FF Procedure 2

11
Finding Optimal Finite Field Strengths Allowing for a Maximum of Precision in the Calculation of Polarizabilities and Hyperpolarizabilities Ahmed A. K. Mohammed, [a] Peter A. Limacher,* [a] and Benoı ˆt Champagne [b] The finite field method, widely used for the calculation of static dipole polarizabilities or the first and second hyperpolarizabilities of molecules and polymers, is thoroughly explored. The application of different field strengths and the impact on the precision of the calculations were investigated. Borders could be defined and characterized, establishing a range of feasible field strengths that guarantee reliable numerical results. The quality of different types of meshes to screen the feasible region is assessed. Extrapolation schemes are presented that reduce the truncation error and allow to increase the precision of finite field calculations by one to three orders of magnitude. V C 2013 Wiley Periodicals, Inc. DOI: 10.1002/jcc.23285 Introduction Nonlinear optical (NLO) properties of molecules and polymers are a subject of high importance in organic chemistry and materials sciences. [1] The theoretical prediction of NLO proper- ties based on quantum chemical calculations is challenging and the development of new, along with improvement of existing methodology remains a substantial demand for com- putational chemistry. [2,3] The difficulties that occur in such cal- culations are related to the fact that the NLO properties render themselves very fragile quantities being high-order derivatives of the energy or of lower-order response properties. If the de- pendence of the energy E on an external homogeneous static electric field F is written as a Taylor expansion of the form EðFÞ¼ Eð0Þþ @E @F 0 F þ 1 2! @ 2 E @F 2 0 F 2 þ 1 3! @ 3 E @F 3 0 F 3 þ 1 4! @ 4 E @F 4 0 F 4 þ (1) ¼ Eð0Þ lF 1 2 aF 2 1 6 bF 3 1 24 cF 4 þ (2) the nonlinear responses, namely the first and second hyperpo- larizabilities b and c, are the third and fourth derivatives of E with respect to F . The other quantities are the permanent dipole moment l and the dipole polarizability a. The finite field (FF) method is a straightforward, easy-to- implement technique for the calculation of hyperpolarizabil- ities. In contrast to other methods such as the sum over states approach, coupled-perturbed HF, or response theory, the prop- erty of interest can be calculated simply from the knowledge of the energy at certain field strengths, and no additional information is needed about excited states or analytical deriva- tives with respect to the field components. These benefits render the FF method an universally applicable technique to any level of theory, with no additional requirements to a quan- tum chemical program other than to allow for an extra electric field term in the Hamiltonian. [4] It is thus common practice to first implement the FF method for the evaluation of hyperpo- larizabilities in newly designed programs or homebuilt quan- tum chemistry codes. [5–12] On the other hand, the FF method has also certain down- sides. It is, for example, limited to the evaluation of static mo- lecular properties, as time-dependent fields cannot be handled in a straightforward manner. Most crucial, however, is the de- pendence of the outcome of an FF calculation on the initially chosen field strength. It is well-known that too small fields introduce numerical noise on the calculated properties due to the finite convergence thresholds for the energy and wave function optimization. These parameters are thus generally tightened by orders of magnitude for hyperpolarizability evalu- ations compared to an ordinary single point energy calcula- tion. The problem with too small fields persists, however, due to the finite precision arithmetics of every computer, and can only be mitigated at most. At the other end of the scale, there is also an upper bound for feasible field strengths. A field chosen too strong leads to an inaccurate evaluation of molecular properties driven by two different factors: First of all, the higher-order terms in the Tay- lor expansion of eq. (2) lead to non-negligible contributions to the overall energy with increasing field strength. This effect is systematic and can be cancelled out by linear combination of [a] A. A. K. Mohammed, P. A. Limacher Department of Chemistry, McMaster University, Hamilton, Ontario L8S 4M1, Canada [b] B. Champagne Laboratoire de Chimie Th eorique, University of Namur, Rue de Bruxelles, 61, B-5000 Namur, Belgium E-mail: [email protected] Contract/grant sponsor: NSERC; Contract/grant sponsor: Swiss National Science Foundation fellowship; contract/grant number: PBEZP2-140249; Contract/grant sponsor: Academy Louvain; Contract/grant sponsor: Belgian Government; contract/grant number: IUAP No P7/05. V C 2013 Wiley Periodicals, Inc. Journal of Computational Chemistry 2013, 34, 1497–1507 1497 FULL PAPER WWW.C-CHEM.ORG

description

Chemistry

Transcript of FF Procedure 2

Page 1: FF Procedure 2

Finding Optimal Finite Field Strengths Allowing for aMaximum of Precision in the Calculation of Polarizabilitiesand Hyperpolarizabilities

Ahmed A. K. Mohammed,[a] Peter A. Limacher,*[a] and Benoıt Champagne[b]

The finite field method, widely used for the calculation of

static dipole polarizabilities or the first and second

hyperpolarizabilities of molecules and polymers, is thoroughly

explored. The application of different field strengths and the

impact on the precision of the calculations were investigated.

Borders could be defined and characterized, establishing a

range of feasible field strengths that guarantee reliable

numerical results. The quality of different types of meshes to

screen the feasible region is assessed. Extrapolation schemes

are presented that reduce the truncation error and allow to

increase the precision of finite field calculations by one to

three orders of magnitude. VC 2013 Wiley Periodicals, Inc.

DOI: 10.1002/jcc.23285

Introduction

Nonlinear optical (NLO) properties of molecules and polymers

are a subject of high importance in organic chemistry and

materials sciences.[1] The theoretical prediction of NLO proper-

ties based on quantum chemical calculations is challenging

and the development of new, along with improvement of

existing methodology remains a substantial demand for com-

putational chemistry. [2,3] The difficulties that occur in such cal-

culations are related to the fact that the NLO properties render

themselves very fragile quantities being high-order derivatives

of the energy or of lower-order response properties. If the de-

pendence of the energy E on an external homogeneous static

electric field F is written as a Taylor expansion of the form

EðFÞ ¼ Eð0Þ þ @E

@F

����0

F þ 1

2!

@2E

@F2

����0

F2 þ 1

3!

@3E

@F3

����0

F3 þ 1

4!

@4E

@F4

����0

F4 þ…

(1)

¼ Eð0Þ � lF � 1

2aF2 � 1

6bF3 � 1

24cF4 þ… (2)

the nonlinear responses, namely the first and second hyperpo-

larizabilities b and c, are the third and fourth derivatives of E

with respect to F. The other quantities are the permanent

dipole moment l and the dipole polarizability a.The finite field (FF) method is a straightforward, easy-to-

implement technique for the calculation of hyperpolarizabil-

ities. In contrast to other methods such as the sum over states

approach, coupled-perturbed HF, or response theory, the prop-

erty of interest can be calculated simply from the knowledge

of the energy at certain field strengths, and no additional

information is needed about excited states or analytical deriva-

tives with respect to the field components. These benefits

render the FF method an universally applicable technique to

any level of theory, with no additional requirements to a quan-

tum chemical program other than to allow for an extra electric

field term in the Hamiltonian.[4] It is thus common practice to

first implement the FF method for the evaluation of hyperpo-

larizabilities in newly designed programs or homebuilt quan-

tum chemistry codes.[5–12]

On the other hand, the FF method has also certain down-

sides. It is, for example, limited to the evaluation of static mo-

lecular properties, as time-dependent fields cannot be handled

in a straightforward manner. Most crucial, however, is the de-

pendence of the outcome of an FF calculation on the initially

chosen field strength. It is well-known that too small fields

introduce numerical noise on the calculated properties due to

the finite convergence thresholds for the energy and wave

function optimization. These parameters are thus generally

tightened by orders of magnitude for hyperpolarizability evalu-

ations compared to an ordinary single point energy calcula-

tion. The problem with too small fields persists, however, due

to the finite precision arithmetics of every computer, and can

only be mitigated at most.

At the other end of the scale, there is also an upper bound

for feasible field strengths. A field chosen too strong leads to

an inaccurate evaluation of molecular properties driven by two

different factors: First of all, the higher-order terms in the Tay-

lor expansion of eq. (2) lead to non-negligible contributions to

the overall energy with increasing field strength. This effect is

systematic and can be cancelled out by linear combination of

[a] A. A. K. Mohammed, P. A. Limacher

Department of Chemistry, McMaster University, Hamilton, Ontario L8S 4M1,

Canada

[b] B. Champagne

Laboratoire de Chimie Th�eorique, University of Namur, Rue de Bruxelles, 61,

B-5000 Namur, Belgium

E-mail: [email protected]

Contract/grant sponsor: NSERC; Contract/grant sponsor: Swiss National

Science Foundation fellowship; contract/grant number: PBEZP2-140249;

Contract/grant sponsor: Academy Louvain; Contract/grant sponsor:

Belgian Government; contract/grant number: IUAP No P7/05.

VC 2013 Wiley Periodicals, Inc.

Journal of Computational Chemistry 2013, 34, 1497–1507 1497

FULL PAPERWWW.C-CHEM.ORG

Page 2: FF Procedure 2

properties obtained at different field strengths, which is usu-

ally applied in literature.[13–18]

A second, more cumbersome effect is the change in the

electron configuration of the system at a certain field

strength.[19] The perturbative potential associated with an

external homogeneous electric field modifies the Hamiltonian

of a system in such a way that its spectrum becomes unbound

and, as a consequence, the true ground state would

correspond to a completely ionized molecular core with all

electrons located at minus infinity. In practice however, the

use of finite, atom-centered basis sets prevents such behavior.

If the field strength in such a calculation is chosen decently

enough, the ground state, corresponding to a metastable state

in the infinite basis set description, will closely resemble the

ground state of a zero field calculation. For too strong fields

however, an excited state at zero field could become favorable

in energy and changes place with the former ground state.

Hence, all properties evaluated at such field strengths reflect

the behavior of this excited state. Moreover, quantum chemical

methods lacking multireference character often fail to con-

verge in the vicinity of the intersection between these states.

In conclusion, choosing a reasonable field strength for an FF

calculation requires to cope with a rather delicate balance

between the two extremes of too weak fields, leading to nu-

merical errors that can be several orders of magnitude larger

than the actual property of interest, and too strong fields

causing the calculation of properties for an excited state.

This publication intends to analyze the FF method in details

for a broad range of field strengths, to characterize the behav-

ior of the noise at the low field limit, and to give prescriptions

about how to find a critical upper limit for the field. Once the

region of feasible field strengths is identified, the issue will be

addressed of how to reach a maximum of precision using

different FF meshes and extrapolation techniques. The final

section contains the application of the different extrapolation

schemes to some exemplary molecules.

Methods

FF equations for the electric field derivatives

The FF formulae needed to compute the different tensor ele-ments of the successive electric properties are straightforwardto derive and treated in-depth in many excellent text booksand articles.[6,20] Nevertheless, they are quickly recapitulatedhere as certain aspects should be highlighted to provide use-ful insights exploited in later sections.

The Taylor expansion of eq. (2) can be split into an evenand an odd part introducing symmetric and antisymmetric lin-ear combinations of the energy at equal positive and negativefields:

ESðFÞ ¼EðFÞ þ Eð�FÞ

2¼ Eð0Þ � 1

2aF2 � 1

24cF4 þ O F6

� �(3)

EAðFÞ ¼EðFÞ � Eð�FÞ

2¼ �lF � 1

6bF3 � 1

120dF5 þ O F7

� �(4)

This allows a completely separate treatment of the polariz-abilities and even order hyperpolarizabilities decoupled from

any effects of the dipole moment and all odd order hyperpo-larizabilities, which characterize noncentrosymmetric mole-cules. The dipole moments and polarizabilities can beobtained directly, rearranging these equations to

lðFÞ ¼ � EAðFÞF

¼ lþ 1

6bF2 þ 1

120dF4 þ O F6

� �(5)

aðFÞ ¼ 2Eð0Þ � ESðFÞ

F2¼ aþ 1

12cF2 þ 1

360eF4 þ O F6

� �(6)

which are already good estimates for l and a under the

assumption that F is chosen small enough.

For the evaluation of b and c, the energy has to be knownat two different field strengths, chosen here to be F and 2F,besides the energy at zero field. This allows to eliminate l anda from eqs. (5) and (6) to yield

bðFÞ ¼ 2lð2FÞ � lðFÞ

F2¼ 2EAðFÞ � EAð2FÞ

F3¼ bþ 1

4dF2 þ O F4

� �(7)

cðFÞ ¼ 4að2FÞ � aðFÞ

F2¼ �2ESð2FÞ þ 8ESðFÞ � 6Eð0Þ

F4

¼ cþ 1

6eF2 þ O F4

� �(8)

The last equality in eqs. (5)–(8) shows that all electric prop-erties can be formulated as Taylor expansions in the samefashion as the energy in eq. (3). Any quantity may thus be cal-culated as well from analytical dipole moments lð0Þ andlð6FÞ using eq. (5), or analytical polarizabilities að0Þ andað6FÞ using eq. (6), etc. This leads to less and less finite differ-

ences to take and consequently to more precise results. There

are, however, other methods to reduce the numerical error

based on different extrapolation schemes.

Error reduction

A well-established method to lift the scaling of the error ineqs. (5)–(8) to higher orders is based on Richardson extrapola-tion.[21] Specifically in quantum chemistry, this approach isequally well known as Romberg differentiation due to its simi-larity with the numerical integration scheme of the samename.[13] Both approaches are in fact exactly the same andwere successfully applied in the past.[15,16,22]

Two given Taylor expansions of an arbitrary quantity q eval-uated at different field strengths F and xF

qðFÞ ¼ qð0Þ þ aF þ bF2 þ cF3 þ dF4 þ… (9)

qðxFÞ ¼ qð0Þ þ axF þ bx2F2 þ cx3F3 þ… (10)

can be combined to obtain

xmqðFÞ � qðxFÞ ¼ ðxm � 1Þqð0Þ þ ðxm � xÞaFþ ðxm � x2ÞbF2 þ… (11)

in order to remove a certain power of F from the expansion

by a proper choice of m. If q is evaluated for a number of field

strengths following a geometric progression qn;0 ¼def

qðxnFÞ withcommon ratio x, then all the fqðxnFÞg form a set of zero-order

approximations for q(0). Refined values for q(0) are obtained

FULL PAPER WWW.C-CHEM.ORG

1498 Journal of Computational Chemistry 2013, 34, 1497–1507 WWW.CHEMISTRYVIEWS.COM

Page 3: FF Procedure 2

by acting on this set with the recursive relation

qn;m ¼def xmqn;m�1 � qnþ1;m�1

xm � 1: (12)

For every refinement step m, the scaling power of the erroris increased by one. Thus, if numerical values of qn;0 are known

for a geometric progression of, for example, five field

strengths, the final q0;4 is an approximation for q(0) with the

error scaling only as O F5ð Þ.Richardson extrapolation is widely used in literature to

refine the precision of FF properties, especially for the higher-order derivatives b and c, which significantly improve duringthe first few iterations.[23] Further refinement, however, is notpossible, as the quality of the result drops, once the range offield strengths spanned by the geometric progression extendsbeyond the feasible region.

A remedy for this is to shrink the ratio x of the progressionsuch that more data points lie within the feasible region.Although it is common practice to work with x = 2, values ofx < 2 were found to improve the precision.[13] To this end, for-

mulae (7) and (8) have to be generalized to

bðFÞ ¼ 6

x2 � 1� lðxFÞ � lðFÞ

F2¼ 6

F3x� xEAðFÞ � EAðxFÞ

x2 � 1(13)

cðFÞ ¼ 12

x2 � 1� aðxFÞ � aðFÞ

F2¼ 24

F4x2x2ESðFÞ � ESðxFÞ

x2 � 1� Eð0Þ

� �:

(14)

which is achieved using eq. (11), this time to remove the

lower-order terms from the energy expansion. Setting x ¼ 2,

the conventional expressions (7) and (8) are retrieved.

For the subsequent Richardson extrapolation, the occur-rence of solely even powers of F in the Taylor expansions (5)–(8) makes it more convenient to change the definition of m ineq. (12) to

qn;m ¼def x2mqn;m�1 � qnþ1;m�1

x2m � 1: (15)

with q ¼ fl; a;b; cg.

Polynomial fitting

Another way to extrapolate a and c from an FF calculation to-ward the values at zero field is to fit an appropriate functionof order r to a given set fFi; Eig of field strengths and ener-

gies.[20] For centrosymmetric systems, a and c are obtained by

fitting a polynomial of the form

EðFÞ � Efit �1

2afitF

2 � 1

24cfitF

4 þ… ¼Xr

k¼0

qk;fitF2k

ð2kÞ! (16)

to these data using the least-squares method. Note that only

even powers of F are needed, if the symmetrized energy of eq.

(3) is used. Thus, choosing r ¼ 1 allows the determination ofEfit and afit, an order of r ¼ 2 yields Efit, afit, and cfit, etc., also if

the inspected molecule is not centrosymmetric. Naturally, the

precision of the extrapolated values depends on the choice of

r for the polynomial as well as on the number of data in the

set. As a general rule, r has to be larger, the broader the con-

sidered range fFig is. The optimal choice of r, however,

depends on the nature and precision of the data, and is not

trivial to find.

Displacement indices

In this study, the fields Fi for the various data points fFi; Eigare generated in three different ways. They are either forming

arithmetic or geometric progressions using the construction

schemes

Fi;lin ¼ Fm � i

m(17)

Fi;exp ¼ F1 �FmF1

� � i�1m�1

(18)

with F1 and Fm as the smallest and largest fields under consid-

eration, or they are determined randomly. The random setsfFi;rndg can be brought in relation to the above progressions

by the definition of a displacement index d. To this end, the

random set is sorted in ascending order of Fi;rnd and then

compared with the corresponding element Fi;lin or Fi;exp. The

displacement index to an arithmetic progression dlin is con-

structed by defining Fm;lin ¼ Fm;rnd and adding up a penalty for

each element of the form ðFi;rnd � Fi;linÞ2. Using eq. (17), one

obtains

dlin ¼Xm�1

i¼1

Fi;rndFm

� i

m

� �2

: (19)

In the same spirit dexp is constructed, this time working withlnðFiÞ instead of Fi. The highest and lowest fields of the geo-

metric progression are defined to match Fm;exp ¼ Fm;rnd andF1;exp ¼ F1;rnd such that finally

dexp ¼Xm�1

i¼2

lnFi;rndF1

� i � 1

m� 1ln

FmF1

� �� �2(20)

is obtained using eq. (18).

The displacement index is of great help, when the similarityof a random set to an arithmetic or geometric progression hasto be assessed. In both definitions (19) and (20), d is a dimen-sionless quantity and d ¼ 0 means a perfect agreement of therandom set fFi;rndg with the corresponding progression,

whereas increasing values of d describe larger and larger devi-

ations from it.

Electronic structure calculations

All the calculations in this article were performed with theDALTON quantum chemistry program.[24] The reference val-ues to which the FF properties are compared were obtainedusing response theory (RT).[25,26] Molecular geometries wereoptimized at the same level of theory as the subsequentevaluation of the hyperpolarizabilities unless otherwisenoted.

Assessing the quality of finite differences requires the com-parison of many almost identical numbers. For a maximal

FULL PAPERWWW.C-CHEM.ORG

Journal of Computational Chemistry 2013, 34, 1497–1507 1499

Page 4: FF Procedure 2

precision of these small numerical differences, care was takenthat all wave functions, molecular energies, and property pa-rameters were tightly converged. All the energies presented inthis study are exact up to at least 2 � 10�12 au ¼ 2e�12 au. As

the smallest molecules considered here have absolute energies

above 40 au, we end up with a relative precision of at least

1e�13 for the energy or, alternatively stated, 13 significant

digits.

The convergence criteria for the response equations werechosen such that at least 10 significant digits for aRT can be

guaranteed. For bRT and cRT, it is nine and eight significant dig-

its, respectively. The errors on the FF properties investigated in

the present study are orders of magnitude higher, so that all

the response properties can essentially be considered as the

correct reference values.

Results and Discussion

Field dependence of the FF quantities

Figure 1 shows cFF of the neon atom as a function of the field

strength, as calculated using eq. (8). Independent from the

quantum chemical method used, all the curves of Figure 1a

have the same characteristics: After an initial noisy region

below 0.001 au, values for cFF begin to stabilize before excited

states become similar in energy and start to alter the slope at

�0.05 au. Finally, all the curves drop down to zero after the

field reaches ionizing strength, which cannot be accounted for

by atom-centered basis sets.

In Figure 1b, the window for stable calculations of c can be

seen to vary if different basis sets are applied. It is well-known

that basis sets lacking diffuse functions only recover a small

fraction of the actual c (for instance, the cc-pVDZ value is

about four orders of magnitude too small when compared to

t-aug-cc-pVQZ) which goes along with the need of stronger

fields before reaching the stable region. For the same reasons,

when going toward large values of F, the end of the stable

region occurs at smaller field strengths for more diffuse basis

sets.

More insight on the field dependence of cFF can be gained

if an accurate reference value for c is known. In our case,

response theory was applied to obtain a cRT from an unrelaxed

CCSD calculation (Fig. 1c) and a HF calculation (Fig. 1d). The

unsigned relative error jecj ¼ jcFF=cRT � 1j reveals more details

about the accuracy of cFF: There is a critical field strength, cor-

responding to ionization or state inversion, below which the

error is well-behaved and, while moving toward smaller F, cFFapproaches more and more the cRT value. This is the region of

field strengths for which extrapolation schemes are meaning-

ful. Then, for even smaller field strengths, there is a clear point

where the noise starts to drop in, systematically worsening the

precision on cFF. The aforementioned shift of optimal field

strengths for different basis sets is also visible here. Neverthe-

less, the relative precision is the same for all basis sets and

reaches a maximum of about three significant digits. This is

generally sufficient for investigations focusing on predicting

and interpreting c values.

Figure 1. Finite field second hyperpolarizabilities cFF for the neon atom as a function of the field strength F using eq. (8). (a) compares different quantum

chemical methods using the aug-cc-pVTZ basis set. (b) shows values for cFF from relaxed CCSD(T) calculations using different basis sets. (c) and (d) show

the unsigned relative error of the FF method in comparison to response theory for the same basis sets as in (b). Data in (c) are unrelaxed CCSD calcula-

tions, those in (d) HF calculations.

FULL PAPER WWW.C-CHEM.ORG

1500 Journal of Computational Chemistry 2013, 34, 1497–1507 WWW.CHEMISTRYVIEWS.COM

Page 5: FF Procedure 2

If the study is extended to polar molecules such as water,

similar observations can also be made for the first hyperpolar-

izability b. In fact, as is shown in Figure 2a, using eqs. (5)–(8),

all the properties that are derivatives of the energy with

respect to the electric field follow the same pattern. The

threshold field strength at which noise starts to corrupt pre-

cise calculations is naturally larger for higher-order derivatives

and the overall precision thus drops to less significant digits.

At the same time, the critical field strength defining an upper

bound for meaningful FF quantities remains unchanged or is

shifted to smaller values as energies acquired at twice the field

strength will enter the calculations, according to eqs. (7) and

(8). This leads to an overall shrinking of the interval of accepta-

ble field strengths when going to higher-order derivatives.

Figure 2b points out that, within the same order of deriva-

tives, any tensor element can be evaluated to the same preci-

sion over the same interval of acceptable field strengths. This

allows us to focus on one distinct tensor element only (cxxxx in

the following). The findings for this quantity can then be gen-

eralized to any other tensor element. Nevertheless, an optimal

field strength in x, y, and z direction has to be found individu-

ally for very prolate or oblate molecules (e.g., planar, p conju-

gated systems) if mixed tensor elements such as cxxyy are

calculated.

It should further be noted from Figure 2 that all the slopes

in the region of acceptable field strengths are nearly parallel

for all quantities and correspond to a quadratic deviation (i.e.,

a slope of two in the double-logarithmic representation) from

the reference value. This means that the error due to the

usage of finite fields in this region is almost exclusively related

to the contribution of the next term in the odd/even power

expansion, which is the quantity’s second-order derivative

(e.g., b in the case of l, or c in the case of a ) and is thus

straightforward to eliminate. In the low field regions of Figure

2, the noise takes overhand. Its characteristic scaling correlates

with the order of the derivative. The constant round-off error

on the energy differences is made field-dependent through

the division by a specific power of the field strength, accord-

ing to eqs. (5)–(8), which causes the error-increase at low

fields. The choice of plotting the unsigned error in Figure 2 has

the benefit of illustrating this noise rather nicely, but on the

other hand also leads to artificial peaks on the high field side,

whenever cFF � cRTð Þ changes sign (visible e.g., around 0.2 au

for l, or around 0.05 au for b).

The low field limit

As the energy is only correct up to a certain threshold dE, thefinite differences in eqs. (5)–(8) cannot be more precise

than that. As all the FF quantities (q) are divided by a certain

power of F, the following relations holds for their absolute

precision (dq):

jdEj � jFdlj � jF2daj � jF3dbj � jF4dcj: (21)

That the noise scales indeed as inverse powers of F can be

seen on the left hand side of Figure 2a. Relation (21) defines

an ultimate limit of how precise an FF quantity can be known.

If, for example, for a molecule a and c are evaluated atF ¼ 0:001 au and its energy is computed to a precision ofdE ¼1e�10 au, then the absolute error introduced on a is of

the order 0.0001 au and likewise c cannot be more precise

than 6100 au, independent of its absolute magnitude.

The high field limit

Taking only the numerical noise into consideration, it would

be profitable to choose the field strength as high as possible.

However, this is prohibited by non-negligible higher-order

terms in the Taylor expansion (2) and an eventual change in

the electronic configuration. Figure 3 illustrates the interplay

between the ground and excited state energies for neon at

rather strong fields and its effect on a and c. It holds in gen-

eral that the more diffuse a basis set, the smaller the required

field strength to cause an abrupt change of the slope of E(F),

which corresponds to an inversion of the electronic ground

state with the first excited state.[19] This relation is well-estab-

lished and is confirmed by Figure 3a for basis sets up to quad-

ruple augmentation. At zero field, the energy is essentially the

Figure 2. Relative error of finite field properties in comparison to response theory for different tensor elements of water at the HF/aug-cc-pVTZ level of

theory. (a) shows the dipole moment l, polarizability a, and first and second hyperpolarizabilities b and c along the direction of the dipole. (b) shows

all nonzero tensor elements for c. The water molecule is oriented such that the dipole moment points along the x axis, and z corresponds to the out-of-

plane axis.

FULL PAPERWWW.C-CHEM.ORG

Journal of Computational Chemistry 2013, 34, 1497–1507 1501

Page 6: FF Procedure 2

same (within 0.0001 au) for all but the nondiffuse cc-pVDZ ba-

sis. Applying weak fields below 0.05 au, the ground state

energy changes are mostly described by the � 12 aF

2 term and

are, therefore, systematic and similar for all basis sets. Beyond

that point, the energy of the q-aug-cc-pVDZ calculation sud-

denly starts to drop rapidly. In the same way, although some-

times less visible, all the other basis sets have a characteristic

field strength at which the energy starts to behave differently.

This sudden drop originates from some excited states that

stabilize faster than the ground state when F increases so that,

beyond a certain field amplitude, they are successively replac-

ing the ground state. This is depicted in Figure 3b for the q-

aug-cc-pVDZ basis where the first state inversion happens at

�0.05 au followed by another one at �0.1 au. The steeper

slope in the interval ½0:05 au; 0:1 au� is attributed to the larger

a value for that state.[27,28] The second inversion happens with

a highly excited state whose zero field energy certainly lies

beyond the ionization potential of neon at EIP � 0:8 au, such

that its corresponding a is physically meaningless.

It is, however, possible to go beyond this point with less

augmented basis sets, and Figure 3c shows proof that indeed

reasonable values for aFF and cFF can be obtained forF > 0:05 au using the aug-cc-pVDZ basis (The same is visible

in Figure 1b for other basis sets). Figure 3d shows what hap-

pens to the polarizability calculated above the critical field

strength. At each crossing point of the ground state,

F > 0:05 au jumps to a new value approximately correspond-

ing to the polarizability of that new state (visible as a large

error in Fig. 3d). As cFF is calculated according to eq. (8), it

contains also the energy at twice the field strength and the

critical point thus appears at half the value of the first ground

state crossing. At even stronger fields, cFF changes sign and is

thus not depicted in the figure anymore.

The comparison of Figures 3c and 3d reveals that it can be

difficult to find the critical field strength only based on inspec-

tion of the ground state energy. In Figure 3a, the graph for

the aug-cc-pVDZ basis set possesses no obvious kink from an

intersection with excited states. Nevertheless, cFF in Figure 3c

is only acceptable for F < 0:1 au. For stronger fields, the error

is of the same magnitude or even larger than c itself.

For small molecules, it is usually affordable to add another

layer of diffuse basis functions in order to clearly locate the

intersection of two states and thus easily define the upper

bound for feasible field strengths. The problem looks different

for cases where the use of diffuse basis sets is discouraged or

even prohibitive due to near-linear dependences within the

overlap matrix.[15,29] A famous, well studied example are poly-

ynes (PY), linear chains of alternating carbon–carbon single

and triple bonds.[30] In these extended systems, even nondif-

fuse basis sets such as cc-pVTZ can have convergence issues

because of the relatively small triple bond length. The basis

sets used to investigate NLO properties are thus usually of

Figure 3. Ground and excited state energies of neon obtained by a CASSCF(8,13) calculation as a function of the external electric field. (a) The ground

state energy using cc-pVDZ with varying amounts of diffuse basis functions. (b) Ground and several excited state energies, applying the q-aug-cc-pVDZ ba-

sis set. (c) The relative energy difference (E(F)/E(0)�1) , as well as the relative error on the polarizability (aFF/aRT�1) and on the second hyperpolarizability

(cFF/cRT�1) obtained with aug-cc-pVDZ. (d) The same data as in (c) for q-aug-cc-pVDZ. [Color figure can be viewed in the online issue, which is available at

wileyonlinelibrary.com.]

FULL PAPER WWW.C-CHEM.ORG

1502 Journal of Computational Chemistry 2013, 34, 1497–1507 WWW.CHEMISTRYVIEWS.COM

Page 7: FF Procedure 2

double-, maximally triple-zeta quality with few or no diffuse

and polarization functions, also in the most recent

literature.[15,16,31,32] In Figure 4, FF calculations are performed

at the HF/6-31G* level for PY chains containing 1 (4a), 5 (4b),

and 10 (4c) triple bonds, respectively. For the small molecule

of Figure 4a, the usage of a nondiffuse basis set makes it diffi-

cult to locate any kink in the field dependence of the ground

state energy that could be assigned to an intersection with

excited states. However, additional information, such as the

HOMO-LUMO gap, is of great assistance in tracking such a

change. Indeed, the minimum value of the gap aroundF ¼ 0:2 au corresponds to soft kinks in the relative error for a,and also for c at half that field strength.

If the chain length is increased, a certain ‘‘diffuseness’’ of the

basis is generated by the fact that neighboring atomic basis

functions overlap substantially and thus account for some

missing features of a small basis set. This effect is also the rea-

son why accurate values for the NLO properties of elongated

systems can be obtained without the aid of diffuse basis func-

tions.[33] Figures 4b and 4c are a verification of that statement.

For C20H2, a kink at �0.018 au is easily visible for the ground

state energy. The discontinuities in the relative error on a and

c are more expressed than in the smaller chains, such that an

overall localization of the critical field strength is facilitated.

C10H2 presents an intermediate behavior with a change of

slope in E(F) around F ¼ 0:05 au.

In conclusion, (i) the upper limit for an electric field that will

lead to meaningful FF quantities is usually found in a straight-

forward manner, if the basis set contains enough diffuse

functions, (ii) extended molecular chains behave well without

diffuse basis functions, and (iii) in problematic cases, additional

quantities such as the excitation energies or the HOMO-LUMO

gap may help to locate the critical field strength.

Figure 4. Field dependence of several quantities for PY chains of 1, 5, and 10 triple bonds, obtained at the HF/6-31G* level of theory. (a) shows the results

for C2H2, (b) for C10H2, and (c) for C20H2. The upper row shows the absolute energy. In the middle, the energy difference between the highest occupied

and the lowest unoccupied canonical molecular orbital (HOMO-LUMO gap) is depicted. On the bottom, the relative energy difference, as well as the error

on a and c is shown for each molecule analog to Figures 3c and 3d. [Color figure can be viewed in the online issue, which is available at

wileyonlinelibrary.com.]

Table 1. Relative error eq5qfit/qfit21 with q5{a, c} for neon at the HF/t-

aug-cc-pVQZ level of theory.

Fixed range [0,0.08] Flexible range

r ea ec Range r ea ec

1 1.4e�2 – [0,0.01] 3 1.1e�7 �1.2e�3

2 �7.9e�4 1.4e�1 [0,0.02] 3 �8.0e�9 �6.5e�6

3 8.3e�5 �2.8e�2 [0,0.04] 4 �8.7e�8 1.3e�4

4 �1.4e�5 7.3e�3 [0,0.06] 5 5.9e�8 �8.7e�5

5 2.8e�6 �2.1e�3 [0,0.10] 8 �7.6e�6 9.9e�3

6 �7.7e�7 7.7e�4 [0,0.02] 3 7.3e�9 �5.2e�5

7 �2.6e�7 2.0e�4 [0.02,0.04] 4 �2.8e�7 2.2e�4

8 �3.3e�6 4.6e�3 [0.04,0.06] 4 1.2e�5 �1.9e�3

9 �3.2e�5 5.8e�2 [0.06,0.08] 4 �1.2e�3 1.7e�1

10 �1.6e�4 3.5e�1 [0.08,0.10] 2 �2.1e�2 7.9e�1

qfit is the coefficient from a least-squares fit of a polynomial of order r to

field-dependent energies generated for a linear mesh of fields with spac-

ing DF ¼ 0.001 au. The ranges [Fmin, Fmax] are given in atomic units.

FULL PAPERWWW.C-CHEM.ORG

Journal of Computational Chemistry 2013, 34, 1497–1507 1503

Page 8: FF Procedure 2

Figure 5. The relative errors |ea| (red) and |ec| (blue) for neon at the HF/t-aug-cc-pVQZ level of theory obtained by fitting a polynomial of optimal order to

five randomly selected field strengths F1 to F5 and the zero field F0 ¼ 0. A total of 40,000 random samples are ordered according to different criteria: (a)

With respect to the highest field strength F5 within a sample. (b) With respect to the deviation of each sample from an arithmetic progression using eq.

(19). (c) the same as (a) but considering only samples with dlin � 0.05. The black line is the logarithmic mean for |ea| and |ec| for a number of samples with

a similar abscissa, indicated by the green bars on the bottom. The dashed black line in (c) corresponds to the black line in (a) for comparison. [Color figure

can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

Figure 6. The same data as in Figure 5 with the samples ordered in view of similarities with a geometric progression. (a) The common ratio, defined asffiffiffiffiffiffiffiffiffiffiffiffiffiF5=F1

4p

according to eq. (18) using m ¼ 5. (b) The deviation from a geometric progression applying eq. (20). (c) the same as (a) considering only samples

with dexp�0.1. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

Page 9: FF Procedure 2

Analyzing different FF meshes

Once we know the upper and lower bounds for feasible field

strengths, the question arises about how the fields fFig should be

chosen within these boundaries in order to maximize the preci-

sion while using only a small number of data points fFi; Eig.For that purpose, a total of 101 FF calculations were per-

formed for neon at the HF/t-aug-cc-pVQZ level, with F being

linearly distributed between 0 and 0.1 au according to eq.

(17). From Figure 1d, the critical field strength for the energy

is determined to be around 0.08 au (it corresponds to two

times the critical field for c around 0.04 au). Table 1 illustrates

the outcome of a polynomial fit for different polynomial orders

as well as for different subsets within these 101 data points. It

is found that the entire range below the critical field is suitable

for an accurate fit. For the 81 data points within the feasible

region, a polynomial of seventh order turns out to be the opti-

mal choice leading to afit and cfit of approximately seven and

four significant digits, respectively. This is about an order of

magnitude better than the values obtainable from Figure 1d.

On the right hand side of Table 1, the results for different

subsets within the range ½0; 0:1 au� are investigated. Only

extrapolations with optimal choices for r are shown. It can be

found that all the fits including the range ½0:08 au; 0:1 au� areof inferior precision, confirming the location of the critical field

strength in that region. Apart from that, the precision of seven

and four significant digits for a and c is met by most of the

subsets if the range around zero is included for the fit.

To mitigate problems with accidentally precise fits (due to

the stochastic nature of the data set), a large number of ran-

dom subsets has been generated by selecting six points (con-

taining the zero-field energy and five arbitrary points) out of a

total of 80 field strengths, linearly distributed between 0.001

au and 0.080 au. In this way, an array of 40,000 relative errorsjeaj and jecj was created by fitting a polynomial (with optimal

choice for r) to each random set. In Figure 5, the change of

these errors with respect to certain properties of the samples

is analyzed. It is found that the relative error is reduced on av-

erage, the more the fields F1 to F5 resemble an arithmetic pro-

gression. Also the choice of F5, the strongest field within a

sample, has an impact on the error. The FF properties can be

improved systematically by about one order of magnitude if

F5 is reduced from the critical field strength down to �0.04

au. For lower fields, an oscillatory behavior sets in, that is par-

tially related to different order r of the optimal polynomial fit.

For example, most sets with F5 ¼ 0:02 au are best fitted with r

¼ 3, whereas for F5 ¼ 0:03 au, fits with r ¼ 3 and r ¼ 4 are

equally likely and thus less precise on average. In Figure 5c,

the restriction to samples with a nearly linear mesh indeed

comes along with a slight reduction of the average error.

Figure 6 shows the same relative errors sorted according to

similarities with a geometric progression. The common ratio

defined by the smallest and strongest fields F1 and F5 is

directly related to parameter x of eqs. (13) and (14) and leads

to minimal errors for values of x that are substantially smaller

than 2. Also, a strong deviation of the sample from a geomet-

ric progression, indicated by dexp, is found to increase the

relative errors rather drastically. However, it should be noted

that about 90% of all the samples have dexp < 3 and in this

region the error increases only slightly. Consequently, restrict-

ing the samples to closely resemble a geometric progression

does not lead to meaningful reductions of the error in Figure

6c, but avoiding a too large deviation from a geometric pro-

gression is generally a benefit.

Extrapolating FF quantities

We now turn to Richardson extrapolation, which has the

advantage over the polynomial fit that no additional parame-

ters, such as the order r, need to be chosen. The effects of this

Figure 7. The minimal relative error |ec| as a function of common ratio x

and refinement step m (all dimensionless quantities), averaged over a data

set of 120 molecules shown on the top. The FF calculations were done

with HF/6-31G* based on CAM-B3LYP/6-31G* geometries. For each mole-

cule, various cn;m were calculated using eq. (15). The minimal relative error

is found, by picking that n which leads to the smallest deviation from the

reference value cRT, while m and x stay fixed. The averaging over the mole-

cules is done by taking the logarithmic mean. [Color figure can be viewed

in the online issue, which is available at wileyonlinelibrary.com.]

FULL PAPERWWW.C-CHEM.ORG

Journal of Computational Chemistry 2013, 34, 1497–1507 1505

Page 10: FF Procedure 2

kind of extrapolation are studied for the second hyperpolariz-

ability of a large set of molecules shown in Figure 7. For every

molecule, a minimal relative error is found by varying the field

strength, while the common ratio x and the degree of refine-

ment m in eq. (15) are fixed. Although very noisy for individual

molecules, the average over multiple molecules produces

smooth graphs that show significant trends.

It is found that the first two refinement steps reduce the rel-

ative error by more than one order of magnitude. Going

beyond m ¼ 2 does not lead to further improvement anymore

as the lowering of the truncation error is superseded by the

accumulated round-off error in eq. (15). The dependence of

the error on x is increased for large m and leads to best results

for x ¼ ½1:2; 1:8�. Values for x < 2, in particular x ¼ffiffiffi2

p, were al-

ready suggested in Ref. [13] based on inspection of an analyti-

cal function and one numerical example. These findings are

greatly confirmed here based on a broad data set of 120

molecules.

Exemplary cases

We finally show the gain in precision that can be achieved

using the methods of polynomial fitting and Richardson

extrapolation for some exemplary molecules. Table 2 contains

the relative error for a, b, and c of tetrachloromethane and

acetonitrile, two reference molecules for NLO measure-

ments.[22] Table 3 shows similar data for increasing chains of

cumulene (CnH4). These conjugated systems have huge NLO

properties with the peculiarity that a sign change for c can

occur depending on the level of theory.[34] This leads to a

rather small cRT for C6H4, whereas all the other properties like

a and the higher-order derivatives of the Taylor expansion are

extremely large. Hence, the precision on cFF is very limited

with only one significant digit in the unrefined case.

As a general observation, the improvement due to extrapo-

lation is larger for a (up to three orders of magnitude increase

for the precision) than for b and c (one to two orders of mag-

nitude). Comparing the different extrapolation schemes, no

preference can be given. However, Richardson extrapolation

with m ¼ 1 needs only four energy calculations, at optimally

chosen fields f0; F; xF; x2Fg, that return a precision similar to a

polynomial fit with 14 data points. The only time when the

polynomial fit outperforms Richardson extrapolation is the

special case of C6H4 where apparently the precision benefits

from the larger array of data points.

It should finally be mentioned that the best extrapolation

estimates shown in Tables 2 and 3 are all obtained for field

strengths between 0.001 au and 0.004 au with a tendency

Table 2. The relative error eq for different electric-field response properties of tetrachloromethane (CCl4) and acetonitrile (CH3CN) computed at the HF/

aug-cc-pVTZ level of theory.

Reference Raw value[a] Polynomial fit[b] Richardson extrapolation[c]

Molecule[d] q cRT/au eq eq;fit x¼ 2 x¼ffiffiffi2

px¼

ffiffiffi24

p

CCl4 axx 65.67 8.0e�6 �5.5e�8 2.9e�8 �7.4e�9 1.7e�8

CCl4 cxxxx 6119 �6.4e�3 3.2e�4 �2.4e�4 �1.4e�4 2.1e�4

CH3CN axx 36.68 8.2e�6 �1.1e�9 �1.7e�8 �2.0e�10 �1.1e�9

CH3CN bxxx 12.41 8.6e�4 �3.3e�5 9.0e�5 8.0e�5 3.9e�5

CH3CN cxxxx 3572 �1.6e�3 �6.5e�6 �1.4e�4 �3.3e�5 �4.8e�5

Different extrapolation schemes are tested for a set containing the energies for 13 field strengths defining a geometric progression in the range

[0.001 au, 0.008 au] (and another 13 fields of opposite sign for the noncentrosymmetric CH3CN) plus the zero field energy.[a] Properties calculated by eqs. (6)–(8) for F ¼ 0.001 au.[b] A polynomial fit of order r¼ 3 (i.e., including terms up to F6 with the coefficients of odd powers set equal zero) was used for afit and cfit. For bfit, apolynomial of order r¼ 2 (including terms up to F4) was fitted to the field dependent dipole moments of eq. (5).[c] Smallest error achievable within above set of 14 data points using a refinement step m¼ 1 and different choices of x.[d] Only tensor elements along the x axis are reported. For CCl4, this coincides with one of the molecule’s C2 axes. In CH3CN, the x axis is defined as the

line connecting all heavy atoms.

Table 3. The relative error ec for methane, ethylene and cumulene chains computed at the HF/aug-cc-pVDZ level of theory.

Reference Raw value Polynomial fit Richardson extrapolation

Molecule cRT=au ec r¼ 3 r¼ 4 x¼ 2 x¼ffiffiffi2

px¼

ffiffiffi24

p

CH4 1350 1.0e�3 2.9e�6 �4.9e�5 �2.4e�5 3.0e�6 �1.6e�5

C2H4 1757 �6.7e�4 1.6e�5 1.7e�4 �6.7e�5 �2.1e�5 �3.2e�5

C3H4 5820 5.2e�4 �7.4e�5 �1.4e�4 �1.5e�5 �7.8e�5 �5.5e�5

C4H4 10633 1.7e�3 �2.2e�4 3.9e�5 �3.4e�5 �1.3e�4 �1.6e�5

C5H4 11523 4.3e�3 �9.1e�4 �1.8e�4 �2.6e�4 �5.6e�5 �2.5e�4

C6H4 1587 7.2e�2 �2.1e�2 1.7e�4 1.7e�3 2.2e�3 3.5e�2

C7H4 �30383 �7.9e�3 2.4e�3 3.1e�5 �3.3e�4 �2.7e�5 3.7e�4

The same field strengths and settings described in Table 2 and footnotes apply also here.

FULL PAPER WWW.C-CHEM.ORG

1506 Journal of Computational Chemistry 2013, 34, 1497–1507 WWW.CHEMISTRYVIEWS.COM

Page 11: FF Procedure 2

toward the lower value for the elongated chains of Table 3.

On the other hand, the critical field strength lies beyond 0.02

au for all cases. So it seems essential to stay at least a factor

of 10 below the critical field strength for performing the most

precise calculations of NLO properties.

Conclusions

The present study assesses the reliability of the FF method for

evaluating the polarizabilities and hyperpolarizabilities and

proposes ways to reduce the numerical error of such calcula-

tions. It is found that for every molecule, an individual region

of feasible field strengths can be highlighted, which is defined

by an upper and a lower bound. Fields chosen below that

region suffer from a too large round-off error when taking the

finite energy differences. This error scales proportionally to the

convergence threshold of the energy and some inverse power

of the field strength, dependent on the property. The upper

bound is imposed by the critical field strength originating

from intersections between the ground and excited state ener-

gies and depends on molecule-specific factors, but also

strongly on the augmentation level of the basis set.

Within the feasible region, the main source of error stems

from higher-order derivatives of the Taylor expansion that can

be removed by means of Richardson extrapolation. The first

two refinement steps reduce the error by one or two orders

of magnitude for c and up to three orders for a. Additionalsteps, however, do not lead to further improvement. A bench-

mark on 120 molecules confirms that a common ratio of x < 2

and a refinement step of m ¼ 2 yields the most precise

results.

A polynomial fit for different meshes of field strengths

within the feasible region reveals the same preference for val-

ues of x < 2. Although a polynomial of sufficiently high order

provides good estimates over the whole range of feasible field

strengths, a clear tendency to improved results is discernible,

when values close to the critical field strength are discarded

from the fit. Similar observations from Richardson extrapola-

tion lead to the conclusion that a factor of about a 10th of

the critical field strength should be the upper limit for an

extrapolation leading to a maximum of precision.

Acknowledgments

The authors acknowledge a generous allocation of computer time

granted by SHARCNET, a partner consortium in the Compute Can-

ada national HPC platform.

Keywords: finite field method � hyperpolarizability � Richardsonextrapolation � error reduction � polynomial fitting

How to cite this article: A. A. K. Mohammed, P. A. Limacher, B.

Champagne, J. Comput. Chem. 2013, 34, 1497–1507. DOI: 10.1002/

jcc.23285

[1] H. Nalwa, S. Miyata, Nonlinear Optics of Organic Molecules and Poly-

mers, CRC Press: Boca Raton, 1996.

[2] J. L. Br�edas, C. Adant, P. Tackx, A. Persoons, B. M. Pierce, Chem. Rev.

1994, 94, 243.

[3] Y. Luo, H. Agren, P. Jorgensen, K. V. Mikkelsen, Adv. Quantum Chem.

1995, 26, 165.

[4] H. D. Cohen, C. C. J. Roothaan, J. Chem. Phys. 1965, 43, S34.

[5] R. J. Bartlett, G. D. Purvis, Phys. Rev. A 1979, 20, 1313.

[6] H. A. Kurtz, J. J. P. Stewart, K. M. Dieter, J. Comput. Chem. 1990, 11, 82.

[7] M. Nakano, I. Shigemoto, S. Yamada, K. Yamaguchi, J. Chem. Phys.

1995, 103, 4175.

[8] G. Maroulis, J. Chem. Phys. 1998, 108, 5432.

[9] K. Kamada, M. Ueda, H. Nagao, K. Tawa, T. Sugino, Y. Shmizu, K. Ohta,

J. Phys. Chem. A 2000, 104, 4723.

[10] F. A. Bulat, A. Toro-Labb�e, B. Champagne, B. Kirtman, W. Yang, J. Chem.

Phys. 2005, 123, 014319.

[11] T. Touma, M. Kobayashi, H. Nakai, Theor. Chem. Acc. 2011, 130, 701.

[12] S. Wouters, P. A. Limacher, D. Van Neck, P. W. Ayers, J. Chem. Phys.

2012, 136, 134110.

[13] M. Medved, M. Stachov�a, D. Jacquemin, J.-M. Andr�e, E. A. Perpete, THE-

OCHEM 2007, 847, 39.

[14] B. Champagne, B. Kirtman, Int. J. Quantum Chem. 2009, 109, 3103.

[15] M. de Wergifosse, B. Champagne, J. Chem. Phys. 2011, 134, 074113.

[16] P. A. Limacher, Q. Li, H. P. Luthi, J. Chem. Phys. 2011, 135, 014111.

[17] L. M. Abreu, T. L. Fonseca, M. A. Castro, J. Chem. Phys. 2012, 136, 234311.

[18] J. B. Robinson, P. J. Knowles, J. Chem. Phys. 2012, 137, 054301.

[19] D. M. Bishop, S. A. Solunac, Chem. Phys. Lett. 1985, 122, 567.

[20] F. Sim, S. Chin, M. Dupuis, J. E. Rice, J. Phys. Chem. 1993, 97, 1158.

[21] L. F. Richardson, J. A. Gaunt, Phil. Trans. R. Soc. Lond. A 1927, 226, 299.

[22] F. Castet, E. Bogdan, A. Plaquet, L. Ducasse, B. Champagne, V. Rodri-

guez, J. Chem. Phys. 2012, 136, 024506.

[23] B. Champagne, B. Kirtman, In Handbook of Advanced Electronic and

Photonic Materials and Devices, Vol. 9; H. S. Nalwa, Ed.; Academic

Press: Burlington, 2001, pp. 63 – 126.

[24] DALTON, a molecular electronic structure program, Release 2.0 (2005),

Available at: http://www.kjemi.uio.no/software/dalton/dalton.html

[25] B. Jansik, P. Salek, D. Jonsson, O. Vahtras, H. Agren, J. Chem. Phys.

2005, 122, 054107.

[26] T. Helgaker, S. Coriani, P. Jørgensen, K. Kristensen, J. Olsen, K. Ruud,

Chem. Rev. 2012, 112, 543.

[27] M. Urban, A. J. Sadlej, Theor. Chim. Acta 1990, 78, 189.

[28] M. M�erawa, M. R�erat, J. Chem. Phys. 1998, 108, 7060.

[29] P. A. Limacher, K. V. Mikkelsen, H. P. Luthi, J. Chem. Phys. 2009, 130, 194114.

[30] S. Eisler, A. D. Slepkov, E. Elliott, T. Luu, R. McDonald, F. A. Hegmann, R.

R. Tykwinski, J. Am. Chem. Soc. 2005, 127, 2666.

[31] J.-W. Song, M. A. Watson, H. Sekino, K. Hirao, J. Chem. Phys. 2008, 129,

024117.

[32] J.-W. Song, M. A. Watson, H. Sekino, K. Hirao, Int. J. Quantum Chem.

2009, 109, 2012.

[33] G. J. B. Hurst, M. Dupuis, E. Clementi, J. Chem. Phys. 1988, 89, 385.

[34] J. O. Morley, J. Phys. Chem. 1995, 99, 10166.

Received: 8 November 2012Revised: 25 February 2013Accepted: 4 March 2013Published online on 4 April 2013

FULL PAPERWWW.C-CHEM.ORG

Journal of Computational Chemistry 2013, 34, 1497–1507 1507