Sunteți pe pagina 1din 14

Gauge origin independent calculations of nuclear magnetic shieldings in relativistic

four-component theory
Miroslav Ilia, Trond Saue, Thomas Enevoldsen, and Hans Jrgen Aa. Jensen
Citation: The Journal of Chemical Physics 131, 124119 (2009); doi: 10.1063/1.3240198
View online: http://dx.doi.org/10.1063/1.3240198
View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/131/12?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Relativistic calculation of nuclear magnetic shielding tensor using the regular approximation to the normalized
elimination of the small component. III. Introduction of gauge-including atomic orbitals and a finite-size nuclear
model
J. Chem. Phys. 129, 224103 (2008); 10.1063/1.3028047
A fully relativistic method for calculation of nuclear magnetic shielding tensors with a restricted magnetically
balanced basis in the framework of the matrix DiracKohnSham equationa)
J. Chem. Phys. 128, 104101 (2008); 10.1063/1.2837472
Relativistic effects on nuclear magnetic shielding constants in H X and CH 3 X(X= Br,I ) based on the linear
response within the elimination of small component approach
J. Chem. Phys. 121, 6798 (2004); 10.1063/1.1787495
Calculation of nuclear magnetic shieldings. XIII. Gauge-origin independent relativistic effects
J. Chem. Phys. 110, 131 (1999); 10.1063/1.478089
Calculation of nuclear magnetic shieldings. XII. Relativistic no-pair equation
J. Chem. Phys. 108, 3854 (1998); 10.1063/1.475788

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

THE JOURNAL OF CHEMICAL PHYSICS 131, 124119 2009

Gauge origin independent calculations of nuclear magnetic shieldings


in relativistic four-component theory
Miroslav Ilia,1,a Trond Saue,2,b Thomas Enevoldsen,3 and Hans Jrgen Aa. Jensen3,c
1

Department of Informatics, Faculty of Management Science and Informatics, University of ilina,


Campus Prievidza, Bakalrska 2, SK-97101 Prievidza, Slovakia
2
Institut de Chimie de Strasbourg, CNRS et Universit de Strasbourg, Laboratoire de Chimie Quantique,
4 rue Blaise Pascal, F-67070 Strasbourg, France
3
Department of Physics and Chemistry, University of Southern Denmark, Campusvej 55,
DK-5230 Odense M, Denmark

Received 24 July 2009; accepted 9 September 2009; published online 29 September 2009
The use of perturbation-dependent London atomic orbitals, also called gauge including atomic
orbitals, has proven efficient for calculations of NMR shielding constants and other magnetic
properties in the nonrelativistic framework. In this paper, the theory of London atomic orbitals for
NMR shieldings is extended to the four-component relativistic framework and our implementation
is described. The relevance of London atomic orbitals in four-component calculations as well as
computational aspects are illustrated with test calculations on hydrogen iodide. We find that the use
of London atomic orbitals is an efficient method for reliable calculations of NMR shielding
constants with standard basis sets, also for four-component calculations with spin-orbit coupling
effects included in the wave function optimization. Furthermore, we find that it is important that the
small component basis functions fulfill the magnetic balance for accurate description of the
diamagnetic shielding and that the role of London atomic orbitals in the relativistic domain is to
provide atomic magnetic balance even in the molecular case, thus greatly improving basis set
convergence. The Sternheim approximation, which calculates the diamagnetic contribution as an
expectation value, leads to significant errors and is not recommended. 2009 American Institute of
Physics. doi:10.1063/1.3240198
I. INTRODUCTION

For properties depending on external magnetic fields, the


introduction of the physically motivated London atomic orbitals LAOs,1 also called gauge including atomic orbitals
GIAOs,2 ensures that the calculated properties are independent of the choice of the gauge origin.310 The motivation for
the present work stems from the physical fact that the LAOs
for nonrelativistic one-electron atoms correctly describe the
first-order changes induced by external magnetic fields,11
thus reducing basis set requirements for a given level of accuracy. One would expect a similar favorable behavior of
LAOs also in the rigorous relativistic realm. Most of the
LAO/GIAO-based methods have been developed and established in a nonrelativistic framework. They include electron
correlation as well using multiconfiguration self-consistent
field MCSCF,12 MllerPlesset MPn, n = 2 4,1315
coupled-cluster CC,16,17 and density functional theory
DFT1822 approaches.
However, relativistic effects play an important role for
systems containing heavy elements.23 They affect the calculated molecular properties, including NMR shielding
constants.24,25 Various theoretical treatments of relativistic
effects have been used for the prediction of magnetic propa

Author to whom correspondence should be addressed. Electronic mail:


ilias@utcpd.sk.
b
Electronic mail: tsaue@chimie.u-strasbg.fr.
c
Electronic mail: hjj@ifk.sdu.dk.
0021-9606/2009/13112/124119/13/$25.00

erties for recent reviews see Refs. 26 and 27. At the twocomponent relativistic level, thus allowing the inclusion of
spin-orbit interaction, schemes for the calculation of magnetic shielding constants in the common gauge origin CGO
have been presented based on the second-order
the
zeroth-order
regular
DouglasKrollHess,2833
approximation,34 the BreitPauli,3543 the normalized elimination of the small component NESC,4447 the
and
the
infinite-order
BaryszSadlejSnijders,33
two-component48 Hamiltonians. Relativistic two-component
methods for calculating NMR shielding tensor with LAOs
included were presented by Fukuda et al.49 and more recently by Hamaya et al.50 and by Autschbach.51 The latter is
based on the regular approximation to the NESC.
At the four-component level, Nakatsuji and co-workers
reported shielding constants calculated at the unrestricted
relativistic HartreeFock HF level with the external magnetic field as a finite perturbation.5254 This approach was
later extended to the four-component relativistic coupledcluster singles-and-doubles CCSD and configurationinteraction singles-and-doubles CISD level by adding the
nuclear magnetic moment as a second finite perturbation.55
The first implementation of four-component LAOs, albeit incomplete to be discussed below, is due to Quiney and
co-workers.5658 They published NMR shielding constants of
the water molecule at the uncoupled HF level. The first full
four-component coupled HF calculations of NMR shielding
tensors of hydrogen halides were reported by Visscher et

131, 124119-1

2009 American Institute of Physics

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-2

J. Chem. Phys. 131, 124119 2009

Ilia et al.

al.59 These NMR shielding constants were, however, obtained using the computationally simpler CGO method. The
same holds true for the more recent computational approaches in the four-component relativistic DFT framework
proposed by Liu and co-workers6062 and Komorovsk
et al.63
In this paper, we present the theory and the implementation of LAOs for NMR shielding constant calculations based
on the four-component DiracCoulomb Hamiltonian. It is an
extension of the four-component linear response method previously reported by two of us H.J.Aa.J and T.S.,64 which
utilizes quaternion algebra for efficient symmetry handling.
The first application of our implementation of the theory
presented here was to the xenon dimer and has been reported
in Ref. 65. In Secs. II and III, we present the theory and
details of the computational implementation. In Sec. IV we
report sample calculations of NMR shielding constants for
hydrogen iodide before concluding in Sec. V. Unless otherwise stated, we use in the following SI-based atomic units.

relativistic. It can be argued, however, that all magnetic interactions vanish in the strict nonrelativistic limit. In practice, that is, in conventional nonrelativistic calculations,
one allows the introduction of external magnetic fields B in
an ad hoc manner without any reference to the generating
mechanism, that is, magnetic induction, which is a truly relativistic effect. This allows the successful calculation of magnetic properties within a nonrelativistic framework. A more
complete discussion of this point is found in Ref. 68.
With the vector potentials given above, the Dirac Hamiltonian takes the form
hDB,mK = hD + B hB + mK hmK ,

where the Zeeman operator hB and the hyperfine operators


hmK appear
1
hB = rG c,
2

II. THEORY

1 rK c
hmK = 2
.
c
rK3

A. Hamiltonian

The point of departure is the one-electron Dirac Hamiltonian with perturbations arising from both a uniform external magnetic field B and internal nuclear magnetic moments
mK. These perturbations are introduced through the vector
potential A = AB + KAmK,
hDA = mc2 + + Vnuc = hD + c A,

according to the principle of minimal electromagnetic


coupling.66,67 The vector potential of a uniform external
magnetic field B is usually chosen in the gauge
ABr = 21 B rG,

rG = r RG ,

where RG is the selected gauge origin. The gauge for the


vector potential associated with a point magnetic dipole at
nucleus K is conveniently chosen as
AmKr =

1 mK rK
.
c2
rK3

The vector potentials AB and AmK are routinely employed in


nonrelativistic calculations and then introduced into the nonrelativistic Hamiltonian by minimal substitution, as in the
relativistic domain. Nonrelativistic results, however, cannot
be reproduced by a four-component relativistic code by letting the speed of light c go to infinity in practice, a very
large value since the vector potential AmK Eq. 3 then
goes to zero. This leads to the somewhat awkward situation
that the results obtained by conventional nonrelativistic calculations can only be obtained by keeping the speed of light
in AmK at its finite value, while letting it go toward infinity
elsewhere in the calculation. A more coherent picture
emerges by recognizing that conventional nonrelativistic calculations employ a nonrelativistic description of particles but
a relativistic coupling to external electromagnetic fields. The
nonrelativistic limit of electrodynamics cannot be explored
by experiment since the world at our disposal is inherently

In contrast to its nonrelativistic counterpart, the perturbed


Dirac Hamiltonian is linear in both the magnetic field B and
the magnetic dipoles mK. There is accordingly no explicit
diamagnetic term, although such a contribution can be extracted from the positive energynegative energy coupling
terms of the linear response function.69
The complete electronic Hamiltonian for use within the
BornOppenheimer approximation can be obtained in
second-quantized form by the introduction of field operators
=
H

1h 1
1d1

1
2

1
2g1,2
1
2d1d2.

The field operators will be expanded in a finite orthonormal


basis
1 = 1a ,

p
p

where a p is an annihilation operator. It is, however, known


that gauge origin invariance is usually not assured in a finite
basis approximation,6,7072 but the reference to an arbitrary
gauge origin RG can be removed by the introduction of
LAOs.37,9,10,73 Consider an expansion of a molecular fourcomponent spinor

p =

X cX p cp

X=L,L,S,s X

in scalar basis functions i.e., with only one nonzero component

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-3

L
0
L =
,
0
0

UMO
B = Bcq0.
q

L =

L
0

S =

0
0

M,XrM ,B = exp iABRM rM,XrM ,

10

X = L,L,S,S ,

1
hBM B = hD + B rG c M B
2

1
= exp iABR M r hD + B r M c M ,
2
11
showing that the reference to the gauge origin in the operator
rG is replaced with the reference to the center of the orbital
r M . In the LAO-based method, the LAOs Eq. 10 replace
the corresponding basis functions in the basis set expansion
in Eq. 8. It should be noted that, in contrast to the nonrelativistic domain, the relativistic LAOs by themselves do not
describe first-order changes in the atomic orbitals due to an
external magnetic field, as seen from the fact that the Zeeman operator Eq. 5 is no longer proportional to the orbital
angular momentum operator. However, the introduction of
LAOs removes reference to an arbitrary gauge origin in the
calculation of magnetic properties and significantly improves
basis set convergence, as will be shown in the following.
In order to ensure orthogonality of orbitals and thus the
same algebra of creation and annihilation operators at all
values of the magnetic perturbation, one must allow the coefficients cpB to be field dependent

The OMOs are connected to the UMOs through an orbital


connection matrix T.10,74 Several choices can be made for
the T, to be discussed in Sec. II C.
We write the B-dependent electronic Hamiltonian in
second-quantized form as
B,m = h B,m aa
H
K
D;pq
K p q
pq

where the extended superscript M , X designates that


M,XrM is an atomic scalar basis function of type X centered at position RM . Operating with the Dirac operator on a
London orbital M of any typeL, L, S, or Syields

pB = BcpB

14

The definition of a magnetic field dependent LAO employed


within the four-component framework is then

It is obvious that the UMOs are not orthonormal when the


magnetic field is turned on. However, one can always define
a set of orthonormalized molecular orbitals OMOs from
the UMOs as

OMO
B = UMO
BTqpB.
p
q

13

S =

J. Chem. Phys. 131, 124119 2009

Gauge origin independent calculations

12

since the overlap matrix over LAOs is field dependent as


well. However, it is more convenient to determine a set of
expansion coefficients at the zero magnetic field limit since
the LAOs in this case reduce to the standard atomic basis
functions. With this set of coefficients, one can construct
unmodified molecular orbitals UMOs

1
gBpqrsaparasaq ,
2 pqrs

15

where the tildes on the matrix elements indicate that the field
operators have been expanded in the B-dependent OMOs.

B. Linear response with perturbation-dependent


orbitals

In this section we first review a general expression for


the linear response function in a perturbation-dependent basis. This allows us to introduce notation and the formalism
we have chosen to use. Our approach follows closely the
derivations in Refs. 64 and 67.
Static molecular properties can be defined as derivatives
of the electronic energy with respect to perturbation
strengths, collected in the vector , in the zero-field limit.
The total first derivative of the energy with respect to a perturbation strength is

E E i
dE
=
+
.
dA A i A

16

We restrict attention to fully variational methods and employ


variational perturbation theory assuming that the energy is
optimized with respect to all variational parameters , that is,

E
=0

17

at all field strengths such that the second term of Eq. 16 is


identically zero. Taking the second derivative of the energy
with respect to perturbation strengths, we obtain

d 2E
2E i
2E
d E
=
=
+
,
dBdA dB A
i A B B A

18

which involves the generalized property gradient 2E / Ai


as well as first-order amplitudes i / B. The latter quantities are obtained from the first-order response equation

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-4

J. Chem. Phys. 131, 124119 2009

Ilia et al.

d E
2E
2E i
=
+
= 0,
dB i
B i i j B

19

where appears again the generalized property gradient as


2
well as the electronic Hessian E2
0,ij = E / i j. We have
deliberately chosen expressions that are asymmetric in the
perturbation strengths A and B since only the response in
terms of the latter is needed and which therefore may be
computationally advantageous. Symmetric expressions are
easily obtained and may provide higher numerical
precision.75,76
In this paper, we focus on NMR shielding constants calculated at the four-component relativistic closed-shell HF
level using LAOs. The individual components of the NMR
shielding tensor for nucleus K are calculated within the
BornOppenheimer approximation as a second derivative of
the electronic energy
K

d 2E
dmK,dB

20

mK=0,B=0

at zero perturbation strength. We employ a unitary exponential parametrization of a closed-shell relativistic 4c or 2c HF


single determinant wave function
SD = exp SD0.

21

We will in the following use indices i , j , k , l for occupied


orbitals, indices a , b , c , d for virtual orbitals, and indices
p , q , r , s , t for general indices. The anti-Hermitian orbital rotation operator can then be expressed in nonredundant
variational parameters as

aiaaai

ai

* aa ,
ai
i a

= OMO exp
p
qp
q
= rUMOBTrqBexp qp

23

rq

= Bci + qTBqi aBai . 24

mK=0,B=0

= Bi hmAi + ihmABi .

25

Inserting the first-order OMO of Eq. 24, we obtain our


working expression
A,B .
A = hmA ;hB0 + hm
ii

26

The first term is recognized as a linear response function


A;B0 = EA1XB,

XB =

xB
,
x B*

B
xBai = ai

27

obtained by contracting the gradient of the property A by the


first-order response vector XB. The latter is obtained by solving the first-order response equation Eq. 19 which can be
written more compactly as
1
E2
0 XB = EB .

28

The Hessian E2
0 as well as the gradients associated with
properties other than a uniform magnetic field have their
usual form see, for example, Refs. 64 and 67 for explicit
expressions. We follow these references and use the following general form for a property gradient in relativistic fourcomponent linear response:
EX1 =

gX
,
g X*

gXai =


EX
*
ai

X0 = FX ,
= 0 ai aa,H
ai
=0

29
where the first-order derivative with respect to property X of
the Fock matrix in OMO basis appears
FX = hX + L
X ,
pq
pq
pq,jj

X = gX gX .
L
pq,rs
pq,rs
ps,rq

30

FB = hB + LB + TB,h + TB,L ,
0 pq
pqjj
pq
pqjj
pq
j

affording the first-order derivative of an occupied orbital


with respect to the external magnetic field B from Eq. 13,

B=0

d 2E
dBTdmA

In the particular case of a uniform magnetic field, the property gradient is extracted from

22

where the tildes again denote that creation and annihilation


operators refer to the B-dependent OMOs Eq. 14. The
corresponding expression for a general transformed OMO is
using Eq. 14

B di
i
dB

A =

The last term is a relaxation term, whereas the first and second terms together constitute a static term consisting of the
direct GIAO contribution and a reorthonormalization term,
respectively.
The first derivative of the HF energy at any field strength
with respect to a nuclear magnetic moment mK is the expectation value of the hyperfine operator hmK Eq. 5 in terms
of the OMOs at the selected field strength. The shielding
tensor is therefore the first-order derivative of this expectation value with respect to the external magnetic field B

31

where we have introduced one-index transformed quantities


T,A pq = T*ptAtq + A ptTtq,
t

T,a pqrs = T*ptatqrs + a ptrsTtq + Trt* a pqts + a pqrtTts,

32

corresponding to reorthonormalization terms for a oneelectron and a two-electron operator, respectively.10 The
property gradient associated with the nuclear magnetic moment evaluates to
mK
mK
= hai
.
gai

33

The second term of the shielding tensor in the LAO formalism Eq. 26 is an expectation value term

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-5

J. Chem. Phys. 131, 124119 2009

Gauge origin independent calculations

i hmii ,B = i hmii ,B + TB,hm ii,


K

34

corresponding to the second-order derivative of matrix elements of the Dirac Hamiltonian Eq. 4 in OMO basis with
respect to the external magnetic field and the selected nuclear
magnetic moment. In the nonrelativistic domain, this term
also contains an explicit diamagnetic contribution, which is
absent in the relativistic domain since the Dirac Hamiltonian
Eq. 4 is linear in the nuclear magnetic moments as well as
the external magnetic field. Explicit expressions for the
atomic matrix elements for UMO appearing in Eqs. 31,
33, and 34 are given in the Appendix, whereas explicit
expressions for the first-order connection matrix TB will be
discussed in Sec. II C.
Our working expression, Eq. 26 can be compared to
the previous expressions employed by Quiney and
co-workers5658 in their calculation of the NMR shielding
tensor for the water molecule at the four-component relativistic HF level with LAOs. Quiney et al. calculated the linear
response function at the uncoupled HF level and without
inclusion of the LAO contributions to the property gradient
Eq. 31. Curiously, they employed the opposite sign of the
vector potential in their definition of the four-component
LAO Eq. 10. This does not remove reference to the arbitrary gauge origin in coupled HF calculations and furthermore changes the sign of the expectation value term Eq.
34 of which they calculated only the first one and not the
reorthonormalization term.
A first particular feature of four-component relativistic
calculations is the presence of negative-energy positronic
orbitals. The response of the HF wave function to a pertur++
bation is therefore not only carried by orbital rotations ai
between virtual and occupied positive-energy orbitals but
+
also by orbital rotations ai
involving the virtual negativeenergy orbitals. In the nonrelativistic limit the contribution
from the latter rotations reduces to the diamagnetic
contribution.69
A second feature to consider is that the coupling between
the large and the small components is modified by the introduction of an external vector potential. In our approach to
four-component relativistic calculations, the small component basis set is generated using the kinetic balance relation.
In this work, we have studied two schemes for the kinetic
balance relationthe unrestricted kinetic balance UKB
Ref. 77 and the restricted kinetic balance RKB.78 The
RKB scheme gives a 1:1 ratio between the sizes of the large
and small component basis sets. On the other hand, starting
from a large component scalar basis function of a given
value of the orbital angular momentum l, the extra small
component basis functions generated by the UKB scheme
include, for the atomic problem, the functions needed for
magnetic balance associated with an external magnetic
field, whereas such functions are generated by the RKB
scheme only by inclusion of large component basis functions
of l + 2.62,69,79,80 In this way, UKB provides a better basis set
for the calculation of magnetic properties,81 albeit at the increased risk of introducing linear dependencies.
Alternatively, one may take magnetic balance explicitly

into account in four-component relativistic response calculations of magnetic properties. This approach has been pursued
by Liu and co-workers6062 and Komorovsk et al.63 and
leads to the introduction of perturbation-dependent basis
functions, albeit for solving a different problem than the
gauge origin problem addressed in the present paper. A different approach has been proposed by Kutzelnigg,82 whereby
the contribution from the negative-energy orbitals to the
shielding constant is reduced from Oc0 to Oc4 through a
unitary transformation, involving the full vector potential, of
the Dirac operator. The formulation furthermore leads to perturbation operators similar to those found in the nonrelativistic domain. Numerical studies by Visscher83 and Xiao and
co-workers60,61 have revealed numerical instabilities in the
original formulation, dubbed full field-dependent unitary
transformation FFUT by the latter authors. This problem is
avoided by restricting the unitary transformation to the vector potential associated with the external field only, leading
to the external field-dependent unitary tranformation
EFUT, although the negative-energy orbitals now contribute to Oc2. However, drawbacks of both methods are that
gauge transformations become more complicated and that
the unitary transformation does not commute with the Gaunt
two-electron interaction.84
C. Explicit and implicit orbital connections

In the previous section, we have developed expressions


for the NMR shielding tensor at the four-component relativistic HF level using a second-quantization formalism and
magnetic field dependent London orbitals. In this section we
discuss explicit choices of the matrix T connecting UMOs
and OMOs Eq. 14. Such choices are implicitly made also
when deriving response theory with perturbation-dependent
orbitals in a first-quantization formalism, but often without
using the terminology employed here. We will therefore elucidate these connections in the present section.
From the orthonormality of the OMOs, it follows that
the Hermitian part of the first-order connection matrix is
equal to the first-order overlap matrix in terms of the UMOs
TB + TB = SB = WB + WB,

35

where we have introduced the matrix


WBpq =

d
pUMO
BB = pBcq0.
q
dB

36

The symmetric connection sc minimizes the difference between the field-dependent OMOs and the corresponding
UMOs at the same magnetic field strength and leads to TB
= 21 SB, which is independent of the chosen gauge origin.
The natural connection nc, on the other hand, minimizes
the difference between the field-dependent OMOs and the
corresponding zeroth-order orbitals and leads to gauge origin
dependent TB = WB.85 However, insertion into Eq. 24
shows that in the nc and in the limit of a complete basis the
reorthonormalization term cancels exactly the direct GIAO
contribution. Moreover, again shown starting from Eq. 24,
in any basis the nc assures that the overlap between any
occupied zeroth- and first-order orbitals is zero, that is,

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-6

J. Chem. Phys. 131, 124119 2009

Ilia et al.

jBi = 0

37

and is thus equivalent to the parallel-transport gauge of Ref.


86. No such simple relations hold in the sc10 or any other
connection; in general, reorthonormalization terms mix with
relaxation terms. Although all connection matrices are, in
principle, equivalent within the finite basis approximation,
these special features of the nc means that it is numerically
more stable because it does not contain large values in the
reorthonormalization terms which must be canceled by the
relaxation terms.
First-quantization approaches typically see for instance
Ref. 87 start directly from field-dependent orbitals of the
form of Eq. 12. First-order perturbed orbitals are expressed
as
B = Bc + cB = Bc + A ,
i
q qi
i
i
i

38

where the final form is obtained by expressing the first-order


expansion coefficients cBi in terms of the zeroth-order ones
and a matrix A of orbital mixing coefficients. Equation 38
may be compared with Eq. 24 from which ones sees that
the occupied-occupied block of the orbital mixing matrix
maps into the corresponding block of the connection matrix
T, whereas the virtual-occupied block in addition contains
relaxation terms, that is,
A ji = TBji,

B
Aai = Tai
Bai .

39

The virtual-occupied block of the orbital mixing matrix is


found from response equations, whereas the occupiedoccupied block is found from the first-order orthonormality
condition Eq. 35. Several choices are thus possible for the
occupied-occupied block. For instance, in Ref. 87 the sc is
chosen. However, the actual choice is arbitrary, as shown by
the following argument: Inserting Eq. 38 into Eq. 25 and
reindexation in the second term gives

A = c* iBhmAi + ihmABci

40

+ A*ji + Aij jhmAi

41

* h + h A .
+ Aai
a mA i
i mA a ai

42

i,

ij

ia

Combined with the first-order reorthonormalization condition Eq. 35, this clearly shows that the term involving the
occupied-occupied orbital mixing matrix is independent of
the chosen connection. A further important conclusion is that
since the first term and the total property are clearly connection independent, this must also then hold true for the third
term involving the occupied-virtual block. Compared with
Eq. 39, we then see that if restrictions are introduced on the
virtual orbitals employed in the response calculation, they
must be imposed on the connection matrix as well. In relativistic calculations, rotations +
ai between virtual negativeenergy orbitals and occupied positive-energy ones may be
suppressed and approximated by an expectation value corresponding to the diamagnetic term in the nonrelativistic limit.

In such cases, reorthonormalization terms involving the


negative-energy orbitals should be suppressed as well to assure connection independence of the total property.
These negative-energy reorthonormalization contributions can, on the other hand, also be suppressed when rotations +
ai are included without introducing large errors. This
can be seen from the following argument: The block of the
Hessian matrix corresponding to pairs of such rotations will
be dominated by orbital energy differences on the order of
2mc2 along the diagonal so that the corresponding firstorder orbital rotations can be approximated by

+;B

ai

FB
pq
,
2mc2

43

where appears the property gradient of Eq. 31. The reorthonormalization contribution to the property gradient can be
reformulated as
TB,Fai = Trt* Laits + LairtTts.

44

The part involving the Fock matrix F can in the case of


virtual negative-energy canonical orbitals be approximated
by
B

B
B
2 B
*
TB,F+
ai = Tai i + aTai iS 2mc Tai ,

45

where we have used the first-order orthonormality condition


Eq. 35, and a 2mc2 + i. Inserted into the approximate
expression for the first-order orbital rotation Eq. 43, we
see that the dominant reorthonormalization contribution to
+;B
will be TBai, which cancels the corresponding reorai
thonormalization term of the first-order orbital in Eq. 24.
III. COMPUTATIONAL DETAILS

The integral module HERMIT Ref. 88 of the DIRAC04


Ref. 89 program suite was extended in order to calculate all
matrix elements appearing in Eqs. 31 and 34. Twoelectron magnetic field derivatives of the Coulomb operator
contributing to the magnetic field property gradient Eq.
31 are implemented as an effective part of the Fock-like
operator. After the SCF optimization,77,78,90 the property gradient elements generated by Eqs. 31 and 33 are transformed to the MO basis and delivered to the linear response
module64 of DIRAC04. We have performed relativistic calculations both with and without the + rotations in the linear
response function; in the latter case, the diamagnetic contribution was calculated as the expectation value of the nonrelativistic diamagnetic operator. The approximations involved
in this are described in Ref. 69.
For the reference Kramers restricted HF calculations we
have used four different models. We have used two different
schemes for the kinetic balance relationthe UKB Ref. 77
and the RKB.78 For the DiracCoulomb four-component
spin-free SF calculations, the elimination of spin-orbit effects in the implementation is identical to the previous work
by one of us.68 As a check of our implementation, we have
also performed nonrelativistic calculations on NMR shieldings in both the CGO- and LAO-based approaches. This was
done by using the four-component LvyLeblond LL

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-7

J. Chem. Phys. 131, 124119 2009

Gauge origin independent calculations

TABLE I. Basis sets used in the present work.


Basis set
I DZ
IIb aug-DZ
IIIc TZ
IVd aug-TZ
Ve
VIf
a

16s12p7d
17s13p7d
21s17p12d1f
22s18p13d2f
27s21p15d4f
28s22p17d5f3g2h

4s1p
5s2p
5s2p1d
6s3p2d
11s6p3d
11s6p3d

Decontracted sp-pVDZ basis for I Ref. 100, and decontracted cc-pVDZ


basis set for H Ref. 101.
b
Decontracted sp-a-pVDZ basis set for I Ref. 100, and decontracted augcc-pVDZ basis set for H Ref. 101.
c
Decontracted sp-pVTZ basis set for I Ref. 100, and decontracted cc-pVTZ
basis set for H Ref. 101.
d
Decontracted sp-a-pVTZ basis for I Ref. 102, and decontracted aug-ccpVTZ basis set for H Ref. 101.
e
Basis sets taken from Ref. 59.
f
Basis sets taken from Ref. 103.
a

Hamiltonian.78 We have compared the results to the corresponding calculations with the standard scalar nonrelativistic
Hamiltonian in the DALTON12,91 program suite, and the results
are identical within numerical accuracy.
The aim of these test calculations is to illustrate and to
emphasize the computational advantages of the use of LAOs
in connection with the rigorous DiracCoulomb based relativistic theory rather than to present very accurate values of
NMR shieldings. This would require inclusion of electron
correlation and is beyond the scope of the present endeavor.
The hydrogen iodide molecule in its experimental geometry 3.040 a.u. Ref. 92 was chosen for this study because
it contains a heavy element where one would expect that
relativistic effects are significant.
NMR shielding constants have been calculated with several decontracted basis sets selected from the literature to
form an approximative sequence of increasing size. They are
shown in Table I. Four gauge origin positions were tested for
the hydrogen iodide molecule coordinates in a.u.: 1 on the
iodine atom 0,0,0 for compatibility with the previous CGO
calculations of Visscher et al.59 2 at the center of mass,
0,0,0.023 952, the typical choice for CGO calculations, 3
on the hydrogen atom 0,0,3.040, and 4 7 a.u. from the
iodine atom 0,0,7. The purpose of the last two positions are
to simulate the effects of LAOs on larger molecules, where
one cannot place the gauge origin close to all atoms. For the
relativistic calculations, we applied a Gaussian finite nucleus
model,93 but in the nonrelativistic LL mode we used the
point nucleus model in order to compare directly to
previous78 and to verifying calculations with the DALTON12,91
program system. All relativistic calculations were done with
the LL LL and SS LL classes of two-electron integrals,
while the SS SS and SL SL Gaunt classes were omitted
L and S denoting large and small component basis functions, respectively, cf. Eq. 9. The used speed of light was
137.035 999 8 a.u. All calculated values of NMR shieldings
are in ppm.
IV. RESULTS AND DISCUSSION

Calculated four-component results for the NMR shielding constants of the HI molecule are collected in Tables II

and III, and in Table IV we report individual contributions to


the final results for selected calculations. We compare four
types of four-component HF wave functions: nonrelativistic
LL, SF DiracCoulomb, as well as regular DiracCoulomb,
the latter with RKB, and UKB the first two approachesLL
and SFimplies RKB. We consider CGO calculations with
four different locations of the gauge origin as well as LAO
calculations with two different orbital connection schemes,
the nc and the sc. We also consider the effect of omitting the
T+ reorthonormalization terms Eq. 14, involving the
negative-energy orbitals.
We note that the nonrelativistic results obtained with the
LL Hamiltonian in basis set V are the same or better, within
rounding/truncation error bars as those given in Table 1 of
Ref. 59. Regarding our relativistic RKB-CGO results in the
same basis set, these are very close to those published in Ref.
59, where all three classes of two-electron integrals were
included. This supports our assumption that the neglect of
the SS SS two-electron integrals in the four-component calculations does not introduce significant errors for the purpose
of the present work.
Since we are in the four-component relativistic domain,
specific features are appearing in connection with LAOs, in
particular, related to the small component basis functions. As
shown in Eq. 14, the orthonormal molecular orbitals are
expanded in the complete set, which for four-component calculations means both positive energy also called electronic or e UMOs and negative energy also called
positronic or p UMOs. We recall that the diamagnetic
shielding in the full four-component formalism is obtained
from the negative energypositive energy linear response solution, which in the nonrelativistic limit becomes
the familiar expectation value expression.69
To gain more insight in the evaluation of the diamagnetic
shielding, we also report some calculations with only the ++
rotations in the linear response function. For these calculations, as well as for all LL and SF calculations, the diamagnetic term is approximately calculated as the expectation
value of the nonrelativistic diamagnetic operator,69,94 with
the same London orbital modifications as in nonrelativistic
implementations.12
The results in Tables II and III for four different gauge
origins clearly illustrate the great advantages of using LAOs.
In particular, one notes that the calculated CGO shieldings
for hydrogen are approximately a factor of 5 too big when
the gauge origin is placed on the hydrogen, and even worse
for the most distant 0,0,7 gauge origin. The results when
the gauge origin is on the iodine atom are reasonable, and
this explains why the four-component results by Visscher
et al.59 and the two-component results by Hamaya et al.50 are
also reasonable. However, for larger molecules we can
clearly not select a gauge origin where the results will be
acceptable for all atoms in a reasonable size basis sets. This
at first surprising result that the hydrogen shielding is so
badly described with the gauge origin on hydrogen can be
understood when one realizes that most of the shielding at
the hydrogen nucleus the proton is caused by the iodine

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-8

J. Chem. Phys. 131, 124119 2009

Ilia et al.

TABLE II. Isotropic proton NMR shielding constants in ppm of hydrogen iodide calculated with nonrelativistic LL, SF, RKB-relativistic, and UKB-relativistic methods; each both with CGO on GO and with LAOs.
The basis sets are specified in Table I. The difference of each value to the corresponding value in the largest
basis set VI is given in parentheses. The experimental value for the 1H shielding constant is 43.86 ppm after
converting the gas-phase shift from Ref. 104 to absolute shielding using CH4 = 30.61 ppm from Ref. 105.

UKB-CGO

GO I
GO CM
GO H
GO 0,0,7

UKB-LAO
UKB-LAOa
RKB-CGO

RKB-CGOb

GO
GO
GO
GO
GO

I
CM
H
0,0,7
I

GO CM
GO H
GO 0,0,7
RKB-LAO
RKB-LAOa,b
SF-CGOb

GO I

SF-LAOa,b
LL-CGOb
LL-LAOb

GO I

II

III

IV

VI

44.03
4.00
45.36
212.58
432.14
45.89
3.45
45.88

44.03
4.00
45.37
213.14
433.43
46.31
3.03
46.31

45.25
2.78
46.46
199.31
400.00
47.03
2.31
47.03

46.09
1.94
47.28
196.85
393.23
47.80
1.54
47.80

47.87
0.16
48.71
153.73
291.61
49.41
0.07
49.40

48.03

37.17
37.93
134.21
260.63
44.69
3.37
46.06
218.44
444.77
39.36
47.19
2.02

39.19
39.96
136.38
262.97
44.25
3.81
45.61
217.18
442.44
40.92
46.49
2.72

42.93
43.68
139.45
265.19
46.16
1.90
47.41
204.24
410.17
44.63
47.93
1.28

45.46
46.21
140.59
264.50
46.33
1.73
47.54
200.50
401.34
47.12
47.92
1.29

47.30
47.84
115.96
205.39
47.99
0.07
48.85
157.12
299.28
48.79
49.27
0.06

29.11
1.84
31.19
0.14

29.58
1.37
31.22
0.17

30.39
0.56
31.10
0.05

30.69
0.26
31.13
0.08

30.91
0.04
31.07
0.02

29.20
1.88
31.32
0.13

29.68
0.40
31.37
0.18

30.53
0.55
31.25
0.06

30.83
0.25
31.28
0.09

31.05
0.03
31.20
0.01

48.68
129.97
236.71
49.34
49.33
47.62
48.03
99.75
167.65
48.06
48.73
133.23
244.17
48.88
49.21

30.95
31.05

31.08
31.19

a
The negative energypositive energy T+ part of the reorthonormalization is neglected holds for UKB-LAO,
RKB-LAO, and SF-LAO approaches, otherwise is present.
b
Diamagnetic term calculated as expectation value of nonrelativistic diamagnetic operator. Otherwise obtained
from the linear response function.

electrons and not by the single hydrogen electron; the


proton is quite deep into the electronic cloud of the iodine
atom.
It should be noted, as follows from Eq. 11, that the
introduction of LAOs in calculations employing explicit
magnetic balance will also correctly shift the gauge origin
appearing in the vector potential appearing in the magnetic
coupling between a large and small component LAO. This is
also a key to an understanding of the improved basis set
convergence of LAOs in the present calculations. Since the
gauge origin appearing in the magnetic coupling is shifted to
the atomic origin of the large component basis functions, the
extra small component basis functions needed for magnetic
balance correspond to those of the atomic problem even in a
molecular calculation. As already discussed in Sec. II B,
starting from an atomic large component scalar basis function of orbital angular momentum l, the necessary small
component basis functions are automatically generated by
UKB, whereas large component basis functions of orbital
angular momentum l + 2 must be included to assure magnetic

balance with RKB, and so we expect considerably slower


basis set convergence with the latter approach.
As discussed in Sec. II C, the orbitals included in the
reorthonormalization terms should be the same as in the orbital rotation relaxation terms. We find that when the T+
contributions are neglected in full LAO response calculations, one gets unphysical results in the nc in the sense that
calculated NMR shielding constants vary for different positions of the gauge origin. Such differences are not observed
with the sc because each individual term is gauge independent. On the other hand, when the diamagnetic term is calculated from the +-rotations, using either the full reorthonormalization terms T++ and T+ or leaving out the T+
contributions, practically identical values are obtained, as
one can see in Tables II and III for the UKB-LAO approach.
This at first somewhat surprising observation is however in
accordance with the discussion in the last part of Sec. II C.
In Table IV we display individual contributions to the
NMR shielding constants for both LAO- and CGO-based
approaches. These data demonstrate the numerical variability

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-9

J. Chem. Phys. 131, 124119 2009

Gauge origin independent calculations

TABLE III. Isotropic 127I iodine NMR shielding constants in ppm of hydrogen iodide calculated with nonrelativistic LL, SF, RKB-relativistic, and UKB-relativistic methods; each both with CGO on GO and with
LAOs. The basis sets are specified in Table I. The difference of each value to the corresponding value in the
largest basis set VI is given in parentheses.

UKB-CGO

GO I
GO CM
GO H
GO 0,0,7

UKB-LAO
UKB-LAOa
RKB-CGO

RKB-CGOb

GO
GO
GO
GO
GO

I
CM
H
0,0,7
I

GO CM
GO H
GO 0,0,7
RKB-LAO
RKB-LAOa,b
SF-CGOb

GO I

SF-LAOb
LL-CGOb
LL-LAOb

GO I

II

III

IV

VI

5673.0
187.4
5675.2
5947.9
6306.1
5905.2
2.0
5905.2

5808.9
51.5
5809.0
5822.2
5839.5
5961.4
54.2
5961.4

5817.8
42.6
5819.1
5977.1
6184.6
5951.6
44.4
5951.6

5852.0
8.4
5852.1
5868.2
5889.4
5921.7
14.5
5921.7

5854.4
6.0
5854.4
5853.2
5851.7
5960.3
53.1
5959.4

5860.4

4753.3
4755.4
5029.3
5388.8
6556.9
217.9
6559.1
6829.8
7185.3
4975.3
6767.6
43.6

4890.6
4890.8
4903.6
4920.4
6690.1
84.7
6690.2
6701.5
6716.3
5027.3
6811.7
0.5

5239.4
5240.7
5399.9
5608.9
6732.9
41.9
6734.1
6890.2
7095.2
5367.1
6845.8
34.6

5278.7
5278.9
5295.7
5317.7
6765.1
9.7
6765.2
6779.5
6798.2
5343.9
6820.1
8.9

5322.8
5322.8
5321.7
5320.2
6769.5
5.3
6769.5
6766.4
6762.4
5422.1
6848.6
37.4

5378.2
5378.2
5377.0
5375.4
6774.8

6295.6
248.2
6465.7
78.2

6472.0
71.8
6509.7
34.2

6490.0
53.8
6540.9
3.0

6540.0
3.8
6542.3
1.6

6540.0
3.8
6540.2
3.7

6543.8

4390.4
148.6
4536.5
2.5

4538.6
0.4
4573.3
43.3

4506.4
32.6
4549.8
10.8

4546.8
7.8
4548.9
9.9

4541.3
2.3
4541.5
2.5

4539.0

5860.4
5859.0
5857.2
5907.2
5908.4

6774.8
6771.6
6767.3
5423.2
6811.2

6543.9

4539.0

a
The negative energypositive energy part T+ of the reorthonormalization is neglected holds for UKB-LAO,
RKB-LAO, and SF-LAO approaches, otherwise is present.
b
Diamagnetic term calculated as expectation value of nonrelativistic diamagnetic operator. Otherwise obtained
from the linear response function.

of individual terms, which is enhanced due to the use of two


different connections, as mentioned in the text above. In particular, we note how the reorthonormalization and paramagnetic terms grow uncontrollably large and with opposite sign with increasing basis set size in the symmetric
connection, which may lead to problems with numerical
accuracy.95
The values for the isotropic proton shielding constant
calculated in the best basis VI with UKB-LAO and RKBLAO, the latter with or without the diamagnetic contribution approximated by an expectation value, agree to
within 0.5 ppm Table II, whereas there is a spread of almost 1400 ppm in the corresponding numbers for iodine
Table III. In order to judge which value is best, we carefully analyze UKB and RKB calculations of the iodine
shielding with the CGO placed at the iodine nucleus. We
analyze, in particular, the validity of the individual steps,
leading to the calculation of the diamagnetic contribution as
the expectation value of the nonrelativistic diamagnetic
operator.69 The full RKB-CGO calculation in basis VI gives
I = 5378.2 ppm, of which the paramagnetic and dia-

magnetic parts are 973.9 and 4404.4 ppm, respectively. The corresponding UKB-CGO numbers are 970.8 and
4889.6 pm, showing that the RKB and UKB paramagnetic
values are very close to each other, and that the major discrepancy arises from the diamagnetic contribution. This illustrates that the magnetic balance is most important for the
diamagnetic type terms from the +-rotations. This is in
agreement with the general observation that Dyalls modified
Dirac equation,78,96 which makes the number of negativeenergy orbitals equal to the number of positive-energy ones
and in fact defines our RKB scheme, does not change the
quality of the positive-energy orbitals relative to UKB. As an
example of this, the converged HF energy in basis I is
7115.7787 hartrees with RKB and 7115.7768 hartrees
with UKB. For all basis sets, the difference between the
RKB and UKB total energy is less than 0.0001 hartree.
Next, solving the response equation Eq. 19, separately for ++- and +-rotations gives RKB-CGO I
= 5379.4 ppm= 975.4 ppm paramagnetic 4404.0 diamagnetic, which means that ignoring the coupling of the
two classes of rotations, which are on the order of Oc1, in

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-10

J. Chem. Phys. 131, 124119 2009

Ilia et al.

TABLE IV. Individual parallel and perpendicular contributions to NMR shielding constants of HI. For LAO only: scsymmetric connection,
ncnatural connection. For CGO calculations GO I. Used basis sets I and VI.
RKB-LAO
Basis

UKB-LAO
H

RKB-CGO

UKB-CGO

1.26
0.23
13.85
1932.53
48.80
51.39

5798.33
5800.58

0.35
0.23

4916.40
4890.85

0.70
0.24

5795.80
5797.14

51.45
51.47

4913.57
4886.99

48.80
51.39

48.18
50.22
31.92
2020.10
2.91
3.58

343.61
629.31

42.41
48.39

341.47
624.59

42.41
48.39

1591.06
1667.63

2.91
3.58

1589.75
1663.25

2.91
3.58

45.89
50.11
45.88
47.81

6141.94
6429.88
7386.86
7464.77

42.76
48.15
48.54
47.88

5257.87
5515.44
6503.32
6550.24

43.11
48.14
45.88
47.81

London

I
1.70
5.36
1.70
VI
3.42
5.73
3.42

I
0.00
0.00
0.00

VI
0.00
0.00
0.00
Reorthonormalizationb
,nc
I
10.54
19.35
2.35
,nc
VI
0.41
11.89
3.55
,sc
I
4.80
3.09
12.60
,sc
VI
10242.17
2058.42
12216.98
,nc and ,sc
I
0.00
0.00
0.00
VI
0.00
0.00
0.00
RKB: diamagnetic expectation value/UKB: diamagnetic response
,nc
I
5 789.49
12.33
4 947.89
,nc
VI
5 796.91
5.83
4 906.52
,sc
I
5 789.49
12.33
4 947.51
,sc
VI
5 796.91
5.83
6 883.51
,nc and ,sc
I
5 795.80
51.45
4 913.57
,nc and ,sc
VI
5 797.14
51.47
4 886.99
Paramagnetic response ee
,nc
I
656.17
48.18
654.13
,nc
VI
683.67
50.22
679.22
,sc
I
671.51
31.92
669.47
,sc
VI
10 926.27
2020.09
10 915.66
,nc and ,sc
I
1 591.06
2.91
1 589.75
,nc and ,sc
VI
1 667.63
3.58
1 663.25
Total

I
6 457.90
46.52
5 606.07

VI
6 484.43
49.88
5 585.62

I
7 386.86
48.54
6 503.32

VI
7 464.77
47.88
6 550.24

5.36
5.73
0.00
0.00
6.38
5.61
5.23
131.95
0.00
0.00

London part is the first term in Eq. 34.


Reorthonormalization part is the second term in Eq. 34.

the electronic Hessian gives a total error of only 1.2 ppm.


Calculating the diamagnetic part by uncoupled rather than
coupled HF also has a negligible effect in accordance with
Ref. 97: The contribution changes by 2.1 ppm to 4406.1
ppm. In an uncoupled formalism the response function
is given by

A =

i,a


1 rA c
i 2
c
rA3

a i

1
rG c i
2

showing that for finite value of the speed of light c, this is


quite a drastic approximation even though it has been
shown97 that only negative-energy orbitals within a fairly
limited interval contribute. The -response function can
now be expressed as

A =
,

46
where the summation over virtual orbitals a is restricted to
negative-energy ones. Following Sternheim,98 the diamagnetic expectation value expression is obtained by approximating the orbital energy difference a i by 2mc2 and
approximating the sum over negative-energy orbitals by a
resolution of identity. We consider these two approximations
separately. We find that the replacement of the orbital energy
difference by a common energy denominator changes the
diamagnetic contribution by 318.2 ppm to 4724.3 ppm,

2 i,p

1 rK
c2
rK3

1
1 rK
i 2
2 i,p+
c
rK3

1
rG i
2

1
rG i ,
2
47

where the first term has a sum over all orbitals, whereas the
sum in the second term is restricted to all positive-energy
orbitals. In the second term the Dirac matrices of the magnetic operators couple the large and small components of
positive-energy orbitals, and this term therefore goes to zero
in the nonrelativistic limit with the small components. The
Sternheim approximation therefore becomes exact in the
nonrelativistic limit. However, ignoring the second term at
the finite value of the speed of light changes the diamagnetic

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-11

J. Chem. Phys. 131, 124119 2009

Gauge origin independent calculations

contribution by 560.0 ppm to 5284.3 ppm. All in all the


Sternheim approximation gives a total error of 878.2 ppm. A
second source of error is basis set incompleteness. Assuming
resolution of identity in the first term of Eq. 47, which
calculates the diamagnetic term as an expectation value,
gives 5799.4 ppm. There is thus a basis set incompleteness
error at the RKB level in basis set VI of 515.1 ppm. In basis
set I the error grows to a staggering 1029.5 ppm. In comparison, the corresponding basis set incompleteness errors with
UKB are 34.7 and 7.0 ppm in basis sets I and VI, respectively. The difference between UKB and RKB basis set completeness can mainly be attributed to the need to include
large component g-functions to obtain magnetic balance for
the occupied iodine d atomic orbitals, in accordance with the
discussion above and in Refs. 62, 79, and 80. We can therefore conclude that the UKB-LAO shielding value is the recommended one and that the difference in the full response
functions with UKB and RKB is due to significant basis set
incompleteness in the latter one, even using the best basis set
VI. Indeed, the difference between the diamagnetic contributions with UKB 4889.6 ppm and RKB 4404.4 ppm is
almost exactly the basis set incompleteness error. It should
furthermore be noted that the Sternheim approximation gives
approximately the same significant error with UKB as with
RKB and can therefore not be recommended. The UKB and
RKB expectation values are almost identical to third decimal in basis set VI, once again showing that the positiveenergy orbitals with RKB and UKB are very similar.
Generally, the convergence of NMR shielding constants
with respect to the basis set is smoother with LAOs than with
standard atomic orbitals CGO approach. This holds for the
nonrelativistic LL calculations, as seen many times before
with nonrelativistic one-component Hamiltonians,12 and for
the relativistic four-component DiracCoulomb RKB and
UKB calculations, which is the main contribution of the paper. The effect of LAOs on the calculated values can be
compared with available two-component data. We observe,
as Hamaya et al.,50 an increase in the isotropic shielding
constants of I and H in hydrogen iodide by employing LAOs.
At the SF level these differences are almost negligible in the
largest basis set. This confirms that it is much easier to get
basis set converged results without spin-orbit relativistic effects. On the other hand, the SF calculations imply that RKB
and the iodine shielding are probably too large by about 500
ppm in basis set VI due to basis set incompleteness error.
V. CONCLUSIONS

A scheme for GIAOs for use in four-component relativistic calculations of second-order magnetic properties has
been implemented and tested. The same good convergence as
in nonrelativistic calculations of calculated NMR shielding
constants with basis set size was observed for the London
orbital perturbation dependent basis sets. It can be rationalized by the fact that the introduction of LAOs in the relativistic domain provides atomic magnetic balance even for the
molecular problem. The extra small component functions
then needed for magnetic balance can be obtained by taking
magnetic balance explicitly into consideration. For an exter-

nal uniform magnetic field, they are also straightforwardly


obtained by UKB, but generally require an extension of the
large component basis for RKB such that the former approach is preferred. We have also implemented LAOs for
four-component relativistic calculations of magnetic susceptibilities and will report this in a separate paper. Our analysis
shows that the Sternheim approximation, which calculates
the diamagnetic contribution as an expectation value, is a
poor approximation and is therefore not recommended.
An obvious future development is to extend the LAOs to
four-component relativistic response calculations at the DFT
level99 as well as at the two-component relativistic level.48
Another future development would be to combine the B-field
dependent LAOs described here with the B-field dependent
explicit 1-1 restricted kinetic magnetic balance schemes,
introduced by Liu and co-workers6062 and Komorovsk et
al.63
ACKNOWLEDGMENTS

M.I. thanks the EU research training network for his


postdoc stay in Odense, the French Ministry of Science and
Technology for his postdoc stay in Strasbourg, and the Slovak Research and Development Agency under Contract No.
APVV-20-018405 for the financial support. The authors also
thank the Danish Center for Scientific Computing for the
computational resources for this work. T.S. would like to
thank Kenneth Ruud Troms for stimulating discussions.
APPENDIX: DERIVATION OF NEW MATRIX
ELEMENTS FOR FOUR-COMPONENT LAOs

Here the derivations of matrix elements of the derivatives of the Dirac Hamiltonian Eq. 4 over LAOs are
given, as we have not been able to find them in literature. For
completeness we also give relevant derivatives of the twoelectron Coulomb term although they are the same as in
nonrelativistic work and can be found in Ref. 10, for example.
In the LAO basis the matrix elements of the one- and
two-electron operators are given by
h = M hN
=


M exp

i
B R MN r
2

1
hD + B rN c
2
+

1
r c
mK K r3
c2 K
K

N ,

A1

g = M PgNQ

= M P exp

i
B R MN r1
2

+ R PQ r2 g NQ ,

A2

where R MN = R M RN. From the Maclaurin expansion in per-

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-12

J. Chem. Phys. 131, 124119 2009

Ilia et al.

turbation strengths, one obtains the following nonzero matrix


elements beyond zeroth-order relevant for this work

1
i
hB = Q MNM rhDN + M rN c N ,
2
2

hmK =

1
M
c2

K
hBm
=

rK c
rK3

N ,

rK c
i
Q MN M r
2c2
rK3

A3
A4

N ,

i
gB = M PQ MNr1 + Q PQr2gNQ,
2

A5

A6

where Q MN is the antisymmetric matrix10


Q MN =

Z MN

Z MN

Y MN

X MN

Y MN

X MN .
0

A7

These matrix elements in AO basis have all been implemented in the HERMIT module in DIRAC04. In the linear response module, they are transformed to the corresponding
matrix elements in UMO basis appearing in Eqs. 31 and
34 using Eq. 13.
F. London, J. Phys. Radium 8, 397 1937.
Aa. E. Hansen and T. D. Bouman, J. Chem. Phys. 82, 5035 1985.
3
H. F. Hameka, Mol. Phys. 1, 203 1958.
4
J. A. Pople, Proc. R. Soc. London, Ser. A 239, 541 1957.
5
J. A. Pople, Proc. R. Soc. London, Ser. A 239, 550 1957.
6
H. F. Hameka, Rev. Mod. Phys. 34, 87 1962.
7
R. Ditchfield, D. P. Miller, and J. A. Pople, J. Chem. Phys. 53, 613
1970.
8
R. Ditchfield, Mol. Phys. 27, 789 1974.
9
K. Wolinski, J. F. Hinton, and P. Pulay, J. Am. Chem. Soc. 112, 8251
1990.
10
T. Helgaker and P. Jrgensen, J. Chem. Phys. 95, 2595 1991.
11
S. T. Epstein, J. Chem. Phys. 58, 1592 1973.
12
K. Ruud, R. Kobayashi, P. Jrgensen, K. L. Bak, and H. J. Aa. Jensen, J.
Chem. Phys. 100, 8178 1994.
13
J. Gauss, Chem. Phys. Lett. 191, 614 1992.
14
J. Gauss, J. Chem. Phys. 99, 3629 1993.
15
J. Gauss, Chem. Phys. Lett. 229, 198 1994.
16
J. Gauss and J. F. Stanton, J. Chem. Phys. 102, 251 1995.
17
J. Gauss and J. F. Stanton, J. Chem. Phys. 104, 2574 1996.
18
G. Schreckenbach and T. Ziegler, J. Phys. Chem. 99, 606 1995.
19
A. M. Lee, N. C. Handy, and S. M. Colwell, J. Chem. Phys. 103, 10095
1995.
20
G. Rauhut, S. Puyear, K. Wolinski, and P. Pulay, J. Phys. Chem. 6310,
6310 1996.
21
J. R. Cheeseman, G. W. Trucks, T. A. Keith, and Mi. J. Frisch, J. Chem.
Phys. 104, 5497 1996.
22
T. Helgaker, P. J. Wilson, R. D. Amos, and N. C. Handy, J. Chem. Phys.
113, 2983 2000.
23
P. Pyykk, Chem. Rev. Washington, D.C. 88, 563 1988.
24
P. Pyykk, Chem. Phys. 74, 1 1983.
25
N. C. Pyper, Chem. Phys. Lett. 96, 204 1983.
26
J. Vaara, Phys. Chem. Chem. Phys. 9, 5399 2007.
27
L. B. Casabianca and A. C. de Dios, J. Chem. Phys. 128, 052201 2008.
28
C. C. Ballard, M. Hada, H. Kaneko, and H. Nakatsuji, Chem. Phys. Lett.
254, 170 1998.
29
H. Fukui and T. Baba, J. Chem. Phys. 108, 3854 1998.
30
T. Baba and H. Fukui, J. Chem. Phys. 110, 131 1999.
31
R. Fukuda, M. Hada, and H. Nakatsuji, J. Chem. Phys. 118, 1015 2002.
1
2

R. Fukuda and H. Nakatsuji, J. Chem. Phys. 123, 044101 2005.


K. Kudo and H. Fukui, J. Chem. Phys. 123, 114102 2005; 124,
209901E 2006.
34
S. K. Wolff, T. Ziegler, E. van Lenthe, and E. J. Baerends, J. Chem. Phys.
110, 7689 1998.
35
M. Hada, R. Fukuda, and H. Nakatsuji, Chem. Phys. Lett. 233, 95
1995.
36
H. Fukui, T. Baba, and H. Inomata, J. Chem. Phys. 105, 3175 1996;
106, 2987E 1997.
37
V. G. Malkin, O. L. Malkina, and D. R. Salahub, Chem. Phys. Lett. 261,
335 1996.
38
S. K. Wolff and T. Ziegler, J. Chem. Phys. 109, 895 1998.
39
P. Manninen, P. Lantto, J. Vaara, and K. Ruud, J. Chem. Phys. 119, 2623
2003.
40
J. I. Melo, M. C. Ruiz De Aza, C. G. Giribet, G. A. Aucar, and R. H.
Romero, J. Chem. Phys. 118, 471 2003.
41
M. C. Ruiz De Aza, J. I. Melo, and C. G. Giribet, Mol. Phys. 101, 3103
2003.
42
J. I. Melo, M. C. Ruiz De Aza, C. G. Giribet, G. A. Aucar, and P. F.
Provasi, J. Chem. Phys. 121, 6798 2004.
43
P. Manninen, K. Ruud, P. Lantto, and J. Vaara, J. Chem. Phys. 122,
114107 2005.
44
M. Filatov and D. Cremer, J. Chem. Phys. 119, 701 2003.
45
K. Kudo, H. Maeda, T. Kawakubo, Y. Ootani, M. Funaki, and H. Fukui,
J. Chem. Phys. 124, 224106 2006.
46
Y. Ootani, H. Yamaguti, H. Maeda, and H. Fukui, J. Chem. Phys. 125,
164106 2006.
47
H. Maeda, Y. Ootani, and H. Fukui, J. Chem. Phys. 126, 174102 2007;
128, 129903E 2005.
48
M. Ilia and T. Saue, J. Chem. Phys. 126, 064102 2007.
49
R. Fukuda, M. Hada, and H. Nakatsuji, J. Chem. Phys. 118, 1027 2003.
50
S. Hamaya, H. Maeda, M. Funaki, and H. Fukuia, J. Chem. Phys. 129,
224103 2008.
51
J. Autschbach, J. Chem. Phys. 128, 164112 2008.
52
Y. Ishikawa, T. Nakajima, M. Hada, and H. Nakatsuji, Chem. Phys. Lett.
283, 119 1997.
53
M. Hada, Y. Ishikawa, J. Nakatani, and H. Nakatsuji, Chem. Phys. Lett.
310, 342 1999.
54
M. Hada, R. Fukuda, and H. Nakatsuji, Chem. Phys. Lett. 321, 452
2000.
55
M. Kato, M. Hada, R. Fukuda, and H. Nakatsuji, Chem. Phys. Lett. 408,
150 2005.
56
H. M. Quiney, H. Skaane, and I. P. Grant, Chem. Phys. Lett. 290, 473
1998.
57
H. M. Quiney, H. Skaane, and I. P. Grant, Adv. Quantum Chem. 32, 1
1999.
58
I. P. Grant and H. M. Quiney, Int. J. Quantum Chem. 80, 283 2000.
59
L. Visscher, T. Enevoldsen, T. Saue, H. J. Aa. Jensen, and J. Oddershede,
J. Comput. Chem. 20, 1262 1999.
60
Y. Xiao, D. Peng, and W. Liu, J. Chem. Phys. 126, 081101 2007.
61
Y. Xiao, W. Liu, L. Cheng, and D. Peng, J. Chem. Phys. 126, 214101
2007.
62
L. Cheng, Y. Xiao, and W. Liu, J. Chem. Phys. 130, 144102 2009; 131,
019902 2009.
63
S. Komorovsk, M. Repisk, O. L. Malkina, V. G. Malkin, I. Malkin
Ondk, and M. Kaupp, J. Chem. Phys. 128, 104101 2008.
64
T. Saue and H. J. Aa. Jensen, J. Chem. Phys. 118, 522 2003.
65
M. Hanni, P. Lantto, M. Ilia, H. J. Aa. Jensen, and J. Vaara, J. Chem.
Phys. 127, 164313 2007.
66
M. Gell-Mann, Nuovo Cimento, Suppl. 4, 159 1956.
67
T. Saue, in Relativistic Electronic Structure Theory. Part 1. Fundamentals, edited by P. Schwerdtfeger Elsevier, Amsterdam, 2002, p. 332.
68
T. Saue, Adv. Quantum Chem. 48, 383 2005.
69
G. A. Aucar, T. Saue, L. Visscher, and H. J. Aa. Jensen, J. Chem. Phys.
110, 6208 1999.
70
S. T. Epstein, The Variational Method in Quantum Chemistry Academic,
New York, 1974.
71
W. Kutzelnigg, J. Mol. Struct.: THEOCHEM 202, 11 1989.
72
C. van Wllen, in Calculation of NMR and EPR Parameters, edited by
M. Kaupp, M. Bhl, and V. G. Malkin Wiley-VCH, Weinheim, 2004, p.
85.
73
E. Dalgaard, Chem. Phys. Lett. 47, 279 1977.
74
T. Helgaker, in Geometrical Derivatives of Energy Surfaces and Molecular Properties, edited by P. Jrgensen and J. Simons Reidel, Dordrecht,
32
33

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

124119-13

Gauge origin independent calculations

1986, p. 1.
H. Sellers, Int. J. Quantum Chem. 30, 433 1986.
76
T. Helgaker and P. Jrgensen, in Methods in Computational Molecular
Physics, edited by S. Wilson and G. H. F. Diercksen Plenum, New York,
1992.
77
T. Saue, K. Fgri, T. Helgaker, and O. Gropen, Mol. Phys. 91, 937
1997.
78
L. Visscher and T. Saue, J. Chem. Phys. 113, 3996 2000.
79
M. Pecul, T. Saue, K. Ruud, and A. Rizzo, J. Chem. Phys. 121, 3051
2004.
80
W. Kutzelnigg and W. Liu, J. Chem. Phys. 131, 044129 2009.
81
A. F. Maldonado and G. A. Aucar, Phys. Chem. Chem. Phys. 11, 5615
2009.
82
W. Kutzelnigg, Phys. Rev. A 67, 032109 2003.
83
L. Visscher, Adv. Quantum Chem. 48, 369 2005.
84
K. G. Dyall and K. Fgri, Jr., Introduction to Relativistic Quantum
Chemistry Oxford University Press, New York, 2007.
85
J. Olsen, K. L. Bak, K. Ruud, T. Helgaker, and P. Jrgensen, Theor.
Chim. Acta 90, 421 1995.
86
X. Gonze, Phys. Rev. A 52, 1096 1995; 52, 4591E 1996.
87
M. Krykunov and J. Autschbach, J. Chem. Phys. 123, 114103 2005.
88
HERMIT, a molecular integral program, written by T. Helgaker, P. R. Taylor, and K. Ruud.
89
H. J. Aa. Jensen, T. Saue, and L. Visscher with contributions from V.
Bakken, E. Eliav, T. Enevoldsen, T. Fleig, O. Fossgaard, T. Helgaker, J.
Lrdahl, C. V. Larsen, P. Norman, J. Olsen, M. Pernpointner, J. K. Pedersen, K. Ruud, P. Salek, J. N. P. van Stralen, J. Thyssen, O. Visser, and
75

J. Chem. Phys. 131, 124119 2009


T. Winther, DIRAC, a relativistic ab initio electronic structure program,
release DIRAC04.0 2004 see http://dirac.chem.sdu.dk.
90
T. Saue and H. J. Aa. Jensen, J. Chem. Phys. 111, 6211 1999.
91
DALTON, a molecular electronic structure program, release 2.0 2005 see
http://www.kjemi.uio.no/software/dalton/dalton.html.
92
K. P. Huber and G. Herzberg, Molecular Spectra and Molecular Structure Van Nostrand Reinhold, New York, 1979.
93
L. Visscher and K. G. Dyall, At. Data Nucl. Data Tables 67, 207 1997;
see also http://www.chem.vu.nl/~visscher/FiniteNuclei/FiniteNuclei.htm.
94
L. Visscher and E. van Lenthe, Chem. Phys. Lett. 306, 357 1999.
95
K. Ruud, T. Helgaker, J. Olsen, P. Jrgensen, and K. L. Bak, Chem. Phys.
Lett. 235, 47 1995.
96
K. G. Dyall, J. Chem. Phys. 100, 2118 1994.
97
A. Maldonado and G. A. Aucar, J. Chem. Phys. 127, 154115 2007.
98
M. M. Sternheim, Phys. Rev. 128, 676 1962.
99
T. Saue and T. Helgaker, J. Comput. Chem. 23, 814 2002.
100
L. Visscher and K. G. Dyall, J. Chem. Phys. 104, 9040 1996.
101
T. H. Dunning, Jr., J. Chem. Phys. 90, 1007 1989.
102
L. Visscher, J. Styszinski, and W. C. Nieuwpoort, J. Chem. Phys. 105,
1987 1996.
103
J. N. P. van Stralen, L. Visscher, and J. F. Ogilvie, Phys. Chem. Chem.
Phys. 6, 3779 2004.
104
W. G. Schneider, H. J. Bernstein, and J. A. Pople, J. Chem. Phys. 28, 601
1958.
105
W. T. Raynes, in Specialist Periodical Report, Nuclear Magnetic Resonance, Vol. 7, edited by R. J. Abraham The Chemical Society, London,
1987, p. 24.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:
130.120.228.135 On: Fri, 20 Jun 2014 15:26:18

S-ar putea să vă placă și