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

Cramer C.J. Essentials of Computational Chemistry Theories and Models

.pdf
Скачиваний:
355
Добавлен:
08.01.2014
Размер:
7.27 Mб
Скачать

8.6 GENERAL PERFORMANCE OVERVIEW OF DFT

291

atoms, Rienstra-Kiracofe et al. (2002) and Bauschlicher and Gutsev (2002) have separately noted that B3LYP with large basis sets seems to be particularly robust.

Finally, atomic and molecular proton affinities (PAs) have also been evaluated for various functionals for ammonia, water, acetylene, silane, phosphine, silylene, hydrochloric acid, and molecular hydrogen. For G2 and G3 theories, the mean unsigned error in PAs is 1.1 and 1.3 kcal mol−1, respectively. At the SVWN, BLYP, BP86, BPW91, B3LYP, B3P86, and B3PW91 levels (using the 6-311+G(3df,2p) basis set), the corresponding errors are 5.8, 1.8, 1.5, 1.5, 1.3, 1.1, and 1.2 kcal mol−1, respectively (quantitatively similar results have also been obtained with more modern functionals). The much cheaper hybrid DFT methods are thus entirely competitive with G2 and G3, although the data set is perhaps too small to come to any firm conclusions on this topic (cf. Pokon et al. 2001).

8.6.2Geometries

Analytic first derivatives are available for almost all density functionals, and as a result geometry optimization can be carried out with facility. The performance of the various functionals is usually quite good when it comes to predicting minimum energy structures. As summarized in Table 8.5, bond lengths at the LDA level for molecules composed of firstand second-row atoms are typically as good as those predicted from MP2 optimizations, with both these levels being somewhat improved over HF theory. The use of GGA functionals does not usually result in much improvement over the LDA level. However, the GGA functionals tend to systematically overestimate bond lengths. As noted in Section 6.4.2, the HF level tends to systematically under estimate bond lengths. Thus, it should come as no surprise that the hybrid ACM functionals, which mix the two, give noticeable improvement in predicted bond lengths (of course, the improvements are on the order of 0.005 A,˚ and it should be noted that most molecular properties are very little affected by such small variations in bond lengths). Very small improvements in geometrical accuracy are usually noted with increasing basis-set size beyond those listed in Table 8.5. Accuracies in bond angles for all flavors of DFT average about 1, the same as is found for HF and MP2. Similarly, the limited amount of data available for dihedral angles suggests that HF, MP2, and DFT all perform equivalently in this area.

Table 8.5 also indicates that the most popular functionals fail to be as accurate for molecules containing third-row main-group elements as they are for molecules made up of elements from the first two rows. The LYP correlation functional seems to perform particularly badly, while the PW91 functional is more robust.

It is, however, for the transition metals themselves that DFT has proven to be a tremendous improvement over HF and post-HF methods, particularly for cases where the metal atom is coordinatively unsaturated. The narrow separation between filled and empty d-block orbitals typically leads to enormous non-dynamical correlation problems with an HF treatment, and DFT is much less prone to analogous problems. Even in cases of a saturated coordination sphere, DFT methods typically significantly outperform HF or MP2. Jonas and Thiel (1995) used the BP86 functional to compute geometries for the neutral hexacarbonyl complexes of Cr, Mo, and W, the pentacarbonyl complexes of Fe, Ru, and Os, and the tetracarbonyl

292

8 DENSITY FUNCTIONAL THEORY

 

 

Table 8.5 Mean absolute errors in bond lengths for

 

 

 

 

 

˚

 

different methods over several different test sets (A)

 

Level of theory

 

 

Test setsa

 

 

 

 

A

B

C

 

 

 

 

 

MO theoretical methods

 

 

 

HF

0.022

0.021

 

 

MP2

 

0.014b

0.014

0.022

 

QCISD

 

0.013b

 

 

 

CCSD(T)

 

0.005b

 

 

 

LSDA functionals

 

 

 

 

 

SVWN

0.017

0.016

 

 

 

 

 

0.013d

 

 

GGA and MGGA functionals

 

 

 

BLYP

0.014

0.021

0.048

 

 

 

 

0.019c

 

 

 

 

 

0.022d

 

 

BP86

 

 

0.018d

 

 

BPW91

0.014

0.017

0.020

 

 

 

 

0.017d

 

 

HCTH

 

 

0.013c

 

 

 

 

 

0.014d

 

 

OLYP

 

 

0.018d

 

 

PBE

0.012

0.016d

 

 

PKZB

 

 

0.027d

 

 

PWPW91

0.012

0.014d

 

 

mPWPW91

0.012

 

 

 

TPSS

 

 

0.014d

 

 

VSXC

 

 

0.013d

 

 

Hybrid functionals

 

 

 

 

 

BH&HLYP

0.015

 

 

 

B1LYP

0.005

 

 

 

B1PW91

0.010

 

 

 

B97-1

 

 

0.008c

 

 

mPW1PW91

0.010

 

 

 

PBE1PBE

0.012

0.010d

 

 

TPSSh

 

 

0.010d

 

 

B3LYP

0.004

0.008c

0.030

 

 

 

 

0.010d

 

 

 

 

 

 

 

8.6 GENERAL PERFORMANCE OVERVIEW OF DFT

293

 

Table 8.5 (continued )

 

 

 

 

 

 

 

 

 

 

 

Level of theory

 

Test setsa

 

 

 

 

 

A

B

C

 

 

 

 

 

 

 

 

 

B3P86

 

0.008d

 

 

 

 

B3PW91

0.008

0.011

0.020

 

 

 

 

0.009d

 

 

 

 

mPW3PW91

0.008

 

 

 

 

 

 

 

 

 

a A: G2 subset (32 molecules containing only first-row

 

 

atoms, see Johnson, Gill, and Pople 1993), 6-311G(d,p)

 

 

basis set unless otherwise specified; B: (108 molecules

 

 

including firstand second-row atoms, see Scheiner, Baker,

 

 

and Andzelm 1997), 6-31G(d,p) basis set; C: (40 molecules

 

 

containing third-row atoms Ga-Kr, see Redfern, Blaudeau,

 

 

and Curtiss 1997).

 

 

 

 

 

 

b 6-31G(d,p) basis set.

 

 

 

 

 

 

c A 40-molecule subset with a polarized triple-ζ

basis set,

 

see Hamprecht et al. (1998).

d A 96-molecule set with the 6-311++G(3df,3pd) basis set, see Staroverov et al. (2003).

complexes of Ni, Pd, and Pt. Over the 10 unique metal–carbon bond lengths for which experimental data are available, they observed no error in excess of 0.01 A˚ except for W, where the error was 0.017 A˚ . At the HF and MP2 levels using equivalent basis sets, the corresponding average absolute errors are 0.086 and 0.028 A,˚ and the maximum deviations are 0.239 and 0.123 A˚ (Frenking et al. 1996).

To the extent DFT shows systematic weaknesses in geometries, it is in those areas where it similarly does poorly for energetics. Thus, van der Waals complexes tend to have interfragment distances that are too large because the dispersion-induced attraction is not properly modeled (although it may accidentally be mimicked by BSSE). Hydrogen bonds are somewhat too short as a rule, and indeed, most charge transfer complexes have their polarities overestimated so that they are too tightly bound. Finally, the tendency noted above in Section 8.5.6 for DFT to overdelocalize structures can show up in geometrical predictions. Thus, for instance, in 1,3-butadiene DFT tends to predict the formal single bond to be a bit too short and the formal double bonds to be somewhat too long (and this extends to other conjugated π systems). As already noted above in Section 8.5.6, this can also lead to a tendency to favor higher symmetry structures over ones of lower symmetry since the former tend to have more highly delocalized frontier orbitals (see also Section 9.1.6). Finally, loose transition state structures can result from this phenomenon; for instance, the C–Cl bond lengths in the TS structure illustrated in Figure 6.12 are 2.45 and 2.39 A˚ at the BLYP/6- 31G(d) and B3LYP/6-31G(d) levels of theory, respectively. Of course, this is a significant improvement over HF theory, and insofar as TS structures tend to be fairly floppy, the remaining geometrical errors may have only small energetic consequences.

Wiest, Montiel, and Houk (1997) have studied carefully a large number of TS structures for organic electrocyclic reactions and, based on comparison to experiment (particularly including kinetic isotope effect studies) and very high levels of electronic structure theory,

294

8 DENSITY FUNCTIONAL THEORY

concluded that the B3LYP functional is particularly robust for predicting geometries in this area. This is consistent with the good behavior of this functional when applied to minimum-energy structures composed only of first-row atoms as already noted above. Cramer and Barrows (1998) have emphasized, however, that overdelocalization problems can arise in ionic examples of such electrocyclic reactions, and caution may be warranted in these instances.

8.6.3 Charge Distributions

Over the 108 molecules in Test Set B of Table 8.5, Scheiner, Baker, and Andzelm computed the mean unsigned errors in predicted dipole moments to be 0.23, 0.20, 0.23, 0.19, and 0.16 D at the HF, MP2, SVWN, BPW91, and B3PW91 levels of theory, respectively, using the 6-31G(d,p) basis set. These results were improved somewhat for the DFT levels of theory when more complete basis sets were employed.

Cohen and Tantirungrotechai (1999) compared HF, MP2, BLYP, and B3LYP to one another with respect to predicting the dipole moments of some very small molecules using a very large basis set, and their results are summarized in Table 8.6. In general the performances of MP2, the pure BLYP functional, and the hybrid B3LYP functional are about equal, although both DFT functionals do very slightly better than MP2 for several cases. HF theory shows its typical roughly 10–15 percent overestimation of dipole moments, and its historically well-known reversal of moment for carbon monoxide.

In addition to the moments of the charge distribution, molecular polarizabilities have also seen a fair degree of study comparing DFT to conventional MO methods. While data on molecular polarizabilities are less widely available, the consensus appears to be that for this property DFT methods, pure or hybrid, fail to do as well as the MP2 level of theory, with conventional functionals typically showing errors only slightly smaller than those predicted by HF (usually about 1 a.u.), while the MP2 level has errors only 25 percent as large. In certain instances, ACM functionals have been more competitive with MP2, but still not quite as good.

Table 8.6 Dipole moments (D) for eight small molecules at four levels of theory using the very large POL basis seta

Molecule

HF

MP2

BLYP

B3LYP

Experiment

 

 

 

 

 

 

NH3

1.62

1.52

1.48

1.52

1.47

H2O

1.98

1.85

1.80

1.86

1.85

HF

1.92

1.80

1.75

1.80

1.83

PH3

0.71

0.62

0.59

0.62

0.57

H2S

1.11

1.03

0.97

1.01

0.97

HCl

1.21

1.14

1.08

1.12

1.11

CO

−0.25

0.31

0.19

0.10

0.11

SO2

1.99

1.54

1.57

1.67

1.63

a From Cohen and Tantirungrotechai 1999.

 

8.6 GENERAL PERFORMANCE OVERVIEW OF DFT

295

 

Table 8.7 Density Functionalsa

 

Abbreviation

Comments

Reference(s)

 

 

 

 

B

Becke’s 1988 GGA exchange functional

Becke, A. D. 1988. Phys. Rev. A,

 

containing one empirical parameter and

38, 3098.

 

 

showing correct asymptotic behavior.

 

 

B0KCIS

One-parameter hybrid functional of B and

Toulouse, J., Savin, A., and

 

 

KCIS incorporating 25% HF exchange

Adamo, C. 2002. J. Chem.

 

 

(B1KCS optimizes the percent HF exchange

Phys., 117, 10465.

 

 

to 23.9%).

 

 

B1B95

One-parameter hybrid functional of B and B95

Becke, A. D. 1996. J. Chem.

 

 

incorporating 28% HF exchange.

Phys., 104, 1040.

 

B1LYP

One-parameter hybrid functional of B and LYP

Adamo, C. and Barone, V. 1997.

 

incorporating 25% HF exchange.

Chem. Phys. Lett., 274, 242.

 

B1PW91

One-parameter hybrid functional of B and

Adamo, C. and Barone, V. 1997.

 

PW91 incorporating 25% HF exchange.

Chem. Phys. Lett., 274, 242.

 

B3LYP

ACM functional discussed in more detail in

Stephens, P. J., Devlin, F. J.,

 

 

Section 8.4.3.

Chabalowski, C. F., and Frisch,

 

 

M. J. 1994. J. Phys. Chem., 98,

 

 

623.

 

B3LYP*

ACM functional discussed in more detail in

Salomon, O., Reiher, M., and

 

 

Section 8.4.3.

Hess, B. A. 2002. J. Chem.

 

 

 

Phys., 117, 4729.

 

B3PW91

ACM functional discussed in more detail in

Becke, A. D. 1993b. J. Chem.

 

 

Section 8.4.3.

Phys., 98, 5648.

 

B86

Becke’s 1986 GGA exchange functional.

Becke, A. D. 1986. J. Chem.

 

 

 

Phys., 84, 4524.

 

B88

Becke’s 1988 GGA correlation functional.

Becke, A. D. 1988. J. Chem.

 

 

 

Phys., 88, 1053.

 

B95

Becke’s 1995 (sic) MGGA correlation

Becke, A. D. 1996. J. Chem.

 

 

functional.

Phys., 104, 1040.

 

B97

Becke’s 1997 GGA exchange-correlation

Becke, A. D. 1997 J. Chem. Phys.,

 

functional containing 10 optimized

107, 8554.

 

 

parameters including incorporating 19.43%

 

 

 

HF exchange.

 

 

B97-1

Hamprecht, Cohen, Tozer, and Handy hybrid

Hamprecht, F. A., Cohen, A. J.,

 

GGA exchange-correlation functional based

Tozer, D. J., and Handy, N. C.

 

on a reoptimization of empirical parameters

1998. J. Chem. Phys., 109,

 

 

in B97 and incorporating 21% HF exchange.

6264.

 

B97-2

Wilson, Bradley, and Tozer’s hybrid GGA

Wilson, P. J., Bradley, T. J., and

 

exchange-correlation functional based on

Tozer, D. J. 2001. J. Chem.

 

 

further reoptimization of empirical

Phys., 115, 9233.

 

 

parameters in B97 and incorporating 21%

 

 

 

HF exchange.

 

 

B98

Schmider and Becke’s 1998 revisions to the

Schmider, H. L. and Becke, A. D.

 

B97 hybrid GGA exchange-correlation

1998. J. Chem. Phys., 108,

 

 

functional to create a hybrid MGGA

9624.

 

 

incorporating 21.98% HF exchange.

 

 

BB1K

Optimization of B1B95 primarily for kinetics

Zhao, Y., Lynch, B. J., and

 

 

of H-atom abstractions by using 40% HF

Truhlar, D. G. 2004. J. Phys.

 

exchange instead of default 28%.

Chem. A, 108, 2715.

 

(continued overleaf )

296

8 DENSITY FUNCTIONAL THEORY

 

Table 8.7 (continued )

 

 

 

 

Abbreviation

Comments

Reference(s)

 

 

 

Bm

A modification of B88 to optimize its

Proynov, E., Chermette, H., and

 

performance with the τ 1 correlation

Salahub, D. R. 2000. J. Chem.

 

functional.

Phys., 113, 10013.

BR

Becke and Roussel’s 1989 MGGA exchange

Becke, A. D. and Roussel, M. R.

 

functional that includes a dependence on the

1989. Phys. Rev. A, 39, 3761.

 

Laplacian of the density in addition to its

 

 

gradient.

 

CAM

Cambridge GGA exchange functional (denoted

Laming, G. J., Termath, V., and

 

as either CAM(A) or CAM(B))

Handy, N. C. 1993. J. Chem.

 

 

Phys., 99, 8765.

CS

Colle and Salvetti’s correlation functional

Colle, R. and Salvetti, O. 1975.

 

(depending on more than only the density)

Theor. Chim. Acta, 37, 329.

 

parameterized to be exact for the He atom.

 

EDF1

Empirical density functional 1 designed as a

Adamson, R. D., Gill, P. M. W.,

 

pure GGA exchange-correlation functional

and Pople, J. A. 1998. Chem.

 

to be used with small basis sets.

Phys. Lett., 284, 6.

FT97

Filatov and Thiel’s GGA exchange functional.

Filatov, M. and Thiel, W. 1997.

 

 

Mol. Phys., 91, 847.

G96

Gill’s 1996 GGA exchange functional.

Gill, P. M. W. 1996. Mol. Phys.,

 

 

89, 433.

H&H

One-parameter hybrid exchange functional

Becke, A. D. 1993b. J. Chem.

 

combining 50% LSDA with 50% HF

Phys., 98, 1372.

 

exchange.

 

HCTH

Hamprecht, Cohen, Tozer, and Handy GGA

Hamprecht, F. A., Cohen, A. J.,

 

exchange-correlation functional based on a

Tozer, D. J., and Handy, N. C.

 

reoptimization/extension of empirical

1998. J. Chem. Phys., 109,

 

parameters in B97 and a removal of HF

6264. (Most recent refinement,

 

exchange. Now a family of functionals with

Boese, A. D., Martin, J. M. L.,

 

optimizations over different numbers of

and Handy, N. C. 2003. J.

 

test-set molecules, typically denoted

Chem. Phys., 119, 3005.)

 

HCTH/n where n is the number of

 

 

molecules in the test set, e.g., HCTH/93,

 

 

HCTH/120, HCTH/147, and HCTH/407.

 

ISM

Imamura, Scuseria, and Martin’s MGGA

Imamura, Y., Scuseria, G. E., and

 

correlation functional based on CS.

Martin, R. M. 2002. J. Chem.

 

 

Phys., 116, 6458.

KCIS

Kriger, Chen, Iafrate, and Savin’s MGGA

Krieger, J. B., Chen, J., Iafrate, G.

 

correlation functional including a

J., and Savin, A. 1999. In

 

self-interaction correction.

Electron Correlations and

 

 

Materials Properties, Gonis, A.

 

 

and Kioussis, N., Eds., Plenum:

 

 

New York, 463.

KMLYP

Kang and Musgrave two-parameter hybrid

Kang, J. K. and Musgrave, C. B.

 

exchange-correlation functional using a

2001. J. Chem. Phys., 115,

 

mixture of Slater and HF (55.7%) exchange

11040.

 

and a mixture of LSDA and LYP (44.8%)

 

 

correlation functionals.

 

 

 

 

 

8.6 GENERAL PERFORMANCE OVERVIEW OF DFT

297

 

Table 8.7 (continued )

 

 

 

 

 

 

Abbreviation

Comments

Reference(s)

 

 

 

 

 

Lap

MGGA correlation functionals that include a

Proynov, E. I., Sirois, S., and

 

 

dependence on the Laplacian of the density

Salahub, D. R. 1997. Int. J.

 

 

in addition to its gradient, typically denoted

Quantum Chem., 64, 427.

 

 

either Lap1 or Lap3 depending on version.

 

 

LG

Lacks and Gordon’s GGA correlation

Lacks, D. J. and Gordon, R. G.

 

functional.

1993. Phys. Rev. A, 47, 4681.

LT2A

Local square kinetic energy density exchange

Maximoff, S. N., Enzerhof, M.,

 

functional depending only on the kinetic

Scuseria, G. E. 2002. J. Chem.

 

energy density (i.e., not at all on the electron

Phys., 117, 3074.

 

 

density).

 

 

LYP

Lee, Yang, and Parr’s GGA correlation

Lee, C., Yang, W., and Parr, R. G.

 

functional based on the CS functional but

1988. Phys. Rev. B, 37, 785.

 

 

depending only on the density.

 

 

mPBE

Adamo and Barone’s modification of PBE

Adamo, C. and Barone, V. 2002.

 

exchange with PBE correlation.

J. Chem. Phys., 116, 5933.

 

mPW

Adamo and Barone’s modification of PW.

Adamo, C. and Barone, V. 1998.

 

 

J. Chem. Phys., 108, 664.

 

MPW1K

Optimization of mPW1PW91 for kinetics of

Lynch, B. J., Fast, P. L., Harris,

 

H-atom abstractions by using 42.8% HF

M., and Truhlar, D. G. 2000. J.

 

exchange instead of default 25% and the

Phys. Chem. A, 104, 4811.

 

 

6-31+G(d,p) basis set.

 

 

mPW1N

Optimization of mPW1PW91 for

Kormos, B. L. and Cramer, C. J.

 

halide/haloalkane nucleophilic substitution

2002. J. Phys. Org. Chem., 15,

 

reactions by using 40.6% HF exchange

712.

 

 

instead of default 25% and the 6-31+G(d)

 

 

 

basis set.

 

 

MPW1S

Optimization of mPW1PW91 for sugar

Lynch, B. J., Zhao, Y., and

 

 

conformational analysis by using 6% HF

Truhlar, D. G. 2003. J. Phys.

 

exchange instead of default 25% and the

Chem. A, 107, 1384.

 

 

6-31+G(d,p) basis set.

 

 

O

Handy and Cohen OPTX GGA exchange

Handy, N. C. and Cohen, A. J.

 

functional including two optimized

2001. Mol. Phys., 99, 403.

 

 

parameters

 

 

O3LYP

ACM functional discussed in more detail in

Hoe, W.-M., Cohen, A. J., and

 

 

Section 8.4.3.

Handy, N. C. 2001. Chem.

 

 

 

Phys. Lett., 341, 319.

 

P

Perdew’s 1986 GGA exchange functional.

Perdew, J. P. 1986. Phys. Rev. B,

 

 

33, 8822

 

P86

Perdew’s 1986 GGA correlation functional.

Perdew, J. P. 1986. Phys. Rev. B,

 

 

33, 8822

 

PBE

Perdew, Burke, and Enzerhof GGA

Perdew, J. P., Burke, K., and

 

 

exchange-correlation functional.

Enzerhof, M. 1996. Phys. Rev.

 

 

Lett., 77, 3865 and erratum

 

 

 

1997. ibid., 78, 1396.

 

(continued overleaf )

298

8 DENSITY FUNCTIONAL THEORY

 

Table 8.7 (continued )

 

 

 

 

Abbreviation

Comments

Reference(s)

 

 

 

PBE1PBE

One-parameter hybrid PBE functional

Adamo, C., Cossi, M., and

 

incorporating 25% HF exchange (sometimes

Barone, V. 1999. J. Mol. Struct.

 

alternatively called PBE0, PBE0PBE, or

(Theochem), 493, 145.

 

PBE1).

 

PKZB

Perdew, Kurth, Zupan, and Blaha’s MGGA

Perdew, J. P., Kurth, S., Zupan,

 

exchange-correlation functional developed

A., and Blaha, P. 1999. Phys.

 

primarily for solids.

Rev. Lett., 82, 2544.

PW

Perdew and Wang’s GGA exchange functional.

Perdew, J. P. and Wang, Y. 1986.

 

 

Phys. Rev. B, 33, 8800.

PW91

Perdew and Wang’s (sic) 1991 GGA

Perdew, J. P. 1991. In: Electronic

 

correlation functional.

Structure of Solids ’91, Ziesche,

 

 

P. and Eschig, H., Eds.,

 

 

Akademie Verlag: Berlin, 11.

τ 1

A MGGA correlation functional.

Proynov, E., Chermette, H., and

 

 

Salahub, D. R. 2000. J. Chem.

 

 

Phys., 113, 10013.

τ HCTH

MGGA extension of the HCTH functional.

Boese, A. D. and Handy, N. C.

 

Comes in both hybrid and pure DFT

2002. J. Chem. Phys., 116,

 

variations.

9559.

TPSS

Tao, Perdew, Staroverov, and Scuseria’s

Tao, J., Perdew, J. P., Staroverov,

 

MGGA exchange-correlation functional.

V. N., and Scuseria, G. E. 2003.

 

 

Phys. Rev. Lett., 91, 146401.

TPSSh

One-parameter ACM of TPSS incorporating

Staroverov, V. N., Scuseria, G. E.,

 

10% HF exchange.

Tao, J., Perdew, J. P. 2003. J.

 

 

Chem. Phys., 119, 12129.

VSXC

van Voorhis and Scuseria’s MGGA

van Voorhis, T. and Scuseria, G.

 

exchange-correlation functional.

E. 1998. J. Chem. Phys., 109,

 

 

400.

VWN

Local correlation functional of Vosko, Wilk,

Vosko, S. H., Wilk, L., and

 

and Nusair fit to the uniform electron gas.

Nussair, M. 1980. Can. J. Phys.,

 

Note that VWN proposed several different

58, 1200.

 

forms for this functional, usually identified

 

 

by a trailing number, e.g., VWN3 or VWN5.

 

 

Different gradient-corrected and hybrid

 

 

functionals built onto the VWN local

 

 

correlation functional may use different

 

 

versions. For example, B3LYP is defined to

 

 

use VWN3, while O3LYP is defined to use

 

 

VWN5.

 

X

GGA exchange functional defined as a

Xu, X. and Goddard, W. A., III.

 

combination of one part LSDA, 0.722 parts

2004. Proc. Natl. Acad. Sci

 

B, and 0.347 parts PW91.

(USA), 101, 2673.

X3LYP

ACM functional discussed in more detail in

Xu, X. and Goddard, W. A., III.

 

Section 8.4.3.

2004. Proc. Natl. Acad. Sci

 

 

(USA), 101, 2673.

a Exchange, correlation, and specific, specially defined combinations or hybrid functionals are contained herein. Routine combinations of exchange and correlation functionals (e.g., BLYP, OP86, or PWPW91) are not included.

8.7 CASE STUDY: TRANSITION-METAL CATALYZED CARBONYLATION

299

8.7Case Study: Transition-Metal Catalyzed Carbonylation of Methanol

Synopsis of Kinnunen and Laasonen (2001), ‘Reaction Mechanism of the Reductive Elimination in the Catalytic Carbonylation of Methanol. A Density Functional Study’.

Acetic acid is made industrially by the condensation of methanol and carbon monoxide catalyzed by either a diiododicarbonylrhodium species or the corresponding iridium complex. The proposed catalytic cycle for this process is illustrated in Figure 8.5. Experimentally establishing a complete catalytic mechanism can be quite challenging, since reactive intermediates in the cycle may be present at such low concentrations that they are very difficult to detect. Theory can therefore play a useful role in establishing the energetic profiles for proposed catalytic steps, with the ultimate goal being the design of improved catalysts based on a fundamental understanding of the mechanism.

To that end, Kinnunen and Laasonen model the reductive elimination pathways from the anionic acetyltriiododicarbonyl rhodium and iridium anions, and from the acetyldiiodotricarbonyl iridium neutral using the B3LYP functional in combination with an unpolarized

O

HI

MeOH

OH

O

H2O

MeI

I

ICO

M

ICO

reductive elimination

CO

 

or

CO

 

 

Me

I

 

I

 

I

CO

 

 

 

M

O

 

M

O

I

M

I

CO

I

CO

CO

 

I

I

 

 

CO

 

 

I

M O

I CO I

CO

Figure 8.5 Catalytic cycle for the metal-catalyzed carbonylation of methanol, with the reductive elimination step highlighted. In the case of iridium, the diiodotricarbonyl species has also been suggested as a possible precursor to reductive elimination. What are the issues of stereochemistry associated with the intermediates? What special basis-set requirements will be involved in modeling this system?

300

8 DENSITY FUNCTIONAL THEORY

double-ζ basis set on C, H, and O, and a valence basis set of similar size for I and the metals combined with relativistic effective core potentials. Happily, from a simplicity standpoint, all species are predicted to be ground-state singlets by large margins, so a restricted DFT formalism can be employed. In this instance, some experimental data are available for species involved in the reductive elimination step, so the adequacy of the theoretical level can be evaluated.

The authors begin by characterizing the relative energies of all possible stereoisomers in the octahedral complexes. For acetyltriiododicarbonyl metal complexes, there are mer,trans, mer,cis, and fac,cis possibilities (mer implies two iodides to be trans to one another while fac implies all I – M – I bond angles to be about 90; trans and cis refer to whether the central iodine atom in the mer arrangement is opposite the acetyl group or adjacent to it) as well as acetyl rotamers to consider. For both rhodium and iridium, a single mer,trans geometry is predicted to be lower than all other possibilities by at least 2.7 kcal mol−1. Experimental IR and NMR data for the Rh system are in accord with this prediction, while IR data for the Ir system suggest the presence of fac,cis, which is the next lowest energy species predicted from the computations. Kinnunen and Laasonen suggest that weak IR bands for the mer,trans isomer may make it difficult to detect experimentally, and infer that it is possible that both may be present experimentally.

Of course, while the intermediate energies are of interest, so long as interconversion between stereoisomers takes place at lower energy than reductive elimination, the latter process may potentially go through any stereoisomer on the way to the lowest energy TS structure for the reaction (the Curtin – Hammett principle). For the Rh system the lowest energy TS structure, which follows from a mer,cis reactant, has an associated 298 K free energy of activation of 20.1 kcal mol−1, which compares well with an experimental value of about 18. In the case of Ir, a fac,cis TS structure is computed to be slightly lower than the mer,cis structure, and the overall free energy of activation is about 8 kcal mol−1 higher than was the case for Rh. In both cases, iodide dissociation is predicted to proceed with a lower barrier than reductive elimination, so stereoisomer scrambling via elimination/addition should be possible prior to reductive elimination.

Kinnunen and Laasonen carry out a similarly thorough analysis for the diiodotricarbonyliridium case. Consideration of all possibilities is complicated (and will depend experimentally on the iodide ion concentration and carbon monoxide pressure) but in essence ‘all’ stationary points corresponding to stereoisomeric minima and transition state structures for dissociation/association and reductive elimination steps are found and characterized energetically (‘all’ in quotes here because in such complicated systems it is essentially impossible to be entirely certain that every stationary point has been found). This exhaustive mapping of the PES provides insight into the catalytic process in a fashion typically not available experimentally, and takes good advantage of DFT’s ability to handle transition metal systems in an efficient manner.

Bibliography and Suggested Additional Reading

Adamo, C. and Barone, V. 1998. ‘Exchange Functionals with Improved Long-range Behavior and Adiabatic Connection Methods Without Adjustable Parameters: The mPW and mPW1PW Models’,

J. Chem. Phys., 108, 664.

Соседние файлы в предмете Химия