25 cm1, the bending fundamental by ca. 7 cm1. The anharmonic parameters derived by Merritt et al.11 are in excellent agreement with the anharmonic effects from our DFT calculations (cf. Tables 3 and 4). From the reported matrix-isolation IR1719 and computed19 data on UO2þ we can estimate the gas-phase value of the asymmetric stretching (νas) fundamental. The matrix-isolation IR data show the following matrix shifts: with respect to Ne there

is a red-shift of 28 cm1 in Ar, 40 cm1 in Kr, and 51 cm1 in Xe. Wang et al.19 carried out DFT calculations on the UO2þ, UO2(Ne)6þ, UO2(Ar)5þ, UO2(Kr)5þ, and UO2(Xe)5þ species to model the effect of matrix on the vibrational frequencies. From their data, we derived somewhat smaller matrix shifts for νas than found experimentally, viz., 27, 32, and 42 cm1 in Ar, Kr, and Xe matrices, respectively, with respect to the frequency in the Ne matrix. The calculations of Wang et al.19 predicted matrix shifts in Ne with respect to the isolated UO2þ molecule values of 2 and 5 cm1 for the asymmetric and symmetric stretching fundamentals, respectively. On the basis of the above information, we estimate the gas-phase νas frequency to be ca. 983 cm1, which is overestimated by our empirical νas by 30 cm1, slightly more than the νs fundamental (vide supra). The consistent overestimation of all three fundamentals by our empirical approach implies that the main source of error lies in the reference bond distance. We get an excellent agreement with the three gas-phase fundamental frequencies using a bond distance of 1.758 Å instead of the SO-CASPT2 value (cf. Tables 3 and 4). The amount and quality of experimental vibrational data on UO2 facilitates a straightforward analysis of these data as well. However, not all the matrix-isolation IR data are useful for our purpose because those obtained in the heavier rare gas matrices suffer from the change of electronic ground state upon interactions with the rare gas atoms.18,48,86 It has been clarified by quantum chemical calculations and new matrix-isolation IR experiments that while the gas-phase electronic ground state (3Φu) is retained in the Ne matrix the interactions with the more polarizable Ar and Kr atoms stabilize the low-lying 3Hg state in those matrices.18,48,86 This is the reason that the observed frequency of UO2 in Ar16,17,86 (775.7 cm1) is considerably lower than the frequency in the Ne matrix15,18,86 (914.8 cm1). Therefore, only the latter value will be discussed in the following. A joint analysis of the 3Φu experimental and theoretical data of UO2 facilitates the estimation of the gas-phase asymmetric stretching frequency and a (possibly) reliable equilibrium bond distance. Assuming a matrix shift of 5 cm1 by Ne, the “gas-phase anharmonic value” of νas would be 920 cm1. The empirical νas frequency using the literature reference bond distance85 is considerably (by 39 cm1) larger, similar to the empirical bending frequency as compared to the gas-phase value from REMPI experiments.10 Utilizing the frequencybond distance relationship (see Supporting Information) and the computed anharmonicity, an equilibrium bond distance of 1.790 Å would fit to the “gas-phase νas value”. In turn, based on this bond distance, the empirical νs and β fundamental frequencies would be 865 and 136 cm1, respectively. This empirical bending frequency is larger by only 15 cm1 than the gas-phase REMPI value, which may partly be attributed to RennerTeller vibronic coupling10 not taken into account by the theoretical model in the DFT calculations. Regarding the other actinide dioxides, some matrix-isolation IR data have been reported for ThO21215 and PuO2.20 The fundamental stretching frequencies of ThO2 have been observed in Ne13,15 and Ar12,14 matrices. The matrix shifts of 21.1 and 21.8 cm1 (Ar with respect to Ne) are consistent for the two stretching fundamentals of ThO2 and are somewhat smaller than the value of 28 cm1 observed for the νas of UO2þ (vide supra). Note that also for ThO somewhat smaller matrix shifts were found than for UO (vide supra), hence the interaction with the matrix seems to depend more on the bonding characteristics of 6652

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Table 4. Available Experimental and Selected Computed Resultsa on the AnO20/þ/2þ Species frequency species ThO2

PaO2þ UO2

UO2

þ

UO22þ

method b

phase/state

νs

νas

β

re

ref

estimated

gas

762

813

IR

Ne matrix

756.8

808.4

IR

Ar matrix

735.0

787.3

PW91

1

759

812

157

1.911

15

DHF

1

761

896

139

1.898

44

B3LYP

1

760

802

159

1.906

13

empirical

1

715

767

151

1.923

present study

empirical DHF

1

A1 Σgþ

762 1158

813 1029

151 130

1.896 1.742

present study 44

empirical

1

Σ0gþ

989

931

108

1.767

present study

REMPI

gas

estimatedb

gas

920

IR

Ne matrix

914.8

18

IR

Ar matrix

776

17

IR

Ar matrix

776.10(5)

IR MP2

Kr matrix 3 þ Σu

B3LYP

3

CASPT2c

3

SO-CASPT2c

3

B3LYPc

3

CCDc

3

MP2c B3LYPc PW91

3

SO-VWN-BP

3

VWN-BP

3

empirical

3

empirical

3

PFI-ZEKE

gas

A1 A1 A1 A1

1

present study 13,15 12,14

121

10 present study

16

767.95 1022

947

294

1.779

16 45

Φu

931

874

138

1.800

18

Φu

932

809

1.806

46

Φ2u

923

948

1.766

46

Φu

937

875

1.794

46

Φu

958

927

168

1.766

47

3

Φu

913

896

149

1.795

47

Φu 3 Φu

933 919

880 856

222

1.764 1.807

47 48

Φu

934

863

1.803

49

Φu

948

853

1.813

49

Φ2u

956

904

150

1.770

present study

Φ2u

920

865

136

1.790

present study

ω = 921(4)

ω = 145.5

ωx = 2.0(22)

ωx = 0.3

11

estimatedb IR

gas Ne matrix

IR

Ar matrix

952.3

17,19

IR

Kr matrix

940.6

17,19

IR

Xe matrix

B3LYP

2

CCDc

2

MP2c

2

B3LYPc CASPT2

2

PW91

2

empirical

2

empirical

2

DHF

1

4-CCSD

1

4-CCSD(T)

1

CCSDc B3LYPc

983 980.1

present study 18,19

929.0

19

Φu

1010

Φu

1031

971

146

1.744

47

Φu

955

901

101

1.80

47

Φu Φu

1001 942

916 858

191

1.746 1.773

47 50

2

Φu

936

148

1.764

18

987

911

1.773

19

1013

946

155

1.745

present study

Φ2.5u

985

916

144

1.758

present study

Σgþ

1294

1234

246

1.650

44

Σ0gþ

1168

1040

180

1.696

52

1.715

52

1.702 1.706

51 51

Φ2.5u

Σ0gþ

974

Σgþ Σgþ

1148 1143

1063 1041

180 150

Σgþ

1140

1041

161

Σgþ

1066

959

Σgþ

1179

1100

1 1

B3LYP

1

CASPT2

1

CCDc

1

6653

194

1.705

18

1.705

50

1.678

47

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A

ARTICLE

Table 4. Continued frequency species

method

phase/state

1095

997

1082

987

Σg

892

Σg

869 939

880

2

1075

969

1.717

49

Φ2.5u Ar matrix

1085 794.2

977

154

1.700

present study 20

1.800

53

SO-VWN-BP

4

VWN-BP

4

empirical

4

NpO22þ

VWN-BP

2

PuO2

empirical IR

H3.5g Φu

IR

Kr matrix

MP2

5

PuO2

PuO22þ

present study

870

1.778

49

868

1.773

49

1.761

present study

185

786.8

20 306

Σgþ

828

792

109

1.870

54

Σgþ

781

772

1.833

49

Σgþ

866

791

1.812

49

758

614

1.88

40

1.818 1.744

39 39

1.744

present study

VWN-BP SO-VWN-BP

5 d

Σgþ 5 Φ1u

5

Φ1u

5

934

862

Φ1.5u

SO-CASPT2 empirical

4

1027

B3LYPc

3

1144

AQCC

3

Φ1.5u Hg

3

Hg

B3LYP SO-CASPT2 SO-CASPT2

3

empirical

3

938

261

1014

229

1005 1100

Hg 3 H4g H4g H4g

184

962

Hg

3

181

773 837

4

c

49

1.710

656

5

VWN-BP

1.720 127

744

5

c

126 144

Σgþ

CCSD(T)c

empirical

47 47

Σ0gþ

empirical

1

þ

ref

Σgþ

1

B3LYPc SO-CASPT2c

re 1.728 1.687

923 1027

VWN-BP

B3LYP

β

1024 1126

1

NpO2

νs

νas

Σgþ 1 þ Σg

MP2c B3LYPc

39

1.704

present study

1.688

51

1.670

55

953

1.703

49

1004 1065

1.678 1.699

39 56

1019 1102

1.704

984

237

1.675

39

1.675

present study

a The vibrational data are given in cm1 and the bond distances in Å. In the rows “empirical” we repeat the empirical anharmonic frequencies from Table 3. In the case of ThO2, UO2, and UO2þ the first set of empirical data refers to the literature reference bond distance, while the second set is based on reference bond distances estimated in the present study. The harmonic experimental vibrational frequencies are indicated by the ωe symbol. Experimental uncertainties are given in parentheses. b Estimated on the basis of matrix shifts. c From the numerous calculated data from the given reference, only selected high-level results are shown here. For the full set of results, see the reference. d Electronic state is not given in the original paper.

the actinide than on the different shape of the molecules (UO2þ is linear while ThO2 is bent57). These two examples suggest that the uranium valence 5f orbitals are more favorable for interaction with the rare gas atoms than the valence 6d orbitals predominating in thorium. Due to this larger interaction, the heavier actinides may generally be more sensitive to matrix effects. Assuming a matrix shift of 5 cm1 in Ne, the “gas-phase anharmonic stretching frequencies” of ThO2 can be estimated to be around 762 and 813 cm1. With respect to these frequencies, the empirical stretching values based on the SO-CASPT2 bond distance are underestimated by ca. 45 cm1. The consistent underestimation of the two fundamentals suggests that the main source of error lies again in the SO-CASPT2 bond distance. This suggestion is supported by our previous experience on ThO and ThOþ, where the theoretical SO-CASPT2 bond distances proved to be larger by ca. 0.02 Å than the experimental ones.57 These examples indicate a general error of this SO-CASPT2 level for thorium oxides. On the basis of the frequencybond distance relationship derived from the DFT calculations and taking into account the

matrix shift and anharmonicity effects, we predict the equilibrium bond distance of ThO2 to be around 1.896 Å. We note that from our DFT calculations the B3LYP results differ only marginally from the “gas-phase frequencies” estimated from the IR data in the Ne matrix using a correction of 10 cm1 for matrix shift and anharmonicity effects. Accordingly, the B3LYP optimized bond distance is 1.898 Å, and the bond angle is 119.0°, being in excellent agreement with the value of 122.5 ( 2° derived from isotope-substitution IR experiments.12 The IR spectra of PuO2 recorded in Ar and Kr matrices call also for a careful interpretation. At first sight our empirical anharmonic asymmetric stretching frequency is considerably larger than the reported values from the matrix-IR spectra.20 Assuming 30 cm1 for the matrix effect in Ar, the “gas-phase value” derived from the matrix-IR data would be lower by ca. 110 cm1 than our empirical frequency. It is unlikely that the SOCASPT2 bond distance of PuO2 would have such a large error. We suspect a change of the electronic ground state upon interaction with the rare gas atoms in the matrix, as shown for UO2 (vide supra). Indeed, recent theoretical results obtained by 6654

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A SO-CASPT2 computations indicate that the 5Φ1u ground and the 5Σ0gþ excited states are very close in energy (at the spin-free level, 5Σgþ is the ground state, while the energy ordering is interchanged upon inclusion of spinorbit effects39), while the bond distances in the two electronic states differ considerably (cf. Table 4). On this basis we suggest that upon interaction with the more polarizable Ar and Kr rare gas atoms the 5Σ0gþ state is stabilized (and is present) in the Ar and Kr matrices. The asymmetric stretching frequencies computed recently for the 5 þ Σg state by Archibong and Ray54 and Liao et al.49 agree reasonably with the matrix-IR values (cf. Table 4) supporting the hypothesis of interchanging ground states in the matrix. In addition to the above-discussed cases, there are a few computed data on PaO2þ, UO22þ, NpO2, NpO22þ, PuO2þ, and PuO22þ as listed in Table 4. Most previous papers report bond distances longer than the reference ones accompanied generally by smaller vibrational frequencies. Liao et al. reported their NpO2 data for the 4Σg state,49 while the ground state of the molecule has been determined to be 4H3.5g by recent SOCASPT2 calculations.57 Nevertheless, the bond distances from the two papers are quite close as are the vibrational frequencies in Table 4.

4. CONCLUSIONS In the present paper, we reported the vibrational frequencies for the electronic ground states of the actinide oxides AnO and AnO2 (An = Th, Pa, U, Np, Pu, Am, Cm). Our consistent CASPT2 calculations on the monoxides revealed no trend in the actinide row but showed a decreasing trend in the bond distance from the neutral molecule to the dication. The frequencies increase generally in this order except for AmO2þ and CmO2þ, which possess a somewhat larger anharmoncity compared to those (mostly 23 cm1) of the others. The fundamental frequencies of the dioxides have been evaluated by an empirical approach which was tested on the accurate gas-phase experimental data of ThO, ThOþ, UO, and UOþ. For the dioxides, the SO-CASPT2 bond distances57 are presently the most sophisticated geometrical data, hence we used them as reference bond distances in the empirical approach. We believe that the reliability of our empirical stretching frequencies is comparable to those of the SO-CASPT2 bond distances (except where vibronic effects appear). The empirical bending frequencies are also promising, as we obtained a very good agreement for UO2þ. We analyzed the available literature experimental data and the found characteristics combined with our computed data facilitated the prediction of molecular properties of several gas-phase species. On the basis of the observed analogous matrix shift in the IR spectra of UO and PuO, we estimated the harmonic gas-phase frequency of PuO. Utilizing the small matrix shifts in Ne and the computed anharmonicities, we derived reliable gas-phase stretching frequencies for ThO2, UO2, and UO2þ. The same differences between the computed and experimental frequencies of UO and PuO suggest that our computed bond distance for PuO (1.818 Å) is likely within 0.002 Å of the experimental equilibrium bond distance. On the basis of the experimental vibrational data and the derived frequencybond distance relationships, we provided good-quality estimates for the equilibrium bond distances of ThO2, UO2, and UO2þ (1.896, 1.790, and 1.758 Å, respectively). We believe that they are a

ARTICLE

better approximation of the (yet unknown) experimental values than the SO-CASPT2 bond distances. Further important information from the present joint analysis is that the electronic ground state of PuO2 in Ar and Kr matrices is probably different from that in the gaseous phase, as observed previously for UO2.18,86

’ ASSOCIATED CONTENT

bS

Supporting Information. The computed molecular geometries and vibrational frequencies of the dioxide species obtained by the eight DFT exchange-correlation functionals, the parameters of the fitted trendlines for the frequency-geometry data, and calculated anharmonic corrections for the dioxides. This material is available free of charge via the Internet at http:// pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected].

’ ACKNOWLEDGMENT The Seventh Framework Programme of the European Commission (Collaboration Project No. 211690, F-BRIDGE) and the Hungarian Scientific Research Foundation (OTKA No. 75972) are acknowledged for financial support. The authors thank Dr. I. Infante and Dr. M. Klipfel for helpful discussions. ’ REFERENCES (1) Konings, R. J. M.; Wiss, T.; Gueneau, C. Nuclear Fuels. In The chemistry of the actinide and transactinide elements; Edelstein, N. M., Fuger, J., Morss, L. R., Eds.; Springer: Berlin, 2010; Vol. 6. (2) Olander, D. Fundamental aspects of nuclear reactor fuel elements. Tech. Rep. 3561 TID-26711- P1, Technical Information Center, Office of Public Affairs Energy Research and Development Administration, 1976. (3) Maeda, K.; Sasaki, S.; Kato, M.; Kihara, Y. J. Nucl. Mater. 2009, 389, 78. (4) Edvinsson, G.; Selvin, L.-E.; Aslund, N. Ark. Fys. 1965, 30, 28. (5) Edvinsson, G.; Lagerqvist, A. Phys. Scr. 1984, 30, 309. (6) Dewberry, C. T.; Etchison, K. C.; Cooke, S. A. Phys. Chem. Chem. Phys. 2007, 9, 4895. (7) Goncharov, V.; Heaven, M. C. J. Chem. Phys. 2006, 124, 064312. (8) Kaledin, L. A.; McCord, J. E.; Heaven, M. C. J. Mol. Spectrosc. 1994, 164, 27. (9) Goncharov, V.; Kaledin, L. A.; Heaven, M. C. J. Chem. Phys. 2006, 125, 133202. (10) Han, J.; Kaledin, L. A.; Goncharov, V.; Komissarov, A. V.; Heaven, M. C. J. Am. Chem. Soc. 2003, 125, 7176. (11) Merritt, J. M.; Han, J.; Heaven, M. C. J. Chem. Phys. 2008, 128, 084304. (12) Gabelnick, S. D.; Reedy, G. T.; Chasanov, M. G. J. Chem. Phys. 1974, 60, 1167. (13) Zhou, M.; Andrews, L. J. Chem. Phys. 1999, 111, 11044. (14) Kushto, G. P.; Andrews, L. J. Phys. Chem. A 1999, 103, 4836. (15) Andrews, L.; Zhou, M.; Liang, B.; Li, J.; Bursten, B. E. J. Am. Chem. Soc. 2000, 122, 11440. (16) Gabelnick, S. D.; Reedy, G. T.; Chasanov, M. G. J. Chem. Phys. 1973, 58, 4468. (17) Hunt, R. D.; Andrews, L. J. Chem. Phys. 1993, 98, 3690. (18) Zhou, M.; Andrews, L.; Ismail, N.; Marsden, C. J. Phys. Chem. A 2000, 104, 5495. 6655

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656

The Journal of Physical Chemistry A (19) Wang, X.; Andrews, L.; Li, J.; Bursten, B. E. Angew. Chem., Int. Ed. 2004, 43, 2554. (20) Green, D. W.; Reedy, G. T. J. Chem. Phys. 1978, 69, 544. (21) Murad, E.; Hildenbrand, D. L. J. Chem. Phys. 1980, 73, 4005. (22) Haire, R. G. J. Alloys Compd. 1994, 213/214, 185. (23) Santos, M.; Marc-alo, J.; de Matos, A. P.; Gibson, J. K.; Haire, R. G. J. Phys. Chem. A 2002, 106, 7190. (24) Santos, M.; Marc-alo, J.; Leal, J. P.; de Matos, A. P.; Gibson, J. K.; Haire, R. G. Int. J. Mass Spectrom. 2002, 228, 457. (25) Gibson, J. K. J. Phys. Chem. A 2003, 107, 7891. (26) Marian, C. M.; Wahlgren, U.; Gropen, O.; Pyykko, P. J. Mol. Struct. (Theochem) 1988, 169, 339. (27) Watanabe, Y.; Matsuoka, O. J. Chem. Phys. 1998, 107, 8182. (28) K€uchle, W.; Dolg, M.; Stoll, H.; Preuss, H. J. Chem. Phys. 1994, 100, 7535. (29) Seijo, L.; Barandiaran, Z.; Harguindey, E. J. Chem. Phys. 2001, 114, 118. (30) Watanabe, Y.; Matsuoka, O. J. Chem. Phys. 2002, 116, 9585. (31) Paulovic, J.; Nakajima, T.; Hirao, K.; Seijo, L. J. Chem. Phys. 2002, 117, 3597. (32) Paulovic, J.; Nakajima, T.; Hirao, K.; Lindh, R.; Malmqvist, P. Å. J. Chem. Phys. 2003, 119, 798. (33) Buchachenko, A. A. J. Chem. Phys. 2010, 133, 041102. (34) Wang, H. Y.; Zhou, Z. H.; Gao, T.; Fu, Y. B.; Wang, X. L.; Sun, Y. Mol. Phys. 2000, 98, 875. (35) Krauss, M.; Stevens, W. J. Chem. Phys. Lett. 1983, 99, 417. (36) Krauss, M.; Stevens, W. J. Mol. Phys. 2003, 101, 125. (37) Paulovic, J.; Gagliardi, L.; Dyke, J. M.; Hirao, K. J. Chem. Phys. 2005, 122, 144317. (38) Gao, T.; Wang, H.-Y.; Yi, Y.-G.; Tan, M.-L.; Zhu, Z. H.; Sun, Y.; Wang, X.-L.; Fu, Y.-B. Acta Phys. Sin. 1999, 48, 2222. (39) La Macchia, G.; Infante, I.; Raab, J.; Gibson, J. K.; Gagliardi, L. Phys. Chem. Chem. Phys. 2008, 10, 7278. (40) Chen, J.; Meng, D. Q.; Du, J. G.; Jiang, G.; Gao, T.; Zhu, Z. H. Acta Phys. Sin. 2010, 59, 1658. (41) Li, Q.; Liu, X. Y.; Gao, T.; Zhu, Z. H.; Fu, Y.-B.; Wang, X. L.; Sun, Y. Acta Phys.-Chim. Sin. 2000, 16, 987. (42) Li, Q.; Wang, H.-Y.; Jiang, G.; Zhu, Z. H. Acta Phys.-Chim. Sin. 2001, 17, 622. (43) Kovacs, A.; Konings, R. J. M.; Raab, J.; Gagliardi, L. Phys. Chem. Chem. Phys. 2008, 10, 1114. (44) Dyall, K. G. Mol. Phys. 1999, 96, 511. (45) Wang, H.-Y.; Gao, T.; Yi, Y. G.; Tan, M. L.; Zhu, Z. H.; Fu, Y. B.; Wang, X. L.; Sun, Y. Acta Phys. Sin. 1999, 48, 2215. (46) Gagliardi, L.; Roos, B. O.; Malmqvist, P.-A.; Dyke, J. M. J. Phys. Chem. A 2001, 105, 10602. (47) Majumdar, D.; Balasubramanian, K.; Nitsche, H. Chem. Phys. Lett. 2002, 361, 143. (48) Li, J.; Bursten, B. E.; Andrews, L.; Marsden, C. J. J. Am. Chem. Soc. 2004, 126, 3424. (49) Liao, M.-S.; Kar, T.; Scheiner, S. J. Phys. Chem. A 2004, 108, 3056. (50) Gagliardi, L.; Roos, B. O. Chem. Phys. Lett. 2000, 331, 229. (51) Ismail, N.; Heully, J.-L.; Saue, T.; Daudey, J.-P.; Marsden, C. J. Chem. Phys. Lett. 1999, 300, 296. (52) De Jong, W. A.; Visscher, L.; Nieuwpoort, W. C. J. Mol. Struct. (Theochem) 1999, 458, 41. (53) Gao, T.; Wang, H.-Y.; Huang, Z.; Zhu, Z. H.; Sun, Y.; Wang, X. L.; Fu, Y. B. Acta Phys.-Chim. Sin. 1999, 15, 1082. (54) Archibong, E. F.; Ray, A. K. J. Mol. Struct. (Theochem) 2000, 530, 165. (55) Maron, L.; Leininger, T.; Schimmelpfennig, B.; Vallet, V.; Heully, J.-L.; Teichteil, C.; Gropen, O.; Wahlgren, U. Chem. Phys. 1999, 244, 195. (56) Clavaguera-Sarrio, C.; Brenner, V.; Hoyau, S.; Marsden, C. J.; Millie, P.; Dognon, J.-P. J. Phys. Chem. B 2003, 107, 3051. (57) Infante, I.; Kovacs, A.; La Macchia, G.; Shahi, A. R. M.; Gibson, J. K.; Gagliardi, L. J. Phys. Chem. A 2010, 114, 6007.

ARTICLE

(58) Karlstr€om, G.; Lindh, R.; Malmqvist, P.-Å.; Roos, B. O.; Ryde, U.; Veryazov, V.; Widmark, P.-O.; Cossi, M.; Schimmelpfennig, B.; Neogrady, P.; Seijo, L. Comput. Mater. Sci. 2003, 287, 222. (59) http://www.teokem.lu.se/molcas/documentation/manual/. (60) Aquilante, F.; Vico, L. D.; Ferre, N.; Ghigo, G.; Malmqvist, P.Å.; Neogrady, P.; Pedersen, T. B.; Pitonak, M.; Reiher, M.; Roos, B. O.; Serrano-Andres, L.; Urban, M.; Veryazov, V.; Lindh, R. J. Comput. Chem. 2010, 31, 224. (61) Roos, B. O. In Advances in Chemical Physics, Ab Initio Methods in Quantum Chemistry - II; Lawley, K. P., Ed.; John Wiley & Sons Ltd.: Chichester, 1987; Chapter 69; p 399. (62) Andersson, K.; Malmqvist, P.-Å.; Roos, B. O.; Sadlej, A.; Wolinski, K. J. Phys. Chem. 1990, 94, 5483. (63) Andersson, K.; Malmqvist, P.-Å.; Roos, B. O. J. Chem. Phys. 1992, 96, 1218. (64) Douglas, N.; Kroll, N. M. Ann. Phys. 1974, 82, 89. (65) Hess, B. Phys. Rev. A 1986, 33, 3742. (66) Roos, B. O.; Lindh, R.; Malmqvist, P. Å.; Veryazov, V.; Widmark, P.-O. Chem. Phys. Lett. 2005, 409, 295. (67) Roos, B. O.; Lindh, R.; Malmqvist, P.-Å.; Veryazov, V.; Widmark, P.-O. J. Phys. Chem. A 2004, 108, 2851. (68) Roos, B. O.; Malmqvist, P.-Å. Phys. Chem. Chem. Phys. 2004, 6, 2919. (69) 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 D.01; Gaussian, Inc.: Wallingford CT, 2004. (70) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (71) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1988, 37, 785. (72) Perdew, J. P. Phys. Rev. B 1986, 33, 8822. (73) Adamo, C.; Barone, V. J. Chem. Phys. 1998, 108, 664. (74) Tao, J. M.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Phys. Rev. Lett. 2003, 91, 146401. (75) Adamo, C.; Barone, V. J. Chem. Phys. 1999, 110, 6158. (76) Becke, A. D. Phys. Rev. A 1988, 38, 3098. (77) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865. (78) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1997, 78, 1396. (79) Hamprecht, F. A.; Cohen, A.; Tozer, D. J.; Handy, N. C. J. Chem. Phys. 1998, 109, 6264. (80) Cao, X. Y.; Dolg, M.; Stoll, H. J. Chem. Phys. 2003, 118, 487. (81) Dunning, J., T. H. J. Chem. Phys. 1989, 90, 1007. (82) Li, Q.; Xu, C. Sichuan Shifan Daxue Xuebao, Ziran Kexueban 2000, 23, 386. (83) Averkiev, B.; Mantina, M.; Valero, R.; Infante, I.; Kovacs, A.; Truhlar, D. G.; Gagliardi, L. Theor. Chem. Acc. 2011in press. (84) Notter, F.-P.; Dubillard, S.; Bolvin, H. J. Chem. Phys. 2008, 128, 164315. (85) Infante, I.; Eliav, E.; Vilkas, M. J.; Ishikawa, Y.; Kaldor, U.; Visscher, L. J. Chem. Phys. 2007, 127, 124308. (86) Infante, I.; Andrews, L.; Wang, X.; Gagliardi, L. Chem.—Eur. J. 2010, 16, 12804. 6656

dx.doi.org/10.1021/jp202538k |J. Phys. Chem. A 2011, 115, 6646–6656