Can Quantum Chemistry Provide Reliable ... - ACS Publications

RoWvo. W(2). = R0(W -. (28) v ( 3 ) =Ro(w- E^yv -. RoEWwW etc., showing that all order ..... 1\\ = (lHjj + 7»iu +7ii;i)/15 with summation convention,...
0 downloads 0 Views 3MB Size
Chapter 2

Can Quantum Chemistry Provide Reliable Molecular Hyperpolarizabilities? Rodney J . Bartlett and Hideo Sekino

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

Quantum Theory Project, Departments of Chemistry and Physics, University of Florida, Gainesville, FL 32611-8435

We analyze the evolution offirstprinciple molecular theory in obtaining reliable values of molecular hyperpolarizabilities. Such quantities place severe demands on the capability of quantum chemistry, as large basis sets, frequency dependence, and high levels of electron correlation are all shown to be essential in obtaining observed, gas phase hyperpolar­ izabilities to within an error of 10%. Ab initio Hartree Fock results are typically in error by nearly a factor of two, while the errors due to the neglect of frequency dependence average ~10-30% depending upon the particular process. It is also shown that for small molecules, stan­ dard semi-empirical approaches, like INDO and INDO/S, will often not even give the correct sign for hyperpolarizabilities. It is demonstrated that using state-of-the-art correlated, frequency dependent methods, it is possible to provide results to within 10%. Excluding any one of the essential elements, though, destroys the agreement. In the late 70's, one of us (RJB) became interested in molecular hyperpolarizabilities as the essential element in non-linear optics (NLO) when Gordon Wepfer at the Air Force Office of Scientific Research called attention to J. F. Ward's results from dcinduced second harmonic generation (dcSHG) experiments and the enormity of their discrepancy with theoretical results. Tables 1 and 2 are extracted from a slightly later, 1979, paper of Ward and Miller (1) demonstrating the problem. Not only was the existing theory results of the time tvpically in error by a factor of 3 to 5 in magnitude for the electric susceptibility, , but even the signs were frequently wrong. (The sign is positive if the measured quantity is positive, where μ is the permanent dipole moment.) When experimental numbers were obtained by other techniques, which should only differ by different dispersion effects that are typically less than 10%; they, too, had little correspondence with the dcSHG data. Note from Table 1 the -3500 value for NH3 from refractivity virial data compared to -209±5 for the dcSHG, 0097-6156/96/0628-0023$18.75/0 © 1996 American Chemical Society

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

24

NONLINEAR OPTICAL MATERIALS

Table 1 x\\ > units of 10~~ esu/molecule. n

Theoretical values f r o m various

33

molecular orbital calculations — semi-empirical ( S E ) , uncoupled H a r t r e e - F o c k ( H F ) , a n d coupled H a r t r e e - F o c k ( C H F ) — a n d a single other experimental value are included for comparison. T h e sign of fix^

is unambiguously determined by

the experiment a n d is independent of the sense chosen for the molecular z axis. T h i s table extracted f r o m J . F . W a r d a n d C . K . M i l l e r . Adapted from ref. 1. THEORY x[,

co

0.112

+

2)

SHG

+129 ±14

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

±0.005

SemiEmpirical

Uncoupled HF

-43.5 (+95)

+879 +420 +387 -438

b

d

Coupled HF

Other EXP

c

c

c

e

NO

0.15872 +147 ±17 ±0.00002

+

H S" +

2

0.974

+47.7

b

-43 ±9

±0.005 N-H

3

+

1.474

-209 ±5

56.4

-44.4 -65.l -19.0 -15.6 -40.8

b

±0.009

H

2

0-

+

1.86 ±0.02

-94 ±4

120

90.6

b

f

f

f

h

-52.5 -79.2 -21.9 -51.6 -48.0

h

f

f

f

h

h

Electric dipole moments in Debye units from Landolt-Bernstein, Zahlenwerte und Funktionen, Neue Serie, Vols. II/4 and II/6 (Springer-Verlag, Berlin) and Ref. 18. The sense of the NO moment is suggested by F. P. Billingsley II, J. Chem. Phys. 63 2267 (1975); 62, 864 (1975). a

N. S. Hush and M . L. Williams, Theoret. (1972). Signs for N H 3 and H 2 O are ambiguous. b

c

Chim.

Acta (Berlin) (25), 346

J. M . O'Hare and R. P. Hurst, J. Chem. Phys. (46), 2356 (1967).

X zzz (0;0,0) from A. D. McLean and M . Yoshimine, J. Chem. Phys. 3682 (1967). d

e

{

(46),

S. P. Liebmann and J. W. Moskowitz, J. Chem. Phys. 54, 3622 (1971).

P. Lazzeretti and R. Zanasi, Chem. Phys. Lett. (39), 323 (1976) Note added in proof. Recent reconsideration of the sign of these entries yields the negative signs now shown here — P. Lazzeretti (private communication). Overall consistency is substantially improved by this change. f

Refractivity virial data from A . R. Blythe, J. D. Lambert, P. J. Petter and H. Spoel, Proc. R. Soc. (London) A 255, 427 (1960). 8

h

G. P. Arrighini, M . Maestro and R. Moccia, Symp. Farad. Soc. 2, 48 (1968).

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

-35008

2.

Reliable Molecular

BARTLETT & SEKINO

25

Hyperpolarizabilities

Table 2 x[| i n units of 10" esu/molecule. Values f r o m other dc-electric field-induced second-harmonic generation ( d c S H G ) , three-wave m i x i n g ( T W M ) , K e r r effect, and t h i r d - h a r m o n i c generation ( T H G ) experiments, along w i t h theoretical results, are included for comparison. T h i s table extracted f r o m J . F . W a r d a n d C . K . M i l l e r (7). Adapted from ref. 1. 39

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

y dcSHG Ward& Miller

dcSHG"

TWM

b

( 3 )

Kerr

THG

C

H

2

65.2 ±0.8

79

-

47 ±5

80 ±12

N

2

86.6 ±1.0

-

104

120 ±10

107 ±17

95.3 ±1.6

110

100

-

-

111.9 ±1.3

-

192

o

2

co

2

CO

144 ±4

NO

235 ±7

H S

865 ±22

NH

511 ±9

2

3

H 0 2

194 ±10

138 322

-

d

f

750

f

±160

Theory 34

e

-

156 ±23

-

-

-

-

Data from G . Mayer, C. R. Acad. Sci. B276, 54 (1968) and G . Hauchecorne, F . Kerberve* and G . Mayer, J . Phys. (Paris)32, 47 (1971), normalized using the d c S H G coefficient for argon from R. S. Finn and J . F. Ward, Phys. Rev. Lett.26, 285 (1971). a

Data from W . G . Rado, Phys. L e t t . l l , 123 (1967), normalized using the d c S H G coefficient for argon from R. S. Finn and J. F. Ward, Phys. Rev. Lett. 26, 285 (1971). b

c

J . F. Ward and G . H . C . New, Phys. Rev.185, 579 (1969).

d

A . D . Buckingham and B . J . Orr, Proc. R. Soc. (London) A 3 0 5 , 259 (1968).

Xzzzz (0;0,0,0) from A . D . M c L e a n and M . Yoshimine, J . Chem. Phys. 46, 3682 (1967). e

A . D . Buckingham, M . P. Bogard, D . A . Dunmur, C . P. Hobbs and B . J . Orr, Trans. Fora. Soc.66, 1548 (1970). f

8

Xzzzz (0;0,0,0) from R. J . Bartlett and G . D . Purvis (private communication).

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

26

NONLINEAR OPTICAL MATERIALS

In Table 2, the discrepancy among different experiments is even more apparent for x||

compared to other d c S H G values, three wave mixing ( T W M ) , Kerr effect,

and third harmonic generation ( T H G ) experiments. The theory is even off by about a factor of 2 from a static value for H2. The one reasonable theoretical number they cite is our static, correlated value for N , one o f our first (unpublished), that is 2

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

(3)

beginning to be i n reasonable agreement with x\\ • This chapter, which is intended to be useful to experimentalists who are trying to assess the reliability of the theory, but also to theoreticians, as it provides an overview of the theory which can and has been used, addresses the demands that hyperpolarizabilities place on first-principle ab initio electronic structure theory; and the level required for the adequate evaluation of molecular hyperpolarizabilities. After doing some basic elementary perturbation theory that underlies all that is done, and which provides definitions and a framework for discussion, we w i l l explain the various levels of quantum chemical application, illustrated by numerical results we have obtained that emphasize the various approximations employed in electronic structure in its application to hyperpolarizabilities. We w i l l also demonstrate how the theory has necessarily evolved to provide a realistic treatment of such properties. The order of presentation w i l l follow our own odyssey that led us to introduce electron correlation into molecular hyperpolarizabilities (2,3), using, then new, manybody perturbation theory ( M B P T ) methods (4,5); to make a prediction for the F H molecule, that had later ramifications (6,7); to use new coupled-cluster (CC) correlated methods including those augmented by triples (8,9); to explore vibrational polarizabilities i n static electric fields (70); and to introduce frequency dependence by developing analytical derivative time-dependent Hartree Fock ( T D H F ) theory (11,12). The latter, which enables a first (decoupled) treatment of both the essential frequency dependent and correlation aspects of the problem (13), culminates i n a uniform study of ten molecules (14). The correlation calculations were assisted by parallel developments i n analytical derivative C C / M B P T theory (15-18). Finally, we coupled correlation and frequency dependence i n the new equation-of-motion ( E O M ) C C method for hyperpolarizabilities (19,20). B y taking this evolutionary viewpoint, we hope the current review w i l l complement the other recent excellent reviews (21-23) on the topic. The theory also offers different, conceptual viewpoints on hyperpolarizability evaluation and interpretation, and this degree of flexibility should be better appreciated when comparing theoretical numbers and addressing N L O design criteria. We w i l l conclude with some recommendations for some future developments in the continuing evaluation of predictive quantum chemical methods for N L O material design.

Perturbation Theory of Molecular Hyperpolarizabilities The quantities of interest are the electric susceptibilities x[| ^ and x[| ^. In the gas phase experiments of interest here, there is no particular distinction between macroscopic and microscopic susceptibilities and the hyperpolarizabilities, as they are simply 2

3

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

Reliable Molecular

BARTLETT & SEKINO

related.

Hyperpolarizabilities

27

Those of particular interest are obtained from electric field (dc) induced

second harmonic generation experiments (7). Namely, a sample o f gas is subjected to a dc field ( £ ) and an optical electric field £cj(e

at frequency u, to

+ e~ )

lu;t

0

ltJt

induce a dipole moment at frequency 2a/. A l l o w i n g for the various manipulations required to relate the molecular quantities to the laboratory fixed observables,

X\ ( - 2 a / ; o, u/, UJ) = xf\~^\ o, a/, a/) + g ^ X | |

(1)

2

X| ^(-2a/;a/,a/) 2

for second

harmonic

generation

(SHG)

X | ^ ( - 2 a / ; o , a / , a ; ) , which is d c S H G , by studying fx

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

3

2 )

(-

2 w

; > w

is separated

from

as a function of T. Numbers

2w

relative to a standard, typically H e gas, are obtained. Revisions o f the H e reference value cause some slight rescaling o f the experimental numbers (22). Furthermore, we relate the susceptibilities to the molecular hyperpolarizabilities via, x [ | ( - 2 a ; ; a ; , a ; ) =/3||(-2a;; a/,a;)/2 2)

(3)

Xy \-2u\o,u,u)

=

(

7\\{-2UJ\O,U,OJ)I§

2

)

= \ [(7«ii + Ttiit + 7u'ij)/15J where the Einstein summation convention is employed, meaning all repeated indices are summed, i.e.

= d x+tiyy+^zzi

The parallel designation (||) means measured

X

parallel to the dc field, while the (_L) component can be similarly obtained. A s and 7y are properties of molecules, they constitute the objective for

first-principle

quantum mechanical evaluation; the subject of this chapter. The basic idea underlying any treatment o f a molecular response to an electric field, static or oscillatory, is the solution of the molecular Schrodinger equation in the presence of the field. We w i l l first consider the static case, generalizing the approach for frequency dependence in Section 7. Our perturbation e -W = e • Q

i^i is given

e

0

i by the interaction of an electron e; at position f; with a static electric field, £o — £xoi> ~f" £yoj "4" ^zo^i where e

XOi

£ , and e yo

zo

(?)

are the field strengths in the JC, y, and z directions of magnitude

In atomic units, e becomes minus unity, so we have a Hamiltonian, Wo)

= H + £ -W = Ho-?o'Y, > 0

0

fi

(

4

)

i where H consists of the usual kinetic energy, - ^ S ^ i i 0

" S ^

+f S ^ +^ S

a n c

* potential energy,

composed of the electron nuclear attraction, the two-

electron, electron-electron repulsion operator, and the proton-proton repulsion. Greek

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

28

NONLINEAR OPTICAL MATERIALS

letters indicate atoms and i,j electrons. W contains the negative sign. The solution to the time independent Schrodinger equation

provides the ground, tj) =

and excited states, {V**}. and associated eigenvalues,

0

for the unperturbed molecule.

Notice these are the exact, many-particle

unperturbed states. We now seek a solution for the perturbed Hamiltonian i n its ground state H{e )Wo)

= E{e )*{e )

0

0

(6)

0

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

The natural way to attempt a solution is perturbation theory. H(e ), 0

^{EO) a n d E(e )

Q

assume the field lies solely i n the z direction, so E (H

+ e W)

0

Hence, we expand,

i n a perturbation series i n E . F o r the time being, we w i l l

0

U

0

+ £ tf

0

0

+ £oV>

( 1 )

X O

= E

+ 4V>

(2)

= 0 and E

Y O

z

o

= e . Then, 0

+ • • )

( 3 )

(7) = (E + e E^

+ ^(2) +

0

+

£

A s the equality has to be true for any power of e

together, we obtain for e j ;

0

- H )TI>M

(E

0

0

(E„

for EI

- H )^

{E„ - Ho)^

= (W-

(8)

EW)TI>

0

(9)

EP>)lP> - Efft+o

= (W-

2)

0

= ( W -

^ d ) + . . .)

(i.e. a l l quantities are linearly

0

independent), gathering terms of a given power o f e

for Ei

e

- E^i>o

E^y™

- ^

V

(

1

- £

(

2

(

2

(10)

)

and for ei (E

0

- «„)V>

= {W-

( 4 )

E^y®

- EPHo

- E ^

(

1

)

¥

2

)

(11)

These are sufficient to take us through the electronic dipole moment, //, the dipole polarizability tensor, a , and the j3_ and 7, hyperpolarizability tensors. Multiplying on the left by (ip \, and using the fact that (tp \(E - H ) = 0, we have = mWtyo) 0

0

0

0

(12)

Similarly, we obtain = (1>o\W\ll>W)

£(2) = {^ \{W 0

- E^\^)

= (1> \W\1>V>) = (^ \W l)

0

E&

= ^ \W\^) 0

= (W\E - H \iP ) 0

0

{2)

-

( > 13

EWtyMtyW)

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

Reliable Molecular

2. BARTLETT & SEKINO

29

Hyperpolarizabilities

where we have used the intermediate normalization condition, (ip \^ ^) = 0 for m =j£ 0. The second form comes from manipulations that demonstrate the 2n+l rule that an rfi order wavefunction w i l l determine E and the analogous 2n rule for even orders. m

0

2 n + 1

1

Instead of straightforward perturbation theory, we can also derive these formulae from explicit differentiation of the expectation value of ^{s ) with respect to e . F r o m equation 6, we have the expectation value, 0

0

(14)

=

E(e )(*(e )\*( )) 0

0

£o

0

0

0

and differentiation gives, (H(e ) 0

-

E{e )) 0

d£ f

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

(H{e )

-

0

\de

0

E(e j) 0

a

(15)

dE de

0

A s we want to evaluate this at e

0

= 0, using (7i — E )\tjj ) = 0, and its complex 0

0

0

conjugate (c.c), gives

£L-WKI*)L- ) = 3 ! £ >

(22)

( 3

(2)

2

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

and the second hyperpolarizability;

dej

-

de

w

del/-*'Y°

0

^ ( 3 ) \

=

4

!

£ ( 4 )

(

2

3

)

That is, 1

1

E(e ) = E - fi e - -yOL z which gives the well known series. 0

0

z

zo

zz

Z0

1

- —ftzzz£

zo

4

(24)

~ -^lzzzz£

+•

zo

M o r e generally, when all components are considered, with the Einstein summa­ tion convention, we have ~ •• •

E = E — \L{Ei — —yOLij£io£j — yPijkEioEjoEko ~ -^7ijkl io jo ko lo 0

£

0

£

£

£

or for the induced dipole moment, V>i{£o) = M°)

+ ij£jo

+

a

Pijk£jo£ko + \ilijkl£jo£ko£lo 3!

+ ...

(26)

Whether the numerical factors are included determines the choice of conventions for polarizabilities. Ward and co-workers use the perturbative definition, meaning that £(2) _ ^ £(3) _ ^ j £ ( 4 ) _ ^ ^ g^ yjhWt power series choice directly a n (

e

m

e

associates the derivatives with the polarizabilities, obviating the numerical factors. Since this choice has direct correspondence with the energy derivatives, it appeals more to theoreticians. To evaluate the polarizabilities, we require a knowledge of the perturbed wavefunctions (wavefunction derivatives). It is convenient to introduce the resolvent op­ erator, R

0

= (E

0

where Q represents the projector (Q =Q) 2

- H )- Q 0

(27)

l

of all functions orthogonal to the unper­

turbed reference, V'o- One such set consists of all other eigenfunctions of H

Q

crete and continuous), making Q =

^ k?0

/(V J^ ,

1

'

• F°

r

t m s

(dis­

particular set,

N

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

(25)

Reliable Molecular

2. BARTLETT & SEKINO

R \r ) 0

k

= (E

-

0

4 ) V*)0)

Q(Eo-

31

Hyperpolarizabilities

0)

B y virtue of excluding

V'o, the inverse Ro operator is well denned. In terms of the resolvent, V>

(1)

= Ro(wW

(2)

v

( 3 )

=

E^)

VO

= R (W

RoWvo (28)

-

0

= R o ( w - E^yv

-

R WW oE

etc., showing that all order wavefunctions are recursively

w

computed from the prior

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

ones. K n o w i n g these solutions, we can compute the energy from (29)

= (1> \W\il> ) M

0

or explicitly, i n a few cases, E&> =

Wo\WR W\i> ) 0

0

£(3) = (i> \WR (w 0

E^RoWtyo)

0

= (1> \WRo(w - EM)RO(W

- EM)ROW\1> )

0

0

-

E^(^ \WR W\i; ) 0

2

0

0

(30) etc. The particular choice of expansion of ij)^ in the set of eigenfunctions of H leads 0

to the well known sum-over-state (SOS) formulas for the z components, -fi

(31a)

= (il> \z\il>o)

z

0

k^o

&o - &

k

(*-tf>)(*-tf>)

~7zzzz —

/

J

fe,/,m#o

(EO - Ef?^ {EO - E\ ^ [EO 0

>™zz

Em^

< 5 { E

0

- E ^

(3 Id) The other tensor elements are obtained from exchanging x and y with z i n all distinct ways. Henceforth, we w i l l recognize that a l l practical molecular quantum chemical calculations employ a finite, discrete basis set, so from the beginning we are limited to this choice, and we need not consider any continuum.

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

32

NONLINEAR OPTICAL MATERIALS

Evaluation of Static Hyperpolarizabilities F r o m the above, it should be abundantly clear that in a basis the derivative approach, equations 20 to 23, and the S O S expressions, equation 31, are simply two equivalent ways of expressing polarizabilities. Furthermore, considering that the choice of excited eigenfunctions of H

0

is just one choice for a complete set representation

of the perturbed wavefunctions, {ip^},

the above S O S forms probably attract more

significance in N L O than warranted. For example, since E -

is the excitation

0

energy, and since ( V ^ I V ^ ) is the z-component of the transition moment, it should be possible to evaluate a

from equation 31b purely from experiment by knowing

u

all electronic excitation energies and transition moments (including those for the

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

continuum in the exact case). The problem is knowing them a l l — a very large number! Instead, attempts to estimate the S O S by the few known excitation energies and transition moments is likely to be very far from the true value (see our contribution later i n this volume (24)). Note for the ground state, all contributions to

from

have the same sign, so there is no potential cancellation among the neglected

ip^

contributions. The problem is further compounded for (3^ and 7 ^ , where in addition to transition moments from the ground to excited states, it is necessary to know the transition moments relating two excited states, and that information is hard to obtain. Note that

can have either sign. j

7777

also can have either sign, since although

the lead term corresponds to (V>o |#o - ^o\i^)

which must be negative (giving

a positive contribution to 7 ^ ) , the second term is positive, attenuating the value of 7

? m

. Lacking a proof that the magnitude of one must be greater than the other,

either sign is possible. F r o m the above we have two viewpoints on the evaluation of static polarizabil­ ities. We can either evaluate energy derivatives of the Schrodinger equation i n the presence of the perturbation, or attempt some approximation to the S O S . Obviously, the former does not require any truncation. ( A s we w i l l see i n Section 8 later, with proper handling neither does the finite basis SOS.) The simplest recipe for evalua­ tion of the derivatives is to use what is called the finite-field technique. That simply means solve the Schrodinger equation in the presence of the perturbation by choosing a small finite value for e of 0.001 a.u., e.g. A d d i n g an electric field quantity to H 0

Q

gives an unbounded H(e )

operator, and i f we obtain its exact solution, the lowest

0

energy state would be the field ionized state, a molecular cation plus an electron; but i n practice we must use a finite basis set for its solution which is effectively like putting the molecule into a box, and this gives a valid E(e ). 0

Repeating the

procedure at e = - 0 . 0 0 1 , we could obtain the dipole from the numerical derivative, 0

E(E ) 0

E{-e )

-

0

2e

(32)

0

The accuracy depends upon the size of the finite field strength. If it is too large, the numerical derivative is not very accurate, while i f too small, there is not a numerically significant change i n E(e ). 0

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

Reliable Molecular

BARTLETT & SEKINO

To obtain all /z, a , expressions like A~4 Janet

= 5 [E(2e ) i0

= 4[E(e ) i0

33

f3_ and 7, we need several more points, so we obtain

- E(-2e )]

+ [E(e )

io

+ £(-e,- )] 0

i0

[E(2e ) io

- E(-e j\ io

- E(-2e j\ io

+

0(e ) 5

0

- 6£(o) +

0(e«)

See (2) for others. Note each expression is accurate to the next odd (even) order since only a , 7, e and /£, /3, 6 are interrelated (2). This exclusion of the next higher-order contribution greatly helps the precision of such a calculation. However, note that the energy needs to be accurate to a couple of significant digits better than ef i f we are to get /?„,. That is, i f £ =0.001, we require energies to be 10" . If Sio =0.01, we would require at least 10' . For 7 „ we would need 10" to 10" . A s molecular integrals in quantum chemical calculations are seldom much more accurate than 1 0 ' , not to mention other parts of the calculation, finite field procedures for hyperpolarizabilities can raise serious precision problems. D

11

t 0

8

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

Hyperpolarizabilities

M I

10

14

12

Another problem lies in the proliferation of tensor elements i n P and 7. Many energy calculations involving field strengths i n different directions are required to evaluate all the numerical derivatives, and at higher levels of sophistication these are quite expensive calculations. The solution to the above problem is to analytically evaluate the derivatives. The , , simplest is —\i = (Vo|^|V o)» which is just an expectation value. For the others, analytical means that while solving the Schrodinger equation for E, we also directly obtain all components of the derivatives, ||s $ ^ , > etc. in about an z

equivalent amount of time. This means we differentiate before evaluation by using equations 20 to 23 and the explicit solutions for the wavefunctions and derivatives, which are proportional to the perturbed wavefunctions given i n equation 28. In a substantial formal and computational achievement of 30 years duration, primarily fueled by the necessity of analytical gradients for atomic displacements i n molecules (15,25,26), such analytical procedures have been developed i n quantum chemistry. Their limitation is that they have not been implemented for all methods. For example, any-order analytical higher derivatives with respect to electric field perturbations have been developed for the Hartree-Fock treatment of hyperpolarizabilities by exploiting the recursive nature of perturbation theory (11,27), equations 28, 29. For correlated methods, analytical second-order perturbative theory, [MBPT(2)], derivatives are available for a and p_ (15). For other methods, even including highly sophisticated correlated methods like coupled-cluster theory (28), the induced dipole /i(e ) can be evaluated analytically, from which numerical derivatives provide a , /3, and 7 (13). In this way, at least one e or two to three orders of magnitude is gained i n the precision of and 7. 0

0

Note that equation 31 represents analytical expressions, too, since no finite field is involved i n their evaluation. The latter viewpoint leads to the analytical evaluation of the (dynamic) polarizability using the equation-of-motion ( E O M ) C C method (20). Obviously, it would be ideal to be able to analytically evaluate third and

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

34

NONLINEAR OPTICAL MATERIALS

fourth derivatives using such powerful C C correlated methods, but the theory and implementation has not yet been developed.

Basis Sets and Hartree-Fock Theory N o w that we know a way to calculate a static hyperpolarizability, we can consider other aspects of the calculation. The first approximation to consider is Hartree-Fock (HF), self-consistent field (SCF) theory.

That is, we evaluate the energy and its

derivatives for the perturbed Hamiltonian by obtaining the energetically best (lowest) single determinant solution, $

=

0

.A( . In 0

N

the absence of e W, that means £ H F = ($o\Ho\$o),

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

0

and further, H ~

H = ] T f(i) k=i

0

Q

N

where / = h + v ^

and W ( l ) = £

e

particle operator v ^(l) are the solutions

/

is chosen to be H

0

e

/

/

(

0

=

a n

d the correlation perturbation

replaces W. Then E&> = ($ |V# V|$ ) = 0

0

0

(9 \V\^). 0

This defines M B P T ( 2 ) . M B P T ( 2 ) is the simplest correlated method, consisting of the initial contribution due to double excitations from occupied to unoccupied orbitals

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

Reliable Molecular

BARTLETT & SEKINO

(i.e.

Hyperpolarizabilities

39

M B P T is usually applied in a given finite order n [MBPT(w)], where

n-4

is the highest frequently used. In the absence of an electric field, proper treatment of correlation i n hyperpolar­ izabilities requires a double perturbation approach (43) where all couplings between V

and e W

are allowed (44), with e W

0

applied i n a given order to describe the

0

particular polarizability, and V is preferably included in all orders. A straightforward double perturbation approach is possible, but usually in practice, the requisite cou­ pling between V and W is handled in two other ways. The first way is by correlating states, and then adding the W perturbation; the route taken in equations 28 to 31: This w i l l be the E O M - C C route described in Section 8. Second, we can take the viewpoint that we w i l l first solve the H F problem, and then add correlation to that H F solution straightforwardly, except that H ,

V, $

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

Q

and tp^

0

are all dependent on e . Q

That is, we evaluate correlation corrections to the electric field perturbed Hamiltonian as f

= fo + £W + v J* as in the C H F method, so e

£

= E /*« = E (

Ho(e ) 0

/

o

+

£

o

W

+

w

(36)

V = H(e )-H (e ) 0

E(e ) 0

=

£

C

H

EcM^o)

F

0

0

E (£ )

+

(2)

=

+ EW(e )

0

0

+ •• •

{*o{£o)\H(e )\*o(eo)) 0

and all evaluation of correlation corrections pertain to V(e )\ such as E^(e ) 0

0

0

0

=

0

and, similarly, in higher orders of M B P T (2,3,7,40). Just

( (£ )\V(£ )\j;W(£ )), 0

as i n H F theory, we have the option of doing this analytically or as a finite field. CC

theory offers a natural generalization of M B P T that sums categories

excitations to infinite order. excitations (i.e.

of

For example, C C S D means all single and double

# C C S D = exp (T\ 4- T2)$

where $

0

0

is the independent particle

model and T\ and T2 are the single and double excitation operators, T * 2

= E

0

( 3 7 )

a

xjj * = 7||/6, so 3

esu/molecule/a.u.

Basic sets listed i n Tables 7 and 8.

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

Reliable Molecular

BARTLETT & SEKINO

M B P T , by letting ^ c c be e dependent, [(V'CC = H(e ) = H + V{e ) + e W and Ecc{e ) = t 0

0

0

Q

0

0

41

Hyperpolarizabilities

e x

0

P (T(£ ))$o(£o)), along with (^o)\n(e )e ^ (e )]. 0

T

0

0

0

Even better, because of the advantage of analytical evaluation even for the induced dipole, we might describe how that is done in C C theory. M B P T follows as a special case (18).

Because ipcc( o)

= exp (T(£ ))$ (£ )

£

0

evaluate the induced dipole p(e ), 0

0

is an infinite series, we do not

0

as

(39)

fco) = (l>Cc(£o)\W\4>cc(£o))/{i>cc(eo)\l>cc(e )) 0

because the expression would have to be truncated. Instead, we can derive the form (18-28), jt(e )

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

0

= .

(40)

pq

where A is a de-excitation operator, complementary to the excitation operator T and j

pq

is the element of the one-particle "relaxed density" matrix (28).

B o t h A and T

have to be determined from the C C equations. Subsequent finite-field differentiation relative to e

provides the various polarizabilities like *ff§°^ = a , d

0

etc.

flgffig^

=

2!/3>

A s mentioned, this greatly improves precision and diminishes the number of

distinct e

values required.

0

Table 4

shows the effect of correlation on static xf\®\0> 0) and x[| ^(0; °> °> ° ) 3

Note the factor of 2 change for x\\^ of N H , and the order of magnitude and sign 3

change for H 2 S . Similarly,

changes substantially. This large correlation effect has

been observed since the initial correlated studies of molecular hyperpolarizabilities (2,3), and makes it apparent that "predictive" ab initio methods for hyperpolarizabil­ ities must include electron correlation.

Frequency Dependence A l l experiments are frequency dependent, and frequency dependence introduces many different processes that become the same in the static limit. Instead of the static field perturbation, e

0

• W,

consider the expansion of the

induced dipole analogous to that in equation 24, for a time-dependent, oscillatory field, e = £0 + 6^ cos IH =

ut,

(9{e,t)\r \9(e t)) i

i

= fli(o) + OLij(o, 0)£ j

+ QLij(-U] Uj)£ j

+

+ ^Ajfc(o; W, -V)£uj£u;k

Q

Pijk(0] O, o)£ j£ 0

+ Pijk{-^\^o)e je k LJ

0

ok

0

COS Ut

(41)

cos ut + jftjfc(-2o;;ci;,a;)e , -e tl

J

wfc

cos ut

+ ••• N o w , in addition to the static terms we have previously considered, we obtain a number of terms that correspond to different incoming and outgoing frequencies.

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

42

NONLINEAR OPTICAL MATERIALS

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

Table 5 Representative N o n - L i n e a r O p t i c a l Processes, w i t h C o r r e s p o n d i n g R e ­ sultant (u>uu)2 • • • ) Frequencies

Non-Linear

Optical

Process

First Static Hyperpolarizability Second Harmonic Generation ( S H G ) Electro-Optics Pockels Effect (EOPE) Optical Rectification (OR)

-UJJ

•UJ2

'U3

0

0

0

-

-2a;

UJ

UJ

-

-UJ

0

UJ

-

0

UJ

-UJ

UJ\

UJ2

-to a

Two-Wave M i x i n g

-

0

0

0

0

-3a;

UJ

UJ

UJ

Intensity-Dependent Refactive Index (IDRI)

-UJ

UJ

UJ

-UJ

Optical Kerr Effect ( O K E )

-UJ\

UJ\

UJ2

-UJ

0

0

UJ

-UJ

-2a;

0

UJ

UJ

-UJ

UJ

0

0

Second Static Hyperpolarizability Third Harmonic Generation ( T H G )

D.C.-Induced Optical Rectification (DCOR) D.C.-Induced S H G ( D C - S H G ) Electro-Optic Kerr Effect ( E O K E Three-Wave M i x i n g D.C.-Induced Two-Wave M i x i n g

2

UJ

UJ2

0

UJ2

{

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

Reliable Molecular

BARTLETT & SEKINO

For example, Pijk(—2uj]uj,uj)

43

Hyperpolarizabilities

corresponds to second harmonic generation ( S H G )

with incoming frequencies UJ + OJ resulting in an outgoing frequency, 2UJ. Similarly, corresponds to the dynamic polarizability, and Pijk(o]u,-UJ)

OLij(-u,u)

is called

optical rectification (OR), and Pijk(-uj\ UJ, o) corresponds to the electro optic Pockels effect ( E O P E ) . If we allowed the frequencies to be different, P(—UJ\

— UJ \ ^ 1 , ^ 2 ) 2

would correspond to two-wave mixing. Similarly, many such components occur in 7. A summary of them is shown in Table 5. A l l become the same in the static limit; hence, without inclusion of frequency dependence in the quantum mechanical method, we cannot distinguish between the different processes. Obviously, we can obtain each of these quantities from an appropriate derivative,

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

d

2

^

(42)

= U (- u,,-u>) jk

0]

etc. Just as in the static case, that derivative can be further related to a quasi-energy derivatives (45,46) from the Frenkel variational principle (77). The first way to augment a static calculation with some measure of dependence is to recognize that it may be rigorously shown (47)

frequency

that for

low

frequencies, UJ, UJ) = P(o; o, o) (1 + AUJ\ + BUJ\ + • • •)

P(-uj

a]

where UJI = W +UJ +UJ 2

2

(43)

and A and B are unknown constants. Without an independent

2

evaluation of A and B, the most we can conclude are the ratios of the various dispersion effects. For example, neglecting the smaller quartic term, the S H G and E O P E values are UJ) = P(o] o, o) (1 +

P(-2w\UJ,

= P(O;O,O)(1

P(UJ;UJ,O)

+

6AUJ ) 2

(44)

2AUJ ) 2

showing that their ratio,

l + 2Aa;

P(-UJ;UJ,O)

(45) 2

Obviously, S H G has a much greater dispersion effect than E O P E . Similarly, we have for 7, 7(-a/,;UJ ,UJ UJ3) 1

2)

=

7

( 0 ; 0 , 0 , 0 ) ( l + A'UJ

2

L

+ B'UJ

4

L

+ - • •)

(46)

The components of 7, O K E , I D R I , D C S H G and T H G correspond to the values 1 plus 2A'UJ , 2

4A UJ , 1

2

6A'UJ

2

and 12^4'a; , respectively, providing similar ratios. 2

Obviously, the degree of dispersion follows the order T H G > D C S H G > I D R I > O K E . However, we must at least know the constants to relate the various processes, particularly, to static quantities. It appears the only way at present to obtain any quantitative relationship is to evaluate the frequency dependent quantum chemical results.

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

44

NONLINEAR OPTICAL MATERIALS

Table 6 Comparison Hyperpolarizabilities

of

Static

and

Dynamic

Hartree

9

Fock

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

b

XM

(Kf

2 )

32

esu/molecule)

Xj| ^ (HF 3

39

esu/molecule)

SCF

TDHF

46

53.6

-

-

61

69.0

-

-

67

76.6

-

-

546

832

CO

9.1

10.5

85

102

HF

-2.3

-2.5

27

30

HC1

-1.3

-1.6

213

270

H 0

-4.7

-5.4

85

100

NH

-6.5

-9.3

200

280

0.58

0.64

470

690

SCF H

2

N

2

C0

TDHF

-

2

C H 2

2

3

H S 2

4

a

Values by T D H F for S H G and d c S H G at 694.3 nm.

b

Basis sets listed i n Tables 7 and 8.

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

BARTLETT & SEKINO

Reliable Molecular

45

Hyperpolarizabilities

The first such viable approach is the time-dependent Hartree-Fock ( T D H F ) method. Just as H F theory provides the simplest approach for static polarizabilities, T D H F provides the simplest ab initio solution to the time-dependent Schrodinger equation H(t)W(t)

= i^-tfp- and, thereby, frequency dependent processes. Just as in

the static case, we can impose a single determinant approximation for \I>(t) « $(£)> and by applying the time-dependent (Frenkel) variational/principle obtain the T D H F (or R P A ) equations.

The R P A equations date from long ago (48);

only pertains to excitation energies and a. to

7, 6, e, etc.

however, this

In 1986, we made the generalization

and showed that analytically

we can evaluate the frequency

dependent hyperpolarizabilities from the derivatives i n equation 36 in any order (11) giving ready T D H F access to hyperpolarizabilities. Others have since implemented

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

equivalent T D H F approaches for p (45) and for ft and

7 (32).

Recently, the

unrestricted Hartree-Fock (open shell) generalization has been made (49). Figs. 1 and 2 graphically illustrate the dispersion behavior for various processes, which reflect the relative numerical proportions. Table 6

shows the effect of dispersion in H F - l e v e l calculations of frequency

dependent polarizabilities to S H G and d c S H G processes. For x[|^ the average change 2

is 2 0 % and x[|^ is 26%. Obviously, from the static values the other dynamic processes 3

can show greater and lesser effects. Just as usual, we can take the derivative or S O S viewpoint i n T D H F . In the latter case, the excited T D H F states are the usual R P A ones

(24).

A t this point, we have a procedure based upon T D H F to evaluate the dispersion, and a procedure to add electron correlation to static hyperpolarizabilities. Clearly, both are critical in obtaining predictive values. Hence, guided by the fact that equa­ tions 37 and 40 are exact in the low-frequency limit, it makes sense to use a percentage T D H F dispersion correction (75) to augment static, correlated hyperpolarizabilities; namely, j8(^;a;i,w2)=j8(0;0 0)x

^ T D H F ( ^ ; ^ I ^ 2 )

>

7^;a;,a;i,(i^)

£HF(0;0,0)

= 7 ( 0 ; 0,0,0) x 7 H F

(47)

( 0 ; 0,0,0)

B y equating the percentage correction to AUJ\ at a particular frequency, a value of A could be extracted, as well; or fitting to several different processes, A and B. In the few cases where frequency dependent correlated results have been obtained (20,50,57), the percentage T D H F dispersion estimate has been well supported. However, it is clear that i f the T D H F = R P A result for the excitation energies are poor, then the slope of the curve in Figs. 1 and 2 w i l l have to eventually change to be able to approach the different asymptotic values of the excitation energy. Using this decoupled T D H F dispersion, static correlation procedure, we obtain the results shown in Tables 7 and 8.

A l l fall within 10% error of the experimental

result except for F H , whose errors are 2 8 % for x\\^ and 2 4 % for xf3^; for

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

C2H4,

46

NONLINEAR OPTICAL MATERIALS

Table 7 T h e o r e t i c a l vs. E x p e r i m e n t a l H y p e r p o l a r i z a b i l i t i e s of Molecules

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

3

b

J)

-

2

TDHF

MBPT(2)

ItfSHG

CCSD

CCSD(T)

EXPERIMENT

CO

10.5

11.2

11.4

11.7

12.9 ±1.4

HF

-2.5

-3.3

-3.2

-3.4

-4.70 ±0.41

HC1

-1.6

-4.0

-3.3

-3.8

-4.22 ±0.50

H 0

-5.4

-8.8

-8.2

-9.1

-9.4 ±0.4

NH

-9.3

-20.1

-18.4

-21.2

-20.9 ±0.5

0.6

-5.0

-3.4

-4.5

-4.3 ±0.9

2

3

H S 2

a

Value corrected for the dispersion effect at 694.3 nm using the formula,

X||corr(" ~ ° )

x

TDHF(u, =

0)

The calculations are performed with basis sets [5s3p2d] for C , N , O and F ; [7s5p2d] for S; and [3s2pld] for H . Lone-pair functions are added for H F , H 0 and N H 3 . For HC1, basis is [8s6p3dlf] for CI and [3s2pld] for H . A l l molecules at experimental geometries and there is no estimate of vibrational corrections. H S values in lone-pair augmented basis are 1.0, -4.4, -2.8 and -3.8, respectively (14). The lone-pair basis HC1 values are -1.0, -3.1, -2.6 and -3.0. 2

2

b

V a l u e obtained by dc-induced Second Harmonic Generation.

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

2.

BARTLETT & SEKINO

Reliable Molecular

47

Hyperpolarizabilities

Table 8 T h e o r e t i c a l vs. Experimental * Hyperpolarizabilities of Molecules 3

1

J * -

I-dcSHG

3

*ll TDHF H Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

N

2

54

2

69 77 822 102

C0

2

C H 2

CO

4

MBPT(2)

fi 'II CCSD -

CCSD(T)

EXP

59.3

59.3

60.5°

88.7

90.6

96

86.6 ±1.0

1.11.5

107.9

110

111.9 ±1.3

960

820

860

758 ±17

151

149

160

144 ±4

58.2

HF

30

52

49

53

70 ±10

HC1

270

364

352

374

347 ±15

H 0

100

180

170

180

194 ±10

NH

280

460

430

470

511 ±9

690

910

870

930

865 ±22

2

3

H S 2

a

Value corrected for the dispersion effect at 694.3 nm using the formula, n)

xj M =

xj|cL (LJ

= 0) x

TDHF(qj) T D H F ( o ; = 0)

The calculations are performed with basis sets [5s3p2d] for C , N , O and F ; [7s5p2d] for S; and [3s2pld] for H . The lone-pair functions are added for H F , H 2 O and N H . For HC1, basis is [8s6p3dlf] for CI and [3s2pld] for H . A l l molecules at experimental geometries and there is no estimate of vibrational corrections. The lone-pair augmented H S basis values are 731, 970, 930 and 980, respectively. The lone-pair basis HC1 values are 274, 369, 356 and 378. 3

2

b

V a l u e obtained by dc-induced Second Harmonic Generation.

E x a c t electronic value [D. M . Bishop, J . Pipin and S. M . Cybulski, Phys. Rev. A 4 3 , 4845 (1991)]. The experimental value of 67.2 includes a significant vibrational effect. c

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

48

NONLINEAR OPTICAL MATERIALS

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

Water-p

10.00

8.00

-

I •••

0.00

I

• • •

0.02

I

. . .

0.04

I

• • •

0.06

I

• • •

0.08

1

0.10

Frequency (a.u.) Figure 1. Water-/?: N L O Processes A s a Function of Frequency (a.u.) Trans Butadiene y

I

0.5 0.00

1 1.31



1 2.62



1 3.94

.

1 5.25

.

1

6.56xl0

2

Frequency(a.u.) Figure 2. Trans Butadiene 7: N L O Processes A s a Function of Frequency (a.u.)

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

Reliable Molecular

2. BARTLETT & SEKINO whose X||

49

Hyperpolarizabilities

is too large by 13.4% and N , and C O , whose error is a modest 11%. 2

We have discussed F H at length elsewhere (7,14), so we w i l l not repeat that here, except to say that we dispute the experimental values, expecting an error i n its determination. Extensive studies of the convergence of F H ' s hyperpolarizabilities to basis set, correlation, frequency dependence and vibrational corrections (14), show no convergence to the experimental values. Instead, we propose that the correct value for x[|^ is -3.6 ± 0 . 3 x l 0 " 2

32

esu/molecule and x[ ^ is 55 ± 5 x l 0 " 3

3 9

esu/molecule. F o r

the similar HC1 molecule, our C C S D ( T ) results fall within our goal of a 10% error. For

C2H4

and N , we can consider more recent or rescaled experimental values 2

(22). These are 76.6 ± 17 and 88.8 ± 1, respectively. This does reduce the C2B4 error to 12.3% and that for N

2

to 7.9%. Using the experimental results i n the tables, the

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

average error for C C S D ( T ) x[

3)

is 9.4%. Excluding F H , it becomes 7.3%. Similarly,

the average error for x\f* is 9.4%, or 5.7% without the F H example, falling within our 10% error target objective. The various contributions are shown i n Table 9. Clearly, correlation and frequency dependence are critical, and C C S D ( T ) is better than C C S D . However, it is gratifying that M B P T ( 2 ) , which is a lot cheaper than C C S D ( T ) , maintains about a 10% error. This level of correlation can be applied to much larger molecules (52) than can C C S D ( T ) .

Correlated Frequency Dependent Polarizabilities Despite the success of the decoupled correlated/TDHF results shown above, the most rigorous method would include the full coupling between correlation and dispersion. There have only been three attempts of this type for hyperpolarizabilities: a secondorder [MBPT(2)=MP(2)] level method (50); a multi-configurational ( M C S C F ) linear response approach (57); and our E O M - C C method (20). The first two are devel­ oped from the energy (or quasi-energy) derivative viewpoint, while the latter refers, conceptually, to the S O S expressions in equation 31. In other words, E O M - C C pro­ vides excited states, {V^},

their excitation energies, ^E — E^^j, and generalized 0

transition moments, ( V ' ^ V i l V ' o ) , from which the S O S expressions could be formally constructed. That form is particularly convenient, since frequency dependence can be trivially added to such an S O S expression. For example, for some frequency UJ,

a

.

i (

u

.

u

)

_ r

v

(^Ki^

Q )

)(^

Q )

hi^)

(^hi^

Q )

)(^

Q )

hi^)i

(

4

8

)

When UJ — E — EJ°^ = uj , we have a pole, whose residue is the dipole strength Q

k

The basic idea of E O M is very simple (53). Consider the solutions to the Schrodinger equation for an excited state,

and for the ground state, i\) , 0

Hoi>o = E if> 0

0

In Nonlinear Optical Materials; Karna, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1996.

(49a)

50

NONLINEAR OPTICAL MATERIALS

= E< tyW

Ho^

N o w we w i l l choose to write ip^ excitations from Vo-

(49b)

0

=

, where K

fc ip k

0

is an operator that creates

k

If we limit ourselves to single and double excitations, we

w i l l define an E O M - C C S D approximation. Inserting the excited state expression into equation 49b, multiplying equation 49a by 1Z from the left and subtracting, we obtain k

{HoHk ~ KkHoWo

=

(50a)

- E )K 1> 0

k

0

Downloaded by UNIV LAVAL on October 10, 2014 | http://pubs.acs.org Publication Date: May 5, 1996 | doi: 10.1021/bk-1996-0628.ch002

[Ho, Kktyo = VkfckTpo

with the commutator between H

and lZ ,

0

"equation-of-motion,"

[i.e.

k

(50b)

expression gives the

(H ,7lk)], 0

from the obvious connection with Heisenberg's form.

To

introduce C C theory, we simply choose for the correlated ground state, i/> = ipcc = 0

e x p ( T i + T 2 ) $ . A s TZk and T are all excitation operators, [JZ ,T ] n

0

k

= 0, and we

n

can commute the operators to give, [ e - f t e , n )$ r

r

0

k

= [Wo, n ]$

0

k

(5i)

= un$

0

k

k

0

recognizing that

: