Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Cundari Th.R. -- Computational Organometallic Chemistry-0824704789

.pdf
Скачиваний:
76
Добавлен:
08.01.2014
Размер:
4.83 Mб
Скачать

368 Li and Bursten

calculation of vibrational properties requires full geometry optimizations to a high degree of accuracy, the PW91 DFT method is anticipated to be an appropriate theoretical tool, especially for metal-containing systems. We have recently shown that this method can be used with good success to calculate the vibrational modes of Pa(Cot)2 (15). Here we will extend these studies to the calculations and assignments of the vibrational spectra to the other An(Cot)2 (An Th–Am) systems, with comparisons to the available experimental data. The present work represents the first comprehensive theoretical study of the vibrational spectra and assignments of the actinocenes.

Based on standard group theoretical analysis (103), among the 93 vibrational modes of actinocenes with D8h symmetry, only the 4 A2u and 6 E1u modes are IR active, whereas 4 A1g, 5 E1g, and 6 E2g modes are Raman active. Table 6 lists the calculated LDA and PW91 frequencies, infrared intensities, and assignments for the vibrational modes of Th(Cot)2. The vibrational frequencies and IR intensities for the other actinocenes are listed in Table 7.

The simple linear-three-mass model for XYX systems (104) can be applied to the An(Cot)2 sandwich complexes when treating the Cot rings as rigid mass points. Using this model, three vibrational modes involving An–Cot interactions are apparent: (1) the symmetric ring–metal–ring stretching mode (A1g), (2) the asymmetric ring–metal–ring stretching mode (A2u), and (3) the ring–metal–ring bending mode (E1u). These three modes represent the lowest-frequency Raman and IR vibrations of actinocenes. As expected, the symmetric ring–metal–ring stretching causes a large change in the molecular polarizability, leading to a strong Raman absorption, observed at 225 cm 1 for Th(Cot)2. The asymmetric ring–metal–ring stretching mode of Th(Cot)2 is observed as a strong IR absorption at 250 cm 1. For the ring–metal–ring bending mode, which has not been observed experimentally for any actinocene, the potential energy surface is so flat that the calculated frequency of this mode becomes imaginary even when a very stringent numerical integration accuracy (INTEGRATION 8.0) is employed. By increasing the accuracy of the numerical integration even more (to INTEGRATION 10.0), we have obtained this frequency as 57 cm 1 for Th(Cot)2. Based on the experimental estimation, this band should appear around 125 cm 1 (102).

The C8H8 rings are, of course, neither structureless point masses nor rigid rings. In addition to the modes involving the metal–ring interactions (for which the rings are largely rigid), there are numerous modes corresponding to C–C and C–H stretching and in-plane ( ) and out-of-plane ( ) C–H bending. We will not detail the descriptions of these modes but only summarize the results in Table 6. We will examine more closely the ring-centered modes that lead to experimentally observed IR and Raman bands with strong intensities.

In addition to the previously mentioned strong absorption at 250 cm 1, the IR spectrum of Th(Cot)2 exhibits strong absorptions at 695, 742, and 895 cm 1.

TABLE 6

Calculated LDA and PW91 DFT Vibrational Frequencies (cm

1

) and Absolute IR Intensities (km/mol, in

 

parenthesis) for the IR and Raman Active Modes of Th(Cot)

 

 

 

 

 

 

 

 

 

 

 

 

2

 

 

 

 

Mode

Exptl.

a

LDA

PW91

 

Assignment

b

 

 

 

 

 

 

 

 

 

 

 

 

 

A

2u

250

(s)

246

(72.7)

220

(76.9)

 

Asymmetric ring–metal–ring stretching

 

 

 

695

(vs)

666

(457)

633

(446)

 

Asymmetric C–H bending ( )

 

 

 

742

(s)

762

(17.2)

747

(0.7)

 

Asymmetric in-plane C–C stretching

 

 

 

3005

(m)

3059

(0.4)

3079

(0.1)

 

Asymmetric C–H stretching

E

1u

(125)

 

 

8

(7.7)

57

(8.9)

 

Asymmetric ring–metal–ring bending

 

 

 

 

 

 

225

(0.0)

235

(0.2)

 

Asymmetric in-plane C–C stretching

 

 

 

 

 

 

743

(35.1)

745

(23.2)

 

Asymmetric C–H bending ( )

 

 

 

895

(s)

915

(69.4)

895

(84.1)

 

Asymmetric ring–metal–ring tilting

 

 

 

1315

(m)

1406

(1.5)

1430

(0.2)

 

Asymmetric C–H bending ( )

 

 

 

2920

(m)

3052

(0.1)

3072

(8.3)

 

Asymmetric C–H stretching

A

1g

225

(s)

227

 

138

 

 

Symmetric ring–metal–ring stretching

 

 

 

 

 

 

680

 

683

 

 

Symmetric C–H bending ( )

 

 

 

 

 

 

770

 

809

 

 

Symmetric in-plane C–C stretching

 

 

 

3045

(m)

3059

 

3058

 

 

Symmetric C–H stretching

E

1g

242

(s)

265

 

260

 

 

Symmetric ring–metal–ring tilting

 

 

 

 

 

 

768

 

762

 

 

Symmetric C–H bending ( )

 

 

 

 

 

 

911

 

898

 

 

Symmetric C–C stretching

 

 

 

3022

 

 

1405

 

1426

 

 

Symmetric C–H bending ( )

 

 

 

(m)

3052

 

3068

 

 

Symmetric C–H stretching

E

2g

 

 

 

237

 

235

 

 

Symmetric CCC bending

( )

 

 

 

391

(m)

388

 

365

 

 

Symmetric in-plane C–C stretching

 

 

 

775

(s)

817

 

819

 

 

Symmetric C–H bending ( )

 

 

 

 

 

 

1187

 

1177

 

 

Symmetric C–H bending ( )

 

 

 

 

 

 

1498

 

1500

 

 

Symmetric in-plane C–C stretching

 

 

 

2905

(m)

3038

 

3066

 

 

Symmetric C–H stretching

 

 

 

 

 

 

 

 

 

 

 

 

 

a

 

 

 

 

 

 

 

 

 

 

 

 

The experimental intensities are shown as very strong (vs), strong (s), and medium (m) based on Ref. 100.

 

b

 

 

 

 

 

 

 

 

 

 

 

 

The and are referred to the Cot plane.

 

 

 

 

 

 

Complexes Organoactinide and DFT Relativistic

369

370

Li and Bursten

TABLE 7 PW91 Vibrational Frequencies (cm 1) and Absorption Intensities (km/mol) for IR and Raman Active Modes of An(Cot)2 (An Pa–Am)

Mode

Pa(Cot)2

U(Cot)2

Np(Cot)2

Pu(Cot)2

Am(Cot)2

A2u

239

232

216

203

195

 

683

735

741

734

722

 

771

751

748

741

746

 

3065

3072

3073

3076

3072

E1u

38

36

45

67

72

 

220

236

235

235

235

 

761

760

759

757

755

 

895

908

910

907

909

 

1438

1438

1438

1438

1439

 

3064

3066

3066

3070

3064

A1g

328

385

252

192

87

 

530

586

700

705

699

 

839

756

716

708

765

 

3090

3133

3125

3125

3072

E1g

253

233

221

221

216

 

787

783

780

778

772

 

902

902

903

901

903

 

1427

1430

1432

1433

1435

 

3063

3063

3065

3067

3065

E2g

219

143

235

250

253

 

410

404

380

376

371

 

823

829

828

831

824

 

1161

1171

1174

1173

1176

 

1513

1516

1518

1516

1516

 

3046

3040

3041

3045

3051

 

 

 

 

 

 

All of these vibrational frequencies are reproduced with reasonable accuracy in the DFT calculations. The 695-cm 1 band is the strongest in the spectrum. The calculated frequencies, 666 cm 1 at the LDA level and 633 cm 1 at the PW91 level, are only in fair agreement with the experimental value. However, the mode is calculated to have the strongest absolute intensity, which is in complete accord with the experimental observation. We are therefore confident in assigning this mode as asymmetric C–H bending, which differs from a previous assignment of it as the asymmetric ring–metal–ring tilting mode (102). The 742-cm 1 and 895cm 1 modes are reproduced with remarkable accuracy at the PW91 level (Table 6). These bands are due to C–C stretching and asymmetric ring–metal–ring tilting, respectively.

Relativistic DFT and Organoactinide Complexes

371

In addition to the strong band at 225 cm 1, the Raman spectrum of Th(Cot)2 shows two other strong absorptions at 242 and 775 cm 1, which are assigned as symmetric ring–metal–ring tilting and a C–C bending. These vibrational frequencies of these bands have been reproduced with reasonable accuracy at both the LDA and PW91 levels of calculation.

Table 6 demonstrates that the PW91 frequencies and IR intensities are overall in fair agreement with the experimental spectra. These theoretical results can be further improved by increasing the quality of the optimized geometry via the use of larger basis sets. By combining the calculated IR intensities and the frequencies, the PW91 theoretical calculations can provide reliable predictions of the positions and absorbency of the IR transitions. Therefore, theoretical calculations of IR spectra can be very useful in helping experimentalists to identify the IR bands and their microscopic origins, even for molecules as large as the actinocenes.

Because the geometric structures and bonding are very similar for the entire series of actinocenes, we might expect their vibrational frequencies to be very similar as well. Table 7 lists the PW91 calculated frequencies for the An(Cot)2 complexes, with An Pa–Am. We see a smooth trend in the vibrational frequencies across the series, with similar values to those discussed earlier for Th(Cot)2. The ability to calculate the predicted changes in the vibrational frequencies across a series of homologous complexes is certainly a strength of the application of DFT to vibrational problems.

From Tables 6 and 7, we see that the lowest A1g vibration (i.e., the symmetric Cot–An–Cot stretch) and the lowest A2u vibration (the asymmetric Cot–An– Cot stretch) show nearly the same trend across the actinide series: the frequencies increase slightly from Th to Pa and then decrease from U to Am. This trend is entirely consistent with the trends in the group orbital overlap integrals mentioned earlier and the An–C8H8 bond energies. Because the frequencies of fully symmetric vibrations are rigorously independent of the mass of An, the changes of the vibrational frequencies of the A1g mode from Th to Am directly reflects the An– Cot bonding strength. For the asymmetric Cot–An–Cot stretching mode, the frequency changes from Th to Am agree well with the increase of the atomic mass from Th to Am and the decrease of the symmetric stretching frequencies.

In conclusion, the calculated PW91 vibrational frequencies are in reasonably good agreement with the experimental data. As we discussed earlier, augmentation of the basis sets can improve the molecular geometries of the actinocenes. We can therefore expect that better agreement with experiments can be achieved if the vibrational frequencies are calculated with larger basis sets. The assignments of the IR and Raman vibrational modes will help us to understand the vibrational spectra of other actinocenes and their microscopic origins. Our calculations in this work indicate that gradient-corrected density functional meth-

372

Li and Bursten

ods, especially the PW91 functional, not only can be used to elucidate the bonding and electronic structure, but can also be of great value in interpreting and predicting the vibrational properties of actinide complexes. Because the symmetric vibrational frequencies are not sensitive to the mass of the central metal, the present vibrational analysis and mode assignments will also be useful in understanding the vibrational spectra of lanthanocenes (105).

7. CONCLUSIONS

Relativistic density functional theory, especially with the inclusion of nonlocal exchange and correlation corrections, has become a powerful predictive tool in actinide chemistry. The methodology is sufficiently efficient to allow experimentally important properties, such as the geometry, vibrational frequencies, and infrared absorption intensities, to be calculated even for large organoactinide systems such as those discussed here. Inasmuch as many aspects of actinide chemistry are experimentally challenging because of the difficulty in handling of the elements, reliable theoretical calculations provide a valuable adjunct to experimental studies.

State-of-the-art DFT methods can provide theoretical interpretations of experimental results and are becoming more and more reliable in predicting physicochemical properties of actinide compounds. In spite of the current shortcomings of the method, new developments and future advances in density functional theory, such as the hybrid exchange-correlation functionals (106), meta-GGA functionals with high-order gradient corrections and kinetic energy density included (107), and time-dependent DFT (108), promise to provide even greater utility with respect to the study of ground-state and excited-state properties of actinide and organoactinide complexes (109).

ACKNOWLEDGMENTS

We gratefully acknowledge support for this research from the Division of Chemical Sciences, U.S. Department of Energy (Grant DE-FG02-86ER13529 to BEB), from Los Alamos National Laboratory, and from the Ohio Supercomputer Center and the Environmental Molecular Sciences Laboratory at Pacific Northwest National Laboratory for grants of computer time.

ABBREVIATIONS

An

actinide element

BLYP

Becke 1988–Lee-Yang-Parr

BP86

Becke 1988–Perdew 1986

Bz

benzene, C6H6

Relativistic DFT and Organoactinide Complexes

373

Cht

cycloheptatrienyl, C7H7

 

Cot

cyclooctatetraene, C8H8

 

Cp

cyclopentadienyl, C5H5

 

Cp*

pentamethylcyclopentadienyl, C5Me5

 

DFT

density functional theory

 

DZP

double-ζ basis set with polarization function

 

GGA

generalized gradient approach

 

HFS

Hartree–Fock–Slater

 

LDA

local density approach

 

NR

nonrelativistic

 

PW91

Perdew–Wang functional (1991)

 

RKS

spin-restricted Kohn–Sham

 

SO

spin-orbit

 

SOCI

spin-orbit configuration interaction

 

SR

scalar relativistic

 

TBB

1,3,5-tri-tert-butylbenzene, C6H3-tBu3

 

TZ2P

triple-ζ basis set with two polarization functions

 

UKS

spin-unrestricted Kohn–Sham

 

NOTES AND REFERENCES

1.(a) TJ Kealy, PL Pauson. Nature 168:1039–1040, 1951. (b) SA Miller, JA Tebboth, JF Tremaine. J Chem Soc 632–635, 1952. (c) EO Fischer, W Phab. Z Naturforsch 76:377–379, 1952. (d) G Wilkinson, M Rosenblum, MC Whiting, RB Woodward. J Am Chem Soc 74:2125–2126, 1952.

2.RH Crabtree. The Organometallic Chemistry of the Transition Metals. 2nd ed. New York: Wiley, 1994.

3.For cyclopolyene and cyclopolyenyl complexes of transition metals, see: (a) G Deganello. Transition Metal Complexes of Cyclic Polyolefins. London: Academic Press, 1979, Chap 1. (b) MLH Green, DKP Ng. Chem Rev 95:439–473, 1995.

4.For recent studies of f-element C6H6, C7H7, and C8H8 sandwich complexes, see, for example: (a) WA King, S Di Bella, G Lanza, K Khan, DJ Duncalf, FGN Cloke, IL Fragala, TJ Marks. J Am Chem Soc 118:627–635, 1996. (b) D Gourier, D Caurant, T Arliguie, Ephritikhine. J Am Chem Soc 120:6084–6092, 1998. (c) M Dolg, P Fulde. Eur J Chem 4:200–204, 1998.

5.(a) H Hoberg, R Krause-Go¨ing, R Mynott. Angew Chem Int Ed Engl 17:123–124, 1978. (b) H Hoberg, C Fro¨hlich. J Organomet Chem 168:C52–C52, 1979.

6.H Hoberg, C Fro¨hlich. J Organomet Chem 197:105–109, 1980.

7.EO Fischer, W Hafner. Z Naturforsch B10:665–667, 1955.

8.U Mu¨ller-Westerhoff, A Streitwieser Jr. J Am Chem Soc 90:7364–7364, 1968.

9.T Arliguie, M Lance, M Nierlich, J Vigner, M Ephritikhine. J Chem Soc Chem Commun 183–184, 1995.

10.(a) Z Liang, AG Marshall, A Pires de Matos, JD Spirlet. In: LR Morss, J Fuger, eds. Transuranium Elements: A Half Century. Washington, DC: American Chemical

374

Li and Bursten

Society, 1992, pp 247–250. (b) WW Yin, AG Marshall, J Marcalo, A Pires de Matos. J Am Chem Soc 116:8666–8672, 1994. (c) J Marcalo, JP Leal, A Pires de Matos, AG Marshall. Organometallics 16:4581–4588, 1997.

11.A large number of review papers and books about ferrocenes have been published. See, for example: (a) A Togni, T Hayashi, eds. Ferrocenes. Weinheim, Germany: VCH, 1995. (b) WP McNutt, CU Pittman Jr. Ferrocene and Ferrocene Derivatives. Redstone Arsenal, Redstone Scientific Information Center, 1966.

12.(a) M Pepper, BE Bursten. Chem Rev 91:719–741, 1991. (b) G Schreckenbach, PJ Hay, RL Martin. J Comp Chem 20:70–90, 1999.

13.J Li, BE Bursten. J Am Chem Soc 121:10243–10244, 1999.

14.J Li, BE Bursten. J Am Chem Soc 119:9021–9032, 1997.

15.J Li, Be Bursten. J Am Chem Soc 120:11456–11466, 1998.

16.A numerical relativistic Dirac–Fock calculation for the uranium atom indicates that

˚ ˚

the spin-orbit averaged radii of the maximum radial density are 0.72 A, 0.86 A,

˚˚

0.56A, and 1.30 A, respectively, for 6s, 6p, 5f , and 6d orbitals. See: JP Desclaux. Atomic Data and Nuclear Data Tables 12:311–406, 1973.

17.PJ Hay, WR Wadt, LR Kahn, RC Raffenetti, DH Phillips. J Chem Phys 71:1767– 1779, 1979.

18.For a general discussion of dynamic and nondynamic electron correlation, see, for example: (a) DKW Mok, R Neumann, NC Handy. J Phys Chem 100:6225–6230, 1996. (b) MA Buijse, EJ Baerends. In: ED Ellis, ed. Density Functional Theory of Molecules, Clusters, and Solids. Dordrecht, The Netherlands: Kluwer Academic, 1995, pp 1–46.

19.For qualitative and quantitative discussions of scalar and spin-orbit relativistic effects, see, for example: (a) KS Pitzer. Acc Chem Res 12:271–275, 1979. (b) P Pyykko¨, JP Desclaux. Acc Chem Res 12:276–281, 1979. (c) P Pyykko¨. Chem Rev 88:563–594, 1988. (d) WHE Schwarz, EM van Wezenbeek, EJ Baerends, JG Snijders. J Phys B22:1515–1530, 1989.

20.J Li, BE Bursten. Unpublished results.

21.CJ Burns, BE Bursten. Comments Inorg Chem 9:61–69, 1989.

22.ER Davidson. Chem Rev 100:351–352, 2000, and references therein.

23.BE Bursten, RJ Strittmatter. Angew Chem Int Ed Engl 30:1069–1085, 1991.

24.WF Schneider, RJ Strittmatter, BE Bursten, DE Ellis. In: JK Labanowski, JW Andzelm, eds. Density Functional Methods in Chemistry. New York: Springer-Verlag, 1991, pp 247–260.

25.(a) RG Parr, W Yang. Density Functional Theory of Atoms and Molecules. New York: Oxford University Press, 1989. (b) RM Dreizler, EKU Gross. Density Functional Theory. Berlin: Springer-Verlag, 1990. (c) ES Kryachko, EV Ludena. Energy Density Functional Theory of Many-Electron Systems. Dordrecht, The Netherlands Kluwer, 1990. (d) JK Labanowski, JW Andzelm, eds. Density Functional Methods in Chemistry. New York: Springer-Verlag, 1991.

26.(a) EKU Gross, RM Dreizler. Density Functional Theory. New York: Plenum Press, 1995. (b) JM Seminario, P Politzer, eds. Modern Density Functional Theory: A Tool for Chemistry. Amsterdam: Elsevier, 1995. (c) RF Nalewajski, ed. Density Functional Theory. In: Topics in Current Chemistry. Vols 180–183. Berlin:

Relativistic DFT and Organoactinide Complexes

375

Springer-Verlag, 1996. (d) BB Laird, RB Ross, T Ziegler. Chemical Applications of Density-Functional Theory. In: ACS Symposium Series, Vol 629. Washington, DC: American Chemical Society, 1996. (e) JF Dobson, G Vignale, MP Das, eds. Electronic Density Functional Theory: Recent Progress and New Directions. New York: Plenum Press, 1998.

27.(a) T Ziegler. Chem Rev 91:651–667, 1991. (b) SB Trickey. In: ES Kryachko, JL Calais, eds. Conceptual Trends in Quantum Chemistry. Dordrecht, The Netherlands: Kluwer Academic, 1994, pp. 87–100. (c) RG Parr, W Yang. Annu Rev Phys Chem 46:701–728, 1995. (d) W Kohn, AD Becke, RG Parr. J Phys Chem 100: 12974–12980, 1996. (e) A Nagy. Phys Reports 298:1–79, 1998.

28.(a) J Li, G Schreckenbach, T Ziegler. J Am Chem Soc 117:486–494, 1995. (b) L Fan, T Ziegler. In: DE Ellis, ed. Density Functional Theory of Molecules, Clusters, and Solids. Dordrecht, The Netherlands: Kluwer Academic, 1995. (c) TV Russo, RL Martin, PJ Hay. J Chem Phys 102:8023–8028, 1995. (d) AP Scott, L Radom. J Phys Chem 100:16502–16513, 1996.

29.For discussions of the performance of DFT methods, see, for example: (a) BG Johnson, PMW Gill, JA Pople. J Chem Phys 98:5612–5626, 1993. (b) B Delley, M Wrinn, HP Lu¨thi. J Chem Phys 100:5785–5791, 1994. (c) SG Wang, DK Pan, WHE Schwarz. J Chem Phys 102:9296–9308, 1995. (d) CW Bauschlicher Jr. Chem Phys Lett 246:40–44, 1995. (e) M Kaupp, OL Malkina, VG Malkin. J Chem Phys 106:9201–9212, 1997. (f) M Kaupp. Eur J Chem 4:2059–2071, 1998. (g) ON Ventura, M Kieninger, RE Cachau. J Phys Chem A 103:147–151, 1999.

30.For lighter molecules, see: AD Becke. J Chem Phys 96:2155–2160, 1992. Good accuracy can also be reached for heavier systems, such as transition metals and lanthanides. See, for example: (a) A Ricca, CW Bauschlicher Jr. Theor Chim Acta 92:123–131, 1995. (b) LA Eriksson, LGM Pettersson, PEM Siegbahn, U Wahlgren. J Chem Phys 102:872–878, 1995. (c) SG Wang, WHE Schwarz. J Phys Chem 99: 11687–11695, 1995.

31.BG Johnson, PMW Gill, JA Pople. J Chem Phys 97:7846–7848, 1992.

32.MC Holthausen, C Heinemann, HH Cornehl, W Koch, H Schwarz. J Chem Phys 102:4931–4940, 1995.

33.PRT Schipper, OV Gritsenko, EJ Baerends. J Phys Chem 111:4056–4067, 1999.

34.AD Becke. In: DR Yarkony, ed. Modern Electronic Structure Theory. Singapore: World Scientific, 1995, pp 1022–1046.

35.For the meaning of KS orbitals and eigenvalues, see: (a) EJ Baerends, OV Gritsenko. J Phys Chem A 101:5383–5403, 1997. (b) R Stowasser, R Hoffmann. J Am Chem Soc 121:3414–3420, 1999.

36.We have found a good linear correlation between the calculated ionization energies and the orbital energies of Pa(Cot)2, IEi 2.106 0.990 i; see Ref. (15).

37.It has been discovered that excitations from the highest occupied orbital are in agreement with differences of Kohn–Sham eigenvalues. See: CJ Umrigar, A Savin, X Gonze. In: JF Dobson, G Vignale, MP Das, eds. Electronic Density Functional Theory: Recent Progress and New Directions. New York: Plenum Press, 1998, pp 167–176.

38.P Politzer, F Abu-Awwad. Theor Chem Acc 99:83–87, 1998.

376

Li and Bursten

39.See, for example: CJ Cramer, FJ Dulles, DJ Giesen, J Almlo¨f. Chem Phys Lett 245:165–170, 1995, and references cited therein.

40.For examinations of basis set effects in DFT calculations, see: (a) KS Raymond, RA Wheeler. J Comp Chem 20:207–216, 1999. (b) AC Scheiner, J Baker, JW Andzelm. J Comp Chem 18:776–795, 1997.

41.(a) EJ Baerends, DE Ellis, P Ros. Chem Phys 2:41–51, 1973. (b) EJ Baerends, P Ros. Chem Phys 2:52–59, 1973. (c) EJ Baerends, P Ros. Int J Quantum Chem Quantum Chem Symp 12:169–190, 1978.

42.(a) JG Snijders, EJ Baerends. J Mol Phys 36:1789–1804, 1978. (b) JG Snijders, EJ Baerends, P Ros. Mol Phys 38:1909–1929, 1979.

43.W Ravenek. In: HJJ te Riele, ThJ DeDekker, HA van de Vorst, eds. Algorithms and Applications on Vector and Parallel Computers. Amsterdam: Elsevier, 1987.

44.C Fonseca Guerra, O Visser, JG Snijders, G te Velde, E J Baerends. In: E Clementi, G Corongiu, eds. Methods and Techniques for Computational Chemistry. Cagliari, Italy: STEF, 1995, p 305.

45.(a) PM Boerrigter, G te Velde, E J Baerends. Int J Quantum Chem 33:87–113, 1988. (b) G te Velde, EJ Baerends. J Comput Phys 99:84–98, 1992.

46.For discussions of the fractional occupation number approach, see: (a) SG Wang, WHE Schwarz. J Chem Phys 105:4641–4648, 1996. (b) FW Averill, GS Painter. Phys Rev B46:2498–2502, 1992. (c) BI Dunlap. In: KP Lawlay, ed. Ab initio Methods in Quantum Chemistry II. New York: Wiley, 1987. (d) JF Janak Phys Rev B18:7165–7168, 1978. (e) JC Slater, JB Mann, TM Wilson, JH Wood. Phys Rev 184:672–694, 1969.

47.(a) JC Slater. Quantum Theory of Molecular and Solids. Vol 4. New York: McGraw-Hill, 1974. (b) SH Vosko, L Wilk, M Nusair. Can J Phys 58:1200–1211, 1980.

48.(a) JP Perdew, Y Wang. Phys Rev B45:13244–13249, 1992. (b) JP Perdew, JA Chevary, SH Vosko, KA Jackson, MR Pederson, DJ Singh, C Foilhais. Phys Rev B46:6671–6687, 1992.

49.The HFS functional uses Slater’s Xα(α 0.7) pure-exchange electron gas formulation with no correlation functional.

50.BP86 exchange functional: AD Becke. Phys Rev A38:3098–3100, 1988. BP86 correlation functional: (a) JP Perdew. Phys Rev B33:8822–8824, 1986. (b) JP Perdew. Phys Rev B34:7406–7406 (erratum), 1986.

51.BLYP correlation functional: C Lee, W Yang, RG Parr. Phys Rev B37:785–789, 1988.

52.The fully relativistic Dirac–Slater atomic code, DIRAC, is one of the auxiliary programs distributed with the ADF code.

53.J Krijn, EJ Baerends. Fit Functions in the HFS Method (Internal Report). Free University of Amsterdam, 1984.

54.(a) JG Snijders, EJ Baerends, P Vernooijs. At Nucl Data Tables 26:483–509, 1981.

(b) P Vernooijs, GJ Snijders, EJ Baerends. Slater-Type Basis Functions for the Whole Periodic System (Internal Report). Free University of Amsterdam, 1984.

55.T Ziegler, EJ Baerends, JG Snijders, W Ravenek. J Phys Chem 93:3050–3056, 1989.

56.For definitions of the mass-velocity, Darwin, and spin-orbit coupling terms, see:

Relativistic DFT and Organoactinide Complexes

377

JG Snijders, EJ Baerends. In: P Coppens, MB Hall, eds. Electron Distribution and the Chemical Bond. New York: Plenum Press, 1981.

57.J Li, BE Bursten. Unpublished results.

58.For exchange-correlation functionals beyond first-order gradient-correction, see, for example: HL Schmider, AD Becke. J Chem Phys 109:8188–8199, 1998.

59.A Avdeef, KN Raymond, KO Hodgson, A Zalkin. Inorg Chem 11:1083–1088, 1972.

60.DJA De Ridder, J Rebizant, C Apostolidis, B Kanellakopulos, E Dornberger. Acta Cryst C52:597–600, 1996.

61.AG Orpen, L Brammer, FH Allen, O Kennard, DG Watson, R Taylor. J Chem Soc Dalton Trans S1–S83, 1989 (Suppl).

62.For a theoretical analysis of the advantages of the PW91 functional, see: (a) JP Perdew, K Burke. Int J Quantum Chem 57:309–319, 1996. (b) K Burke, JP Perdew, M Ernzerhof. Int J Quantum Chem 61:287–293, 1997.

63.J Li, BE Bursten. Relativistic Density Functional and Ab initio Theoretical Studies of Actinide Fluorides. Book of Abstracts, 217th National ACS Meeting, Anaheim, CA, 1999, NUCL 174.

64.(a) MF Zhou, L Andrews, J Li, BE Bursten. J Am Chem Soc 121:9712–9721, 1999. (b) MF Zhou, L Andrews, J Li, BE Bursten. J Am Chem Soc 121:12188– 12189, 1999. (c) L Andrews, B Liang, J Li, BE Bursten. Angew Chem Int Ed. In press.

65.RD Fischer. Theor Chim Acta (Berlin) 1:418–431, 1963.

66.(a) A Streitwieser Jr, SA Kinsley. NATO ASI Ser (Ser C). 155:77–114, 1985, and references cited therein. (b) A Streitwieser Jr. Inorg Chim Acta 94:171–177, 1984.

67.For recent reviews, see: (a) P Pyykko¨. Inorg Chim Acta 139:243–245, 1987. (b) K Balasubramanian. In: KA Gschneider Jr, L Eyring, eds. Handbook on the Physics and Chemistry of Rare Earths. Vol 18. Amsterdam: North-Holland, 1994, pp 29– 158. (c) M Dolg, H Stoll. In: KA Gschneider Jr, L Eyring, eds. Handbook on the Physics and Chemistry of Rare Earths. Vol 22. Amsterdam: Elsevier Science, 1996, pp 607–729.

68.(a) KD Warren. Struct Bond 33:97–138, 1975. (b) DW Clack, KD Warren. J Organomet Chem 122:C28–C30, 1976.

69.(a) P Pyykko¨, LL Lohr. Inorg Chem 20:1950–1959, 1981. (b) P Pyykko¨, LJ Laaksonen, K Tatsumi. Inorg Chem 28:1801–1805, 1989.

70.RB King. Inorg Chem 31:1978–1980, 1992.

71.MG Cory, S Ko¨stlmeier, M Kotzian, N Ro¨sch, MC Zerner. J Chem Phys 100: 1353–1365, 1994.

72.(a) M Dolg, P Fulde, H Stoll, H Preuss, AHH Chang, RM Pitzer. Chem Phys 195: 71–82, 1995. (b) W Liu, M Dolg, P Fulde. J Chem Phys 107:3584–3591, 1997.

(c) W Liu, M Dolg, P Fulde. Inorg Chem 37:1067–1072, 1998.

73.(a) AHH Chang, RM Pitzer. J Am Chem Soc 111:2500–2507, 1989. (b) AHH Chang, K Zhao, WC Ermler, RM Pitzer. J Alloys Comp 213/214: 191–195, 1994.

74.(a) N Ro¨sch, A Streitwieser Jr. J Organomet Chem 145:195–200, 1978. (b) N Ro¨sch, A Streitwieser Jr. J Am Chem Soc 105:7237–7240, 1983. (c) N Ro¨sch. Inorg Chim Acta 94:297–299, 1984.

75.PM Boerrigter, EJ Baerends, JG Snijders. Chem Phys 122:357–374, 1988.

Соседние файлы в предмете Химия