University of Groningen

An ab initio two-component relativistic method including spin- orbit coupling using the regular approximation Faas, S.; van Lenthe, J. H.; Hennum, A. C.; Snijders, J. G.

Published in: Journal of Chemical Physics

DOI: 10.1063/1.1288387

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record

Publication date: 2000

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA): Faas, S., van Lenthe, J. H., Hennum, A. C., & Snijders, J. G. (2000). An ab initio two-component relativistic method including spin- orbit coupling using the regular approximation. Journal of Chemical Physics, 113(10), 4052 - 4059. [PII [S0021-9606(00)32034-7]]. https://doi.org/10.1063/1.1288387

Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

Download date: 21-05-2019 An ab initio two-component relativistic method including spin–orbit coupling using the regular approximation S. Faas, J. H. van Lenthe, A. C. Hennum, and J. G. Snijders

Citation: J. Chem. Phys. 113, 4052 (2000); doi: 10.1063/1.1288387 View online: https://doi.org/10.1063/1.1288387 View Table of Contents: http://aip.scitation.org/toc/jcp/113/10 Published by the American Institute of Physics

Articles you may be interested in Relativistic total energy using regular approximations The Journal of Chemical Physics 101, 9783 (1994); 10.1063/1.467943

Molecular density functional calculations in the regular relativistic approximation: Method, application to coinage metal diatomics, hydrides, fluorides and chlorides, and comparison with first-order relativistic calculations The Journal of Chemical Physics 109, 392 (1998); 10.1063/1.476576

SPOCK.CI: A multireference spin-orbit configuration interaction method for large molecules The Journal of Chemical Physics 124, 124101 (2006); 10.1063/1.2173246

Quasirelativistic theory equivalent to fully relativistic theory The Journal of Chemical Physics 123, 241102 (2005); 10.1063/1.2137315

An improved ab initio relativistic zeroth-order regular approximation correct to order The Journal of Chemical Physics 113, 9957 (2000); 10.1063/1.1323266

Density-functional thermochemistry. III. The role of exact exchange The Journal of Chemical Physics 98, 5648 (1993); 10.1063/1.464913 JOURNAL OF CHEMICAL PHYSICS VOLUME 113, NUMBER 10 8 SEPTEMBER 2000

An ab initio two-component relativistic method including spin–orbit coupling using the regular approximation S. Faas, J. H. van Lenthe,a) and A. C. Hennumb) Group, Debye Institute, Utrecht University, Padualaan 14, 3584 CH Utrecht, The Netherlands J. G. Snijders Theoretical Chemistry, Materials Science Centre, Rijksuniversiteit Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands ͑Received 30 March 2000; accepted 16 June 2000͒ In this paper we present the implementation of the two-component scaled zeroth-order regular approximation ͑ZORA͒ method in the molecular electronic structure package GAMESS-UK.Itisthe first application of this method, which was earlier investigated in the context of density functional theory, in molecular ab initio basis set calculations. The performance of the method is tested in atomic calculations, which we can compare with numerical results, on xenon and radon and in molecular calculations on the molecules AgH, HI, I2, AuH, TlH, and Bi2. In calculations on the I2 molecule we investigated the effect of the different approaches regarding the internal Coulomb matrix used in the ZORA method. For the remaining molecules we computed harmonic frequencies and bond lengths. It is shown that the scaled ZORA approach is a cost-effective alternative to the Dirac–Fock method. © 2000 American Institute of Physics. ͓S0021-9606͑00͒32034-7͔

I. INTRODUCTION basis sets optimized originally for Dirac–Fock calculations, see, e.g., Ref. 15. The inclusion of relativistic effects is essential in the In the present paper we present the first implementation quantum chemical description of compounds containing one and applications of the two-component ZORA Hamiltonian or more heavier elements, see, e.g., Ref. 1. Highly accurate, within an ab initio basis set framework. Unlike the earlier but unfortunately extremely costly, calculations can be per- described scalar ZORA Hamiltonian,7 this Hamiltonian in- formed using the full theory of quantum electrodynamics. cludes spin–orbit coupling. We give a short review of the Due to the computational requirements, these calculations theory followed by details of the implementation of the can only be performed on small atomic systems. To be able method in the GAMESS-UK package.9 Within this implemen- to treat systems of chemical interest it is therefore necessary tation there is room for several approaches regarding the so- to make approximations. called internal Coulomb matrix. An explanation of this con- An accurate method is based on the solution of the four- cept and the different approximations used will be given. component Dirac–Fock equations, the relativistic analog of These approximations are tested in calculations on the I2 Hartree–Fock. This method is still very costly mainly be- molecule. cause of the use of large basis sets needed for a proper de- We test our implementation by calculations on the xenon scription of the small component. An attractive alternative is and radon atoms. These atomic calculations are compared to transform the four-component Dirac equation to a two-, or with numerical ZORA calculations using our ZORA imple- even one-, component formalism. The ZORA ͑zeroth-order mentation in the GRASP2 program.10 regular approximation͒ or CPD method, originally developed In molecular calculations on the diatomics I2, HI, AgH, 2,3 by Chang, Pellisier, and Durand and by Heully et al. and AuH, TlH, and Bi2 we investigate the performance of the later generalized to molecular density functional theory by ZORA Hamiltonian. The quality of the wave functions is 4,5 van Lenthe, Baerends, and Snijders and in its scalar form asserted by comparison of orbital energies (I2) with four- to ab initio methods by the present authors,6,7 performs such component Dirac–Fock results, obtained using the MOLFDIR 11 a transformation. A calculation based on the ZORA method program package. The calculations on the equilibrium ge- can therefore, in principle, be performed using only a large ometries and harmonic frequencies of the molecules HI, component basis, thereby circumventing the use of kineti- AgH, AuH, TlH and Bi2 are used to test the validity of the cally balanced basis sets. Since the ZORA orbitals resemble one-center approximation to the internal Coulomb matrix. the Dirac large components spinors,8 we can make use of Furthermore the results are compared with experiment and full Dirac–Fock results obtained with the DIRAC program.12 The comparison with scalar ZORA calculations unveils the a͒Electronic mail: [email protected] b͒Permanent address: Department of Chemistry, University of Oslo, P.O. effect of the inclusion of spin–orbit coupling in the self- Box 1033 Blindern, N-0315 Oslo, Norway. consistent field ͑SCF͒ procedure.

0021-9606/2000/113(10)/4052/8/$17.004052 © 2000 American Institute of Physics J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Two-component relativistic method 4053

II. THEORY ␳͑ Ј͒ϭ ␺†͑ ͒␺ ͑ Ј͒ ͑ ͒ r,r ͚ i r i r . 7 Using an expansion in E/(2c2ϪV) on the Foldy– iocc Wouthuysen transformed Dirac–Fock equation one can de- It implies rive the ab initio ZORA and scaled ZORA equations.6 The ϩ Ϸ ͑ ͒ equations, which still contain the large and small compo- J␾␾ J␹␹ J␺␺ . 8 nents of the Dirac spinor, read In the same spirit we assume

1 K␾␾ϷK␺␺ , FZORA␺ ϭ⑀ZORA␺ , FZORA␺ ϭ⑀ZORA ␺ , ͑1͒ i i i S i scaled,i i ͑9͒ K␾␹ ,K␹␾Ϸ0. where ␺ is the two-component ZORA orbital. The ZORA i This leads to the removal of the off-diagonal and small Fock operator is given by component Coulomb and exchange operators. The resulting 1 ZORA Fock operator is given by ZORAϭ͑ ␴ Ϫ ͒ ͑ ␴ Ϫ ͒ F c p K␾␹ 2 ␾␹ c p K␹␾ • ͑2c ϪV ϩK␹␹͒ • c2 c ZORA ␺ F ϭ␴ p ␺ ␴ pϩV ϪK␺␺ ͑10͒ ϩ ␾␹Ϫ ͑ ͒ • 2c2ϪV • c Vc K␾␾ 2 c and the scaling factor by with ␺ϭ ϩ ͑ ͒ Sϭ ϩ͗␺ ͉͑ ␴ Ϫ ͒ Vc Vnuc J␺␺ 11 1 i c •p K␾␹ and the scaling factor 1 ϫ ͑ ␴ Ϫ ͉͒␺ ͘ ͑ ͒ ͑ 2Ϫ ␾␹ϩ ͒2 c •p K␹␾ i . 3 c2 2c Vc K␹␹ Sϭ1ϩ͗␺ ͉͑␴ p͒ ␺ ͑␴ p͉͒␺ ͘. ͑12͒ i • ͑2c2ϪV ͒2 • i We use the abbreviation c ␾␹ We can now perform two-component ZORA calcula- V ϭV ϩJ␾␾ϩJ␹␹ . ͑4͒ c nuc tions using only a large component basis set. The large and small components of the original Dirac ␾ ␹ ␺ spinor are denoted by i and i . The relation between i , ␾ ␹ III. IMPLEMENTATION i , and i is determinated by the Foldy–Wouthuysen transformation6 and has become energy independent by the The two-component ZORA method is implemented in a use of the expansion in E/(2c2ϪV). J and K are the Cou- separate module in the GAMESS-UK package9 that allows only lomb and exchange operators, c is the speed of light, ␴ are SCF calculations. The difference from the previously pre- the Pauli matrices. Equations ͑2͒ and ͑3͒ differ from the local sented scalar ZORA scheme is that the ZORA Fock operator density form derived earlier4,5 by the presence of the ex- ͑10͒ contains the spin operators ␴. Hence, the orbitals have change operators K␾␹ and K␹␾ coupling the large and small to be expanded in a set of spin orbitals ͕␾␨͖, with ␾ the components. Another difference can be found in the density spatial and ␨ the spin part ͑either ␣ or ␤͒. The ith ZORA used in the different exchange potentials in the denominator orbital can be written as of the ZORA kinetic energy term. In the present formulation ␺ ͑ ͒ϭ ␣ ␾ ͑ ͒␣͑ ͒ϩ ␤ ␾ ͑ ͒␤͑ ͒ ͑ ͒ based on Dirac–Fock, we have an exchange potential K␹␹ i r,s ͚ c␮,i ␮ r s c␮,i ␮ r s , 13 only due to the small–small density ␮ where r denotes the spatial coordinate and s the spin coordi- ␳SS͑ Ј͒ϭ ␹†͑ ͒␹ ͑ Ј͒ ͑ ͒ ␣ ␤ r,r ͚ i r i r , 5 nate. Note that the coefficients c␮,i and c␮,i can be, and in iocc general are, complex. The total density matrix can be written whereas the exchange–correlation potential used in the den- as ͑ ͒ sity functional theory DFT case contains the total density p␣␣ p␣␤ ͩ ͪ ͑ ͒ ␤␣ ␤␤ 14 ␳͑ ͒ϭ ␾†͑ ͒␾ ͑ ͒ϩ␹†͑ ͒␹ ͑ ͒ ͑ ͒ p p r ͚ i r i r i r i r . 6 iocc with Furthermore, the remaining exchange term in the ab ini- ␨␩ ϭ ␨ ␩ ␨ ␩ϭ␣ ␤ ͑ ͒ tio Fock operator contains the large component density. The p␮␯ ͚ c␮*,ic␯,i , , or . 15 density functional formulation again uses an exchange– iocc correlation potential containing the total density. These matrices have the properties At this point the density is still calculated from the large p␣␣ϭ͑p␣␣͒†, p␤␤ϭ͑p␤␤͒†, p␤␣ϭ͑p␣␤͒† ͑16͒ and small component spinors obtained from the backtrans- formed ZORA orbitals. An easier and more practical SCF and in the closed shell case we have the additional relation procedure is obtained when we lose all dependence on the p␣␣ϭ(p␤␤)†. The real parts of p␣␣ and p␤␤ are symmetric, small components.6 In order to achieve this we make use of whereas the imaginary parts of p␣␣ and p␤␤ and the real and an approximate density derived directly from the ZORA or- imaginary parts of p␤␣ and p␣␤ are antisymmetric. The total bitals density can be written as 4054 J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Faas et al.

occ cancel in the second and third terms of Eq. ͑22͒. TZORA then ␳͑ ͒ϭ ͵ ͉␺ ͑ ͉͒2 2 r ͚ i r,s ds reduces to p /2, the nonrelativistic kinetic energy operator. If i s we would have inserted the resolutions directly into Eq. ͑21͒ we would only have had a reliable representation of the ma- ϭ ͑ ␣␣ϩ ␤␤͒␾ ͑ ͒␾ ͑ ͒ ͑ ͒ ͚ p␮␯ p␮␯ ␮ r ␯ r . 17 trix elements if the resolutions were complete in the range of ␮,␯ ␴" Ϫ 2 p and 1/(1 Vc/2c ). This is a far more severe require- The properties ensure the fact that the total density, as ment. given in Eq. ͑17͒, remains real in spite of the complex den- Second, the splitting off of the nonrelativistic kinetic en- ␣␣ sity matrices, even in an open shell case where (p ergy ensures us of having the correct answer in the nonrela- ␤␤ ϩp )␮␯ can have imaginary elements. Moreover we see tivistic limit (c→ϱ). Direct use of the resolutions in Eq. that the two-electron part of the Fock matrix is Hermitian. ͑21͒ leads, in this limit, to the approximation The total two-electron Fock matrix can be written as ␣␣ ␣␣ ␣␤ 1 Ϫ1 J ϪK K T␮␯ϭ ͚ ͗␾␮␨␮͉␴"p͉␾␭␨␭͘S␾ ͗␾␭␨␭͉␴"p͉␾␯␨␯͘ ͩ ͪ ͑ ͒ 2 ␭ ␤␣ ␤␤ ␤␤ 18 K J ϪK ͑23͒ with of the nonrelativistic kinetic energy, again requiring the reso- lution to be complete in the range of ␴"p. In Eq. ͑22͒ the last ␣␣ ␤␤ ␣␣ ␤␤ J␮␯ ϭJ␮␯ ϭ͚ ͑p␬␭ ϩp␬␭ ͒͑␾␮␾␯͉␾␭␾␬͒ ͑19͒ two terms cancel in the nonrelativistic limit since 1 ␬,␭ Ϫ 2 Vc/2c reduces to 1. and We now show that we can rewrite all terms in the ZORA kinetic energy ͑22͒ in terms of already available gradient ␨␩ ␨␩ K␮␯ϭ͚ p␬␭͑␾␮␾␬͉␾␭␾␯͒, ͑20͒ integrals. To this end we realize that the basis set ␬,␭ functions used depend on the difference of the electronic ͑r͒ ␾ Ϫ where ␨ and ␩ are either ␣ or ␤. The two-electron matrices and nuclear (rN) coordinates only: ␮(r rN). So ץ -originating from the symmetric ͑parts of the͒ density matri ces can be constructed using the, already available, unre- ͗␾␮␣͉␴"p͉␾␯␣͘ϭϩi͗␾␮͉ ͉␾␯͘, Zץ stricted Hartree–Fock matrix builders. The antisymmetric N ͑24͒ ץ -parts of the͒ density matrix, however, required the develop͑ ment of new code. ͗␾␮␤͉␴"p͉␾␯␤͘ϭϪi͗␾␮͉ ͉␾␯͘, Zץ The one-electron part of the Fock matrix requires the N evaluation of the matrix elements of TZORA, and ץ ץ 1 1 ZORAϭ ͗␾ ␨ ͉␴" ␴" ͉␾ ␨ ͘ ͑ ͒ T␮␯ ␮ ␮ p p ␯ ␯ . 21 ͗␾␮␣͉␴"p͉␾␯␤͘ϭi͗␾␮͉ Ϫi ͉␾␯͘. ͑25͒ Yץ Xץ Vc 2 1Ϫ N N 2c2 Where we used the standard representation of the spin op- These matrix elements can be calculated most easily by in- erators, the Pauli matrices sertion of resolutions of identity13. Furthermore we split off Ϫ the nonrelativistic kinetic energy. We write 01 0 i 10 ␴ ϭͩ ͪ , ␴ ϭͩ ͪ , ␴ ϭͩ ͪ . ͑26͒ x 10 y i 0 z 0 Ϫ1 1 1 ZORAϭ ␾ ␨ ͉ 2͉␾ ␨ Ϫ ␾ ␨ ͉␴" ͉␾ ␨ T␮␯ ͗ ␮ ␮ p ␯ ␯͘ ͚ ͗ ␮ ␮ p ␭ ␭͘ ␣ϭ 1 ␤ 2 2 ␭ The spin functions are then represented by (0) and ϭ(0). Consider, e.g., the ␣␣ part of the second term of Eq. Ϫ1 1 ϫS␾ ͗␾␭␨␭͉␴"p͉␾␯␨␯͘ ͑22͒. Using the relations in Eqs. ͑24͒ and ͑25͒ we write 1 ϩ ␾ ␨ ͉␴" ͉␾ ␨ Ϫ1 Ϫ1 ͚ ͗ ␮ ␮ p ␭ ␭͘S␾ ͚ ͚ ͗␾␮␣͉␴"p͉␾␭␨͘S␾ ͗␾␭␨͉␴"p͉␾␯␣͘ 2 ␭,␬ ␭ ␨ϭ␣,␤ V ϭ " ϩ " ϩ " ϩ ͑ " Ϫ " ͒ ͑ ͒ c Ϫ1 Dz Dz Dx Dx Dy Dy i Dx Dy Dy Dx , 27 ϫ͗␾␭␨␭͉1Ϫ ͉␾␬␨␬͘ 2c2 where Di denotes the derivative of the overlap matrix with Ϫ1 ϫS␾ ͗␾␬␨␬͉␴"p͉␾␯␨␯͘, ͑22͒ respect to nuclear coordinate iN , Ϫ ץ .where S 1 is the inverted metric in the nonorthogonal basis ␾ ͑ ͒ ϭ͗␾ ͉ ͉␾ ͘ ϭ ͑ ͒ ␯ , iN XN ,Y N ,ZN 28 ץ Di ␮␯ ␮ ͒ ͑ In the last term of Eq. 22 we have used an inner projection, i replacing the matrix of an inverse operator by the matrix N inverse of the operator. and the dot is used to represent a standard matrix multipli- This procedure has two advantages. First, we now need a cation in a nonorthogonal basis basis set in the resolution of the identity that is complete in the range of V /2c2. In the region where V /2c2 is negligible ͑ " ͒ ϭ ͑ ͒Ϫ1 ͑ ͒ c c A B ␮␯ ͚ A␮␬ S␾ ␬␬ЈB␬Ј␯ . 29 the terms arising from the incompleteness of the resolution ␬␬Ј J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Two-component relativistic method 4055

␣␤ ␤␣ ␤␤ The , , and contributions can be rewritten in the electron part (Vnuc) with only intra-atomic parts. same way. The last term of Eq. ͑22͒ contains no extra diffi- ͑2͒ Use the density of the atomic startup for evaluation ␾ ␨ ͉ Ϫ 2 ͉␾ ␨ culties since ͗ ␭ ␭ 1 (Vc/2c ) ␬ ␬͘ is diagonal in spin of the Coulomb matrix in the ZORA corrections and keep ͑ ␨ ϭ␨ " ϩ " ϩ " so ␭ ␬). The real part, Dz Dz Dx Dx Dy Dy , is also the ZORA corrections constant during the remainder of the present in our implementation of the scalar ZORA method. SCF procedure. As noted earlier,7 it is convenient to use two different basis sets in our calculations. The first one, the external ba- IV. BASIS SETS sis, is the one used in the electronic structure calculation. In the calculations on the xenon atom we have used two The latter, the internal basis, is the one used in the resolution different basis sets. The first one, which we will call‘‘ad- of the identity. There are two aspects in the construction of justed Huzinaga,’’ was originally developed for nonrelativis- the internal basis. tic calculations14 and was augmented with extra high expo- ͑1͒ The ͗␾͉p͉␹͗͘␹͉( ) products of integrals should ¯ nent s- and p-type functions and with diffuse d- and f-type be correctly represented. Consider a matrix element ͗s͉p͉p͘, functions to account for the relativistic contraction and ex- where s and p denote an s- and p-type Gaussian basis func- pansion of the corresponding orbitals. The basis set is used tion. The momentum operator working on a high exponent uncontracted. The second basis set is taken from Dyall15 and ͗s͉ produces a ͗p͉ with the same exponent. This high expo- is the large component basis of a basis originally developed nent p-type Gaussian is, in general, not present in our exter- for Dirac–Fock calculations. It is again used uncontracted. In nal basis set. The internal basis, however, should contain this the case of the radon atom we followed the same recipe as function. Generalizing this argument to p-, d-, and f-type for the xenon ‘‘adjusted Huzinaga’’ set. In the corresponding functions we see that the matrix elements are represented Dirac–Fock calculation the small component functions were ͕␹͖ʚ "͕␾͖ exactly if we define our internal basis as p . generated by kinetic balance. ͑ ͒ ␺ ϭ ϩ 2 The Coulomb matrix VC Vnuc J␺␺ appearing in In the case of the I2 molecule we used a basis from the inverse operator needs to be evaluated in the internal Poirier, Kari, and Csizmedia16 again augmented with extra basis. This is generally done by projecting the external den- functions and used uncontracted. sity matrix onto the internal basis and building the Coulomb The calculations on the HI molecule were performed potential from the resulting density. This projection is exact with a Dyall set15 on I. The basis sets used on the silver, if the internal basis contains the external basis. Hence the gold, thallium, and bismuth atoms are again based on non- ͕␹͖ʚ͕␾͖ condition: ensures an exact projection. relativistic basis sets taken from Huzinaga and adjusted to In practice we have chosen to construct the internal basis incorporate relativistic effects. In all cases we used a SV by copying all functions from the external basis onto the 3-21G set on hydrogen.17 internal basis, thereby fulfilling the second condition com- The adjusted Huzinaga and Poirier sets are available pletely. The internal basis is successively augmented with upon request. All molecular calculations have been per- p-type functions with exponents of external s- functions formed using spherical harmonic functions. The atomic cal- higher than the highest external p-type functions; with d-type culations are performed using a Cartesian basis. functions with exponents of external p-functions higher than the highest external d-type functions; and so on. This means V. RESULTS that we approximate the first condition. The resulting internal ͑ A. Atomic calculations basis size Nint) is usually considerably bigger than the ex- ͑ ternal basis size Next). Table I lists the orbital energies for the xenon atom from The evaluation of the Coulomb matrix in the internal numerical as well as basis set Dirac–Fock and scaled ZORA basis is done using a direct algorithm. The screening of in- calculations. The basis set results are obtained with the ad- tegrals ensures the fact that we only compute integrals that justed Huzinaga set and the larger basis set from Dyall. are contracted with a nonzero element of the density matrix. Looking at the numerical results we see that, except for Consequently, integrals of the type ͑␹␹͉␹␹͒ are avoided and the deep core region, the scaled ZORA method reproduces the calculation of the internal Coulomb matrix is an opera- the Dirac–Fock results closely. The error ranges from 1.73 2 2 4 tion of order NextNint instead of Nint. hartree for the 1s to 0.0001 hartree for the 5p3/2 . To be able to reduce the computational work even fur- The basis set error can be compared for the adjusted ther we have implemented several possibilities for the con- Huzinaga and Dyall basis sets. For both sets it is seen that struction of the internal Coulomb matrix. the basis set error made by scaled ZORA is of the same order ͑1͒ Project the external density matrix onto the internal in magnitude as the basis set error in the Dirac–Fock calcu- basis and build the full or one-center Coulomb matrix. The lation in the same set. Sometimes, however, the error in the latter is of course an approximation of the first, but, since scaled ZORA results is slightly larger. This is probably due 2 Vc/2c is only large near the nuclei, we expect that this extra to the incompleteness of the internal basis. Of course, the approximation does not result in a serious loss of accuracy. absolute value of the basis set error, for both methods, is The effect of the one-center approximation has been investi- smaller in the larger basis set from Dyall. gated and will be commented on in Sec. V. This approxima- The relativistic effect in the orbital energies of the xenon tion has been made to reduce the computational cost of the atom is 50 hartree for the 1s and decreases to 0.06 hartree, evaluation of the two-electron Coulomb matrix in the inter- which is still significant, for the 5s orbital. The spin–orbit nal basis. Note that it is also necessary to build the one- splitting in the 5p shell is still 0.05 hartree. 4056 J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Faas et al.

TABLE I. Orbital energies for the xenon atom obtained with the Dirac–Fock and scaled ZORA methods.

Dirac–Fock ZORA Dirac–Fock ZORA Dirac–Fock ZORA numeric numeric Huzinagaa Huzinagaa Dyallb Dyallb

1s Ϫ1277.3687 Ϫ1275.6352 Ϫ1276.2021 Ϫ1274.8530 Ϫ1277.2511 Ϫ1275.5423 2s Ϫ202.4784 Ϫ202.2698 Ϫ202.3546 Ϫ202.2127 Ϫ202.4603 Ϫ202.2547 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 2p1/2 189.6793 189.6160 189.6442 189.4510 189.6729 189.5154 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 2p3/2 177.7038 177.5994 177.7118 177.6218 177.7003 177.6275 3s Ϫ43.0131 Ϫ42.9738 Ϫ42.9836 Ϫ42.9724 Ϫ43.0057 Ϫ42.9673 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 3p1/2 37.6598 37.6513 37.6609 37.6068 37.6549 37.6120 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 3p3/2 35.3251 35.3078 35.3345 35.3215 35.3208 35.3156 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 3d3/2 26.0232 26.0232 26.0162 26.0045 26.0181 26.0117 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 3d5/2 25.5369 25.5311 25.5382 25.5313 25.5323 25.5308 4s Ϫ8.4305 Ϫ8.4232 Ϫ8.4254 Ϫ8.4246 Ϫ8.4255 Ϫ8.4185 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 4p1/2 6.4525 6.4520 6.4525 6.4340 6.4479 6.4315 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 4p3/2 5.9827 5.9804 5.9848 5.9870 5.9785 5.9823 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 4d3/2 2.7113 2.7125 2.7095 2.7066 2.7069 2.7046 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 4d5/2 2.6337 2.6339 2.6337 2.6350 2.6294 2.6316 5s Ϫ1.0102 Ϫ1.0094 Ϫ1.0092 Ϫ1.0098 Ϫ1.0070 Ϫ1.0061 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 5p1/2 0.4926 0.4927 0.4924 0.4871 0.4888 0.4840 Ϫ Ϫ Ϫ Ϫ Ϫ Ϫ 5p3/2 0.4398 0.4397 0.4398 0.4417 0.4367 0.4385

aModified, based on Ref. 14. bReference 15.

The method has also been tested on the heavier radon orbitals, respectively. However, even here the error is small atom, Table II. The relativistic effects in this atom are even compared to the relativistic effect. The agreement in the larger, ranging from 414 hartree on the 1s orbital to 0.2 ͑sub͒ valence shells is very good. hartree on the 6s. The spin–orbit splitting is as large as 101 Comparing the basis set results for Dirac–Fock and hartree for the 2p shell and decreases to 0.16 hartree in the scaled ZORA, obtained in the adjusted Huzinaga set, with valence shell. the numerical results, we see that the basis set errors are The numerical scaled ZORA results deviate significantly again of the same order in magnitude. The basis set errors are from the numerical Dirac–Fock results in the deep core re- always significantly smaller that the size of the relativistic gion. The errors are 9.1 and 1.3 hartree for the 1s and 2s effect.

TABLE II. Orbital energies for the radon atom obtained with the Dirac–Fock, scaled ZORA, and nonrelativ- istic method.

Dirac–Fock ZORA Dirac–Fock ZORA Nonrel. numeric numeric Huzinagaa Huzinagaa numeric

1s Ϫ3644.8056 Ϫ3635.6847 Ϫ3635.2927 Ϫ3627.6551 Ϫ3230.3128 1s 2s Ϫ669.3867 Ϫ668.0729 Ϫ665.0317 Ϫ666.0655 Ϫ556.9131 2s Ϫ Ϫ Ϫ Ϫ Ϫ 2p1/2 642.3553 641.7975 641.1660 639.2556 536.6770 2p Ϫ Ϫ Ϫ Ϫ 2p3/2 541.0813 540.5567 541.7030 541.4071 3s Ϫ166.9675 Ϫ166.6788 Ϫ165.2106 Ϫ165.9139 Ϫ138.4219 3s Ϫ Ϫ Ϫ Ϫ Ϫ 3p1/2 154.9012 154.7955 154.6073 154.1382 128.6716 3p Ϫ Ϫ Ϫ Ϫ 3p3/2 131.7246 131.6164 131.8994 131.9312 Ϫ Ϫ Ϫ Ϫ Ϫ 3d3/2 112.5611 112.5389 112.3169 112.1751 110.7013 3d Ϫ Ϫ Ϫ Ϫ 3d5/2 107.7535 107.7045 107.7484 107.6409 4s Ϫ41.3487 Ϫ41.2782 Ϫ40.8703 Ϫ41.0651 Ϫ33.9207 4s Ϫ Ϫ Ϫ Ϫ Ϫ 4p1/2 36.0209 36.0002 35.9615 35.8068 29.4912 4p Ϫ Ϫ Ϫ Ϫ 4p3/2 30.1186 30.0962 30.1860 30.2115 Ϫ Ϫ Ϫ Ϫ Ϫ 4d3/2 21.5464 21.5465 21.4956 21.4501 21.3313 4d Ϫ Ϫ Ϫ Ϫ 4d5/2 20.4371 20.4305 20.4580 20.4345 Ϫ Ϫ Ϫ Ϫ Ϫ 4 f 5/2 9.1927 9.1990 9.2241 9.2120 10.1076 4 f Ϫ Ϫ Ϫ Ϫ 4 f 7/2 8.9270 8.9293 8.9644 8.9575 5s Ϫ8.4168 Ϫ8.4025 Ϫ8.3145 Ϫ8.3572 Ϫ6.9058 5s Ϫ Ϫ Ϫ Ϫ Ϫ 5p1/2 6.4090 6.4066 6.4021 6.3465 5.2252 5p Ϫ Ϫ Ϫ Ϫ 5p3/2 5.1752 5.1724 5.1949 5.2058 Ϫ Ϫ Ϫ Ϫ Ϫ 5d3/2 2.1892 2.1910 2.1844 2.1712 2.3263 5d Ϫ Ϫ Ϫ Ϫ 5d5/2 2.0161 2.0168 2.0232 2.0194 6s Ϫ1.0727 Ϫ1.0707 Ϫ1.0559 Ϫ1.0620 Ϫ0.8740 6s Ϫ Ϫ Ϫ Ϫ Ϫ 6p1/2 0.5403 0.5404 0.5398 0.5216 0.4280 6p Ϫ Ϫ Ϫ Ϫ 6p3/2 0.3839 0.3839 0.3866 0.3896

aModified, based on Ref. 14. J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Two-component relativistic method 4057

TABLE III. Orbital energies of selected orbitals of I2 at 5.0 bohr, obtained bohr͒ displayed in Fig. 1 is obtained from fitting our I with the scaled ZORA method, using different Coulomb options and the 2 Dirac–Fock method. data. The largest spin–orbit splitting can, of course, be found ZORA ZORA ZORA in the 2␲ orbital. Since the spin–orbit splitting is large and Dirac–Fock full one center atomic start the interaction splitting, i.e., the splitting between gerade and 1␴ Ϫ1225.9171 Ϫ1224.3790 Ϫ1224.3786 Ϫ1224.3839 ungerade orbitals, is very small, the observed energy levels should be interpreted as being atomic. The two 2p orbitals 2␴ Ϫ193.0169 Ϫ192.8347 Ϫ192.8346 Ϫ192.8355 1/2 are combined in the highest 2␲ orbital, for which the gerade Ϫ180.4957 Ϫ180.2990 Ϫ18.2989 Ϫ180.2998 and ungerade combination have the same energy. The re- 2␲ Ϫ169.5547 Ϫ169.4885 Ϫ169.4884 Ϫ169.4886 maining 2p orbitals are combined in four orbitals, for Ϫ Ϫ Ϫ Ϫ 3/2 169.5538 169.4874 169.4874 169.4876 which the ungerade and gerade combinations are again de- 4␴ Ϫ7.7508 Ϫ7.7548 Ϫ7.7458 Ϫ7.7458 generate. These levels are slightly split by the linear part of Ϫ5.8583 Ϫ5.8416 Ϫ5.8416 Ϫ5.8416 the electric field of the nuclei. The spin–orbit splitting cal- ␲ Ϫ Ϫ Ϫ Ϫ 4 5.4490 5.4541 5.4541 5.4541 culated with the Dirac–Fock method is 10.94 hartree. It is Ϫ5.4384 Ϫ5.4428 Ϫ5.4428 Ϫ5.4428 reproduced by scaled ZORA within 0.15 hartree. Ϫ2.3304 Ϫ2.3298 Ϫ2.3298 Ϫ2.3298 It is remarkable that the absence of splitting between the Ϫ2.3134 Ϫ3.3128 Ϫ2.3128 Ϫ2.3128 gerade and ungerade orbitals persists up to the subvalence ␦ Ϫ Ϫ Ϫ Ϫ 4 2.2655 2.2686 2.2686 2.2686 shell. Only in the valence shell is a significant interaction Ϫ2.2566 Ϫ2.2595 Ϫ2.2595 Ϫ2.2595 Ϫ2.2427 Ϫ2.2454 Ϫ2.2454 Ϫ2.2454 induced splitting observed. Table V gives the equilibrium distances for the HI, AgH, ␴ Ϫ Ϫ Ϫ Ϫ 5 0.9749 0.9751 0.9751 0.9751 AuH, TlH, and Bi molecules at the nonrelativistic, scalar 5␴* Ϫ0.8421 Ϫ0.8421 Ϫ0.8421 Ϫ0.8421 2 5␴ Ϫ0.4678 Ϫ0.4645 Ϫ0.4645 Ϫ0.4645 ZORA, and two-component ZORA levels of calculation. ␲ Ϫ Ϫ Ϫ Ϫ 5 1/2 0.4589 0.4558 0.4558 0.4558 Looking at the corresponding one-center and full Coulomb ␲ Ϫ Ϫ Ϫ Ϫ 5 3/2 0.4307 0.4326 0.4326 0.4326 results we can conclude that the one-center approximation ␲ Ϫ Ϫ Ϫ Ϫ 5 1/2* 0.3700 0.3719 0.3719 0.3719 has an effect of 0.02–0.03 Å on equilibrium bond lengths of ␲* Ϫ0.3477 Ϫ0.3501 Ϫ0.3501 Ϫ0.3501 5 3/2 compounds containing a sixth row element ͑AuH and TlH͒. This error is of the same magnitude as the error which is usually regarded as allowable in bond length calculations. For Bi2 the effect is as large as 0.11 Å for scalar ZORA and 0.10 Å for spin–orbit ZORA. Even for the fifth row element B. Molecular calculations Ag, the effect is 0.01 Å. The effect of spin–orbit coupling is only noticeable on the TlH and Bi molecules. For TlH we In calculations on the I2 molecule we investigated the 2 effects of the different types of internal Coulomb matrix op- have a spin–orbit contraction of 0.03 Å for scaled ZORA. tions. Table III shows the orbital energies of selected orbitals For these molecules we have also computed the har- for Dirac–Fock and scaled ZORA basis set calculations. The monic frequencies, Table VI. The spin–orbit effect is pro- Ϫ1 scaled ZORA results are produced using the full, one-center, nounced for Bi2, where we have a decrease of 13 cm . For Ϫ1 and atomic-start-up density Coulomb options. We see that TlH an increase of 51 cm is found. The other molecules do the effects of the use of the different Coulomb options are not show a significant change. The use of the one-center only visible in the core region. However, even there the dif- approximation for the TlH and Bi2 molecules introduces, ex- ferences between the scaled ZORA results are very small and cept for the scalar ZORA TlH result, an error ranging from the orbital energies are equally close to the Dirac–Fock ones. 15% to 30%. ͑ We note that the I molecule is a simple nonpolar ex- Comparing the most accurate ZORA calculations two- 2 ͒ ample. It might be expected that the cheapest option, the component scaled ZORA with the full Coulomb option with atomic start-up option, does not perform very well in ionic Dirac–Fock values and experiment, we see that the error molecules since it uses a Coulomb potential, in the ZORA made by scaled ZORA compared to experiment is compa- corrections, based on the superposition of the densities of the rable to the error made by Dirac–Fock. This holds for both neutral atoms. the equilibrium bond lengths and the harmonic frequencies. The orbital energies of the I2 molecule produced by Dirac–Fock and one-center scaled ZORA calculation are VI. CONCLUSION listed in Table IV. Except for the deepest core levels the agreement between Dirac–Fock and scaled ZORA is very The two-component ZORA method has been imple- good. The deviation of the scaled ZORA results from the mented in the GAMESS-UK package. In atomic calculations Dirac–Fock numbers is of the same order of magnitude as we showed that the error of ZORA orbital energies obtained calculated in our numerical xenon calculations. This suggests in a basis set calculation, compared to numerical calcula- that the one-center approximation as well as the incomplete- tions, is of the same order of magnitude as the basis set ness of the internal basis have not resulted in a loss in accu- errors in the corresponding Dirac–Fock calculation. More- racy. However, it must be noted that the error introduced by over, we demonstrated that the deviations of the scaled the one-center approximation can be much larger for other ZORA results from the Dirac–Fock numbers is in every case interatomic distances, Fig. 1. The equilibrium distance ͑5.33 much smaller than the size of the relativistic effect. The basis 4058 J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Faas et al.

TABLE IV. Orbital energies for the I2 molecule at 5.0 bohr, obtained with the scaled ZORA method, with the one-center Coulomb matrix option, and the Dirac–Fock method.

Dirac–Fock ZORA ⌬E Dirac–Fock ZORA ⌬E

1␴ Ϫ1225.9171 Ϫ1224.3786 Ϫ1.5385 Ϫ1225.9171 Ϫ1224.3786 Ϫ1.5385 1␴*

2␴ Ϫ193.0169 Ϫ192.8346 Ϫ0.1823 Ϫ193.0169 Ϫ192.8346 Ϫ0.1823 2␴*

Ϫ180.4957 Ϫ180.2989 Ϫ0.1968 Ϫ180.4957 Ϫ180.2989 Ϫ0.1968 2␲ Ϫ169.5547 Ϫ169.4884 Ϫ0.0662 Ϫ169.5547 Ϫ169.4884 Ϫ0.0662 2␲* Ϫ169.5538 Ϫ169.4874 Ϫ0.0665 Ϫ169.5538 Ϫ169.4874 Ϫ0.0665

3␴ Ϫ40.5096 Ϫ40.4772 Ϫ0.0325 Ϫ40.5096 Ϫ40.4772 Ϫ0.0325 3␴*

Ϫ35.3098 Ϫ35.2597 Ϫ0.0502 Ϫ35.3098 Ϫ35.2597 Ϫ0.0502 3␲ Ϫ33.1988 Ϫ33.1951 Ϫ0.0036 Ϫ33.1988 Ϫ33.1951 Ϫ0.0036 3␲* Ϫ33.1946 Ϫ33.1906 Ϫ0.0040 Ϫ33.1946 Ϫ33.1906 Ϫ0.0040

Ϫ24.1560 Ϫ24.1520 Ϫ0.0040 Ϫ24.1560 Ϫ24.1520 Ϫ0.0040 Ϫ24.1505 Ϫ24.1465 Ϫ0.0041 Ϫ24.1505 Ϫ24.1465 Ϫ0.0041 3␦ Ϫ23.7216 Ϫ23.7221 0.0005 Ϫ23.7216 Ϫ23.7221 0.0005 3␦* Ϫ23.7193 Ϫ23.7198 0.0004 Ϫ23.7193 Ϫ23.7198 0.0004 Ϫ23.7146 Ϫ23.7150 0.0003 Ϫ23.7146 Ϫ23.7150 0.0003

4␴ Ϫ7.7508 Ϫ7.7458 Ϫ0.0050 Ϫ7.7507 Ϫ7.7458 Ϫ0.0050 4␴*

Ϫ5.8583 Ϫ5.8416 Ϫ0.0167 Ϫ5.8582 Ϫ5.8415 Ϫ0.0167 4␲ Ϫ5.4490 Ϫ5.4541 0.0052 Ϫ5.4488 Ϫ5.4539 0.0051 4␲* Ϫ5.4384 Ϫ5.4428 0.0044 Ϫ5.4383 Ϫ5.4428 0.0044

Ϫ2.3304 Ϫ2.3298 Ϫ0.0006 Ϫ2.3303 Ϫ2.3296 Ϫ0.0006 Ϫ2.3134 Ϫ2.3128 Ϫ0.0007 Ϫ2.3134 Ϫ2.3126 Ϫ0.0007 4␦ Ϫ2.2655 Ϫ2.2686 0.0031 Ϫ2.2652 Ϫ2.2684 0.0031 4␦* Ϫ2.2566 Ϫ2.2595 0.0029 Ϫ2.2564 Ϫ2.2593 0.0029 Ϫ2.2427 Ϫ2.2454 0.0027 Ϫ2.2426 Ϫ2.2454 0.0027

5␴ Ϫ0.9749 Ϫ0.9751 0.0002 Ϫ0.8421 Ϫ0.8421 Ϫ0.0001 5␴*

5␴ Ϫ0.4678 Ϫ0.4645 Ϫ0.0033 ␲ Ϫ Ϫ Ϫ Ϫ Ϫ ␲ 5 1/2 0.4589 0.4558 0.0031 0.3700 0.3719 0.0019 5 1/2* ␲ Ϫ Ϫ Ϫ Ϫ ␲ 5 3/2 0.4307 0.4326 0.0019 0.3477 0.3501 0.0024 5 3/2*

set error of the employed sets is also much smaller. This In calculations of the equilibrium geometries and har- brings the calculation on molecules of chemical significance monic frequencies of the molecules HI, AgH, TlH, AuH and within reach. Bi2, it transpires that the spin–orbit effect on the bond For the I molecule it was shown that already a pure 2 lengths is only noticeable in the TlH and Bi2 molecules. atomic ZORA correction is significantly accurate.

FIG. 1. Error made by the one-center Coulomb option with respect to the full Coulomb option with the scaled

ZORA method for the I2 molecule. The vertical dashed line indicates the geometry used in Tables III and IV. J. Chem. Phys., Vol. 113, No. 10, 8 September 2000 Two-component relativistic method 4059

TABLE V. Equilibrium distances ͑Å͒ from nonrelativistic, scalar ͑sr͒, and For ionization potentials the method gives acceptable results ͑ ͒ ͑ two-component tc scaled ZORA using the one-center and full Coulomb except for the deepest levels, e.g., the 1s and 2sp shells of option͒ calculations. Comparison with experiment and Dirac–Fock values. radon. The suitability of ZORA for other properties requires

AgH HI AuH TlH Bi2 further investigation.

Non-Rel. 1.74 1.61 1.78 1.93 2.61 ACKNOWLEDGMENTS sr ZORA One center 1.68 1.60 1.65 1.93 2.64 tc ZORA One center 1.68 1.60 1.65 1.92 2.63 We thank Professor I. P. Grant and Professor Marketos sr ZORA Full 1.67 1.60 1.62 1.91 2.53 10 for providing us with the GRASP2 package, and Professor tc ZORA Full 1.67 1.60 1.62 1.88 2.53 Dirac–Fock 1.67 1.60 1.62 1.87 2.59 W. C. Nieuwpoort and Dr. W. A. de Jong for making the 11 Experiment 1.617a 1.608a 1.5237a 1.87a 2.66b MOLFDIR program available to us. a Reference 18. 1 P. Pyykko¨, Chem. Rev. 88, 563 ͑1988͒. b Reference 19. 2 Ch. Chang, M. Pelissier, and Ph. Durand, Phys. Scr. 34,394͑1986͒. 3 J.-L. Heully, I. Lindgren, E. Lindroth, S. Lundquist, and A.-M. Ma˚rtensson-Pendrill, J. Phys. B 19, 2799 ͑1986͒. Upon inclusion of spin–orbit coupling the calculated bond 4 E. van Lenthe, E. J. Baerends, and J. G. Snijders, J. Chem. Phys. 99, 4597 ͑ ͒ length for TlH decreases with 0.03 Å. 1993 . 5 E. van Lenthe, E. J. Baerends, and J. G. Snijders, J. Chem. Phys. 101, The effect of the one-center approximation on the bond 1272 ͑1994͒. lengths is about 0.03 Å for compounds containing one sixth 6 S. Faas, J. G. Snijders, J. H. van Lenthe, E. van Lenthe, and E. J. Baer- row element. For Bi the errors are 0.11 and 0.10 Å for scalar ends, Chem. Phys. Lett. 246, 632 ͑1995͒. 2 7 and two-component ZORA, respectively. In the computation S. Faas, J. G. Snijders, and J. H. van Lenthe, Quantum Systems in Chem- istry and Physics, Basic Problems and Model Systems, Vol. 1 ͑Kluwer of harmonic frequencies, the one-center approximation pro- Academic, Dordrecht, 2000͒,p.251. 8 duces a serious error in the TlH and Bi2 molecules. These A. J. Sadlej, J. G. Snijders, E. van Lenthe, and E. J. Baerends, J. Chem. deviations are a result of the variation of the error introduced Phys. 102, 1758 ͑1995͒. 9 by the one-center approximation with the bond distance, Fig. GAMESS-UK is a package of ab initio programs written by M. F. Guest, J. H. van Lenthe, J. Kendrick, K. Schoffel, P. Sherwood, and R. J. Harrison, 1. It can be concluded that the one-center approximation can with contributions from R. D. Amos, R. J. Buenker, M. Dupuis, N. C. only be used in systems with only one fifth row atom. Handy, I. H. Hillier, P. J. Knowles, V. Bonacic-Koutecky, W. von Nies- For the systems considered, the two-component scaled sen, V. R. Saunders, and A. J. Stone. The package is derived from the ZORA method is able to achieve the same degree of accu- original GAMESS code due to M. Dupuis, D. Spangler, and J. Wendoloski, NRCC Catalog, Vol. 1, Program No. QG01 ͑GAMESS͒͑1980͒. racy as is obtained using the full four-component Dirac– 10 F. A. Parpia, I. P. Grant, and C. F. Fischer, Comput. Phys. Commun. 94, Fock method. These results suggest that the two-component 249 ͑1996͒. 11 scaled ZORA approach is a cost-effective alternative to L. Visscher, O. Visser, P. J. C. Aerts, H. Merenga, and W. C. Nieuwpoort, Comput. Phys. Commun. 81, 120 ͑1994͒. 12 Dirac–Fock for the calculation of potential energy surfaces. ‘‘DIRAC, a relativistic ab initio electronic structure program, Release 3.1 ͑1998͒,’’ T. Saue, T. Enevoldsen, T. Helgaker, H. J. Aa. Jensen, J. K. Laerdahl, K. Ruud, J. Thyssen, and L. Visscher ͑http://dirac.chem.sdu.dk͒. Ϫ TABLE VI. Harmonic frequencies ͑cm 1͒ from nonrelativistic, scalar ͑sr͒, 13 E. van Lenthe, R. van Leeuwen, E. J. Baerends, and J. G. Snijders, Int. J. and two-component ͑tc͒ scaled ZORA ͑using the one-center and full Cou- Quantum Chem. 57, 281 ͑1996͒. lomb option͒ calculations. Comparison with experiment and Dirac–Fock 14 S. Huzinaga, J. Andzelm, M. Klobukowski, E. Radzio-Andzelm, Y. Sakai, values. and H. Tatewaki, Gaussian Basis Sets for Molecular Calculations ͑Elsevier, Amsterdam, 1984͒. 15 ͑ ͒ AgH HI AuH TlH Bi2 K. G. Dyall, Theor. Chem. Acc. 99, 366 1998 . 16 R. Poirier, R. Kari, and I. G. Csizmadia, Handbook of Gaussian Basis Nonrel. 1561 2421 1608 1534 238 Sets, a Compendium for Ab Initio Molecular Orbital Calculations sr ZORA One center One 1693 2432 2110 1459 293 ͑Elsevier, Amsterdam, 1985͒. tc ZORA One center 1694 2431 2110 1631 301 17 J. S. Binkley, J. A. Pople, and W. J. Hehre, J. Am. Chem. Soc. 102, 939 sr ZORA Full 1752 2349 1979 1397 240 ͑1980͒. tc ZORA Full 1751 2346 1978 1448 227 18 Table of Interatomic Distances and Configuration in Molecules and Ions, Dirac–Fock 1691 2426 1981 1477 191 edited by A. D. Mitchell and L. C. Cross ͑Thanet, Markate, Kent, UK, Experimenta 1760 2309 2305 1391 173 1958͒. 19 K. P. Huber and G. Herzberg, Constants of Diatomic Molecules ͑Van aReference 19. Nostrand Reinhold, New York, 1979͒.