<<

Coupled cluster theory: Analytic derivatives, molecular properties, and response theory

Wim Klopper and David P. Tew

Lehrstuhl für Theoretische Chemie Institut für Physikalische Chemie Universität Karlsruhe (TH)

C4 Tutorial, Zürich, 2–4 October 2006

Variational and non-variational wavefunctions

A wavefunction is referred to as variational if the electronic • energy function E(x, λ) fulfills the condition

∂E(x, λ) = 0 for all x ∂λ x is the molecular geometry or any other perturbational • parameter and λ represents the molecular electronic wavefuntion parameters (e.g., MO coefficients, CC amplitudes).

The electronic gradient vanishes at all geometries. • The variational condition determines λ as a function of x. • The molecular electronic energy is obtained by inserting the optimal λ into the energy function. ∗ Examples of variational wavefunctions

The Hartree–Fock wavefunction is variational since the orbital • rotation parameters κ (MO coefficients) are variational,

∂E (x, κ) HF = 0 for all x ∂κ

The MCSCF wavefunction is variational since the variational • condition is fulfilled both for the orbital rotation parameters κ (MO coefficients) and the state transfer parameters p (CI coefficients),

∂E (x, κ, p) ∂E (x, κ, p) MCSCF = 0, MCSCF = 0 for all x ∂κ ∂p

CI wavefunctions are not variational since the variational • condition is not fulfilled for the orbital rotation parameters κ,

∂E (x, κ, p) ∂E (x, κ, p) CI = 0, CI = 0 for all x ∂κ ! ∂p

Derivatives of variational wavefunctions

The molecular electronic energy is obtained by inserting • the optimal electronic wavefunction parameters (λ ) into ∗ the energy function,

ε(x) = E(x, λ ) ∗ We are interested in the first derivative • dε(x) ∂E(x, λ) ∂E(x, λ) ∂λ = + dx ∂x ∂λ ∂x ∂E(x, λ) dε(x) ∂E(x, λ) If = 0, then = ∂λ dx ∂x

We do not need the response of the variational • wavefunction! Hellmann–Feynman theorem

Assume that the (variational) energy function can be • written as an expectation value,

E(x, λ) = λ Hˆ (x) λ # | | $

We then obtain • dε(x) ∂Hˆ (x) = λ λ dx ! " ∂x " # " " " " " " This is the Hellmann–Feynman" theorem" . • Although originally stated for geometrical distortions, it holds for any perturbation.

Hellmann–Feynman theorem for Hartree–Fock Consider the Hamiltonian of a in a static electric • field , E Hˆ ( ) = Hˆ (0) µ E − · E Thus, in Hartree–Fock theory, the z-component of the • molecule’s dipole can be computed as

dε(x) ∂Hˆ ( ) = λ E λ = λ µz λ d z ! " ∂ z " # − # | | $ E " E " " " Concerning Hartree–Foc" k calculations" in a finite basis, • " " note that the Hellmann–Feynman theorem does not hold for a geometrical distortion (Ax),

dε(x) ∂Hˆ (x) = λ λ dAx ! ! " ∂Ax " # " " " " " " " " Hellmann–Feynman force and corrections

In Hartree–Fock theory, a geometrical distortion (A ) yields • x the Hellmann–Feynman force plus corrections,

dε(x) ∂Hˆ (x) = λ λ + corrections dAx ! " ∂Ax " # " " " N " " " x A = Z" λ" i − x λ + . . . − A r3 i=1 ' " A " ( & " " " " The reason is that the parameters" of the one-electron" • basis (exponents and contraction coefficients) are non-variational electronic wavefunction parameters. The corrections are sometimes called Pulay terms. •

Second derivative of variational wavefunctions

The variational condition also simplifies the calculation of • second derivatives,

d2ε(x) ∂2E ∂2E ∂λ ∂2E ∂λ 2 = + 2 + dx2 ∂x2 ∂x∂λ ∂x ∂λ2 ∂x $ % $ % The term ∂E/∂λ(∂λ2/∂x2) is eliminated by the variational • condition. We need the first-order response (∂λ/∂x) of the wave • function to calculate the energy to second order. 2n+1 rule: The derivatives of the wavefunction to order n • determine the derivatives of the energy to order 2n+1. Response equations

The calculation of second derivatives requires the • knowledge of the first-order response ∂λ/∂x. This first-order response is obtained by differentiating the • equations that determine the electronic wavefunction parameters λ. For variational wavefunctions, the variational condition • ∂E/∂λ = 0 determines the parameters λ. Thus,

d ∂E ∂2E ∂2E ∂λ = + = 0 dx ∂λ ∂x∂λ ∂λ2 ∂x $ % $ % We obtain a set of linear equations (response equations) • from which the first-order response may be determined.

Derivatives of non-variational wavefunctions

As an example, we consider the gradient of the CI energy, • which is variational w.r.t. the configuration coefficients p, but not w.r.t. the orbital rotations κ, ∂E (x, κ, p) ∂E (x, κ, p) CI = 0, CI = 0 ∂p ∂κ !

Therefore, if we differentiate the CI energy function w.r.t. x, • we do not obtain the simplifications of the 2n+1 rule,

dε (x) ∂E (x, κ, p) ∂E (x, κ, p) ∂κ CI = CI + CI dx ∂x ∂κ ∂x It appears that we need the first-order response of the • orbitals, ∂κ/∂x. First-order response of the orbitals

The orbital rotation parameters κ are determined by the • variational Hartree–Fock condition ∂E (x, κ) HF = 0 for all x ∂κ Thus, the first-order response of the orbitals ∂κ/∂x can be • determined by differentiating the Hartree–Fock condition with respect to x,

∂2E (x, κ) ∂κ ∂2E (x, κ) HF = HF ∂κ2 ∂x − ∂κ∂x $ % There is one set of response equations for each • perturbation, that is, for each independent geometrical distortion.

Lagrange’s method of undetermined multipliers

By regarding the variational Hartree–Fock condition as a • set of constraints in the optimization of the CI energy, we introduce the Lagrangian function

∂E (x, κ) L (x, κ, κ¯, p) = E (x, κ, p) + κ¯ HF CI CI ∂κ κ¯ are the Lagrange multipliers. The form of L is different • CI from ECI, but it gives the same energy when the Hartree–Fock condition is fulfilled. We adjust the multipliers so that L becomes variational • CI in all variables. The price we pay for this is that there is a larger number of variables. The variational Lagrangian

The Lagrangian function is variational in all variables, • ∂L (x, κ, κ¯, p) ∂E (x, κ, p) CI = CI = 0 ∂p ∂p ∂L (x, κ, κ¯, p) ∂E (x, κ) CI = HF = 0 ∂κ¯ ∂κ ∂L (x, κ, κ¯, p) ∂E (x, κ, p) ∂2E (x, κ) CI = CI + κ¯ HF = 0 ∂κ ∂κ ∂κ2 The last equation determines the Lagrange multipliers in • such a way that the Lagrangian is variational in κ. With the Lagrangian function, we have a completely • variational formulation of the CI energy, and the total derivative of the Lagrangian w.r.t. x is simply the corresponding partial derivative.

The total derivative of the Lagrangian

The total derivative of the CI energy can be computed from • the Lagrangian,

dε (x) dL (x, κ, κ¯, p) ∂L (x, κ, κ¯, p) CI = CI = CI dx dx ∂x ∂E (x, κ, p) ∂2E (x, κ) = CI + κ¯ HF ∂x ∂x∂κ The multipliers κ¯ are obtained from the equation • ∂2E (x, κ) ∂E (x, κ, p) HF κ¯ = CI ∂κ2 − ∂κ which does not depend on the perturbation x. The perturbation independent formulation is also known as • “Z-vector” or “interchange” method. The coupled-cluster Lagrangian

The coupled-cluster energy is neither variational in the • orbital rotations κ nor in the amplitudes tµ. Thus, we must introduce Langrange multipliers κ¯ and t¯µ,

∂E (x, κ) L (x, κ, κ¯, t , t¯ ) = E (x, κ, t )+κ¯ HF +t¯ Ω (x, κ, t ) CC µ µ CC µ ∂κ µ µ µ

where Ωµ(x, κ, tµ) is the coupled-cluster vector function

Ω (x, κ, t ) = µ Hˆ T (x) HF = µ exp( Tˆ)Hˆ (x) exp(Tˆ) HF µ µ # | | $ # | − | $ The coupled-cluster amplitudes equations are

Ωµ(x, κ, tµ) = 0 for all µ

Orbital-unrelaxed coupled-cluster properties

Let us first consider orbital-unrelaxed molecular properties. • Imagine that the perturbation is switched on only after the Hartree–Fock calculation. Thus, the orbitals are not changed by the perturbation and it suffices to consider the unrelaxed Lagrangian

LCC,unrelaxed(x, κ, tµ, t¯µ) = ECC(x, κ, tµ) + t¯µΩµ(x, κ, tµ)

The property can be obtained from • dL ∂L ∂E ∂Ω CC,unrelaxed = CC,unrelaxed = CC + t¯ µ dx ∂x ∂x µ ∂x (Here and in the following we omit arguments for clarity.) A simple unrelaxed one-electron property

Consider (again) the Hamiltonian of a molecule in a static • electric field , E Hˆ ( ) = Hˆ (0) µ E − · E The (orbital-unrelaxed) z-component of the molecule’s • dipole can be computed as ∂E ∂Ω µ = CC + t¯ µ z ∂ µ ∂ Ez Ez Note that the z-component of the molecule’s dipole can • also be computed by means of finite by adding the operator µ after the Hartree–Fock − zEz calculation has finished and before the coupled-cluster calculation has begun.

The coupled-cluster multipliers

The coupled-cluster multipliers are obtained by requiring • that the coupled-cluster Lagrangian is variational in the amplitudes,

∂LCC,unrelaxed ∂ECC ∂Ωµ ∂ECC = + t¯µ = + t¯µΩµν = 0 ∂tν ∂tν ∂tν ∂tν

where Ωµν is the coupled-cluster Jacobian,

Ω = µ exp( Tˆ) [Hˆ (x), τˆ ] exp(Tˆ) HF µν # | − ν | $ Furthermore, •

∂ECC = HF Hˆ (x)τˆν CC ∂tν # | | $ The coupled-cluster Hellmann–Feynman theorem Consider the following partial derivatives: • ∂E ∂Hˆ (x) CC = HF CC ∂x ! " ∂x " # " " " " ∂Ωµ " " ∂Hˆ (x) t¯µ = t¯µ µ" exp( T"ˆ) CC ∂x ! " − ∂x " # " " " " " " Thus, if we define a bra state" "

Λ = HF + t¯ µ exp( Tˆ) # | # | µ# | − we can write the total derivative of the Lagrangian as

dL ∂Hˆ (x) CC,unrelaxed = Λ CC dx ! " ∂x " # " " " " " " " "

A variational coupled-cluster energy

The usual expression for the coupled-cluster energy is • (now omitting the x-dependence of Hˆ (x))

E = HF Hˆ CC = HF Hˆ T HF CC # | | $ # | | $ Alternatively, we may compute the energy from • E = Λ Hˆ CC = HF Hˆ T HF + t¯ µ Hˆ T HF CC,var # | | $ # | | $ µ# | | $ HF + t¯ µ is the left eigenvector and HF is the right • # | µ# | | $ eigenvector of the similarity-transformed Hamiltonian Hˆ T . Of course, E is nothing but the CC Lagrangian. • CC,var The CC energy is less sensitive to numerical errors in the • amplitudes and multipliers when evaluated from ECC,var. Coupled-cluster density matrices

Recall that the Hamiltonian in is • ˆ 1 H = hnuc + hP QaP† aQ + 2 gP QRSaP† aR† aSaQ &P Q P&QRS Hence, the energy E = HF Hˆ CC can be written as CC # | | $ ¯ 1 ¯ ECC = DP QhP Q + 2 dP QRSgP QRS &P Q P&QRS D¯ = HF a† a CC , d¯ = HF a† a† a a CC P Q # | P Q| $ P QRS # | P R S Q| $ The coupled-cluster density matrices are not Hermitian • and may give complex eigenvalues upon diagonalization. For the energy, it is sufficient to consider the real symmetric part.

Coupled-cluster Lagrangian density matrices

The energy E = Λ Hˆ CC can be written as • CC,var # | | $ ¯ Λ 1 ¯Λ ECC = DP QhP Q + 2 dP QRSgP QRS &P Q P&QRS Λ Λ D¯ = Λ a† a CC , d¯ = Λ a† a† a a CC P Q # | P Q| $ P QRS # | P R S Q| $ In terms of the Lagrangian densities, we may calculate • coupled-cluster first-order properties in the same way as for variational wavefunctions, contracting the density matrix elements with the molecular integrals. The Lagrangian density matrices are also known as the • variational or relaxed density matrices. A biorthogonal basis

We introduce the notation •

(µ = µ exp( Tˆ) = HF τˆ† exp( Tˆ) | # | − # | µ − ν) = exp(Tˆ) ν = exp(Tˆ)τˆ HF | | $ ν| $ These states form a biorthogonal set, • (µ ν) = δ | µν For convenience, we identify τˆ as the identity operator, • 0

(0 = HF τˆ† exp( Tˆ) = HF exp( Tˆ) = HF = (HF | # | 0 − # | − # | | 0) = exp(Tˆ)τˆ HF = exp(Tˆ) HF = CC = HF) | 0| $ | $ | $ |

Matrix representation of the Hamiltonian

We consider the matrix representation of the molecular • electronic Hamiltonian Hˆ in the biorthogonal basis,

H H H = 00 0ν with µ, ν > 0 H H $ µ0 µν % Hˆ is an unsymmetric real matrix. • It follows that • H = (0 Hˆ 0) = HF exp( Tˆ)Hˆ exp(Tˆ) HF = E 00 | | # | − | $ CC H = (µ Hˆ 0) = µ exp( Tˆ)Hˆ exp(Tˆ) HF = Ω = 0 µ0 | | # | − | $ µ H = (0 Hˆ ν) = HF exp( Tˆ)Hˆ exp(Tˆ) ν = ∂E /∂t 0ν | | # | − | $ CC ν Left and right eigenvectors

E ∂E /∂t Apparently, H = CC CC ν with µ, ν > 0 0 H $ µν % 1 E is an eigenvalue of H with right eigenvector . • CC 0 $ % The left eigenvector (1 t¯ ) must fulfill • µ

∂ECC + t¯µ(Hµν δµνECC) = 0 ∂tν − Recall the multipliers equation, •

∂ECC + t¯µΩµν = 0 Ωµν = Hµν δµνECC ∂tν → −

The coupled-cluster Jacobian

Earlier, we have encountered the Jacobian • ∂Ωµ Ωµν = = µ exp( Tˆ) [Hˆ , τˆν] exp(Tˆ) HF ∂tν # | − | $ = (µ [Hˆ , τˆ ] HF) = (µ Hˆ ν) (µ τˆ Hˆ HF) | ν | | | − | ν | = H (µ τˆ Hˆ HF) µν − | ν | We invoke the resolution of the identity to show that • (µ τˆ Hˆ HF) = µ τˆ Hˆ T HF = µ τˆ ρ ρ Hˆ T HF | ν | # | ν | $ # | ν| $# | | $ ρ & = µ τˆ HF HF Hˆ T HF = δ E # | ν| $# | | $ µν CC Thus, the CC Jacobian occurs in the matrix representation • of the similarity-transformed Hamiltonian. Equation-of-motion CC theory (EOM-CC)

IN EOM-CC theory, we expand the excited states in the • space spanned by all µ), | c ) = ck µ) = ck exp(Tˆ) µ | k µ| µ | $ µ µ & & = ck exp(Tˆ)τˆ HF = ck τˆ exp(Tˆ) HF µ µ| $ µ µ | $ µ µ & & = ck τˆ CC = exp(Tˆ) ck τˆ HF µ µ| $ µ µ| $ µ µ & & The EOM-CC may be regarded as being • generated from a conventional expansion in Slater determinants by the application of an exponential operator containing the ground-state amplitudes.

The EOM-CC eigenvalue problem

In the biorthogonal basis, we may set up EOM-CC • wavefunctions of the form

c ) = ck ν), (c¯ = c¯k (µ | k ν | k| µ | ν µ & & and express the energy as a pseudo-expectation value

E = (c¯ Hˆ c ), with c¯Tc = δ k k| | k i j ij 0 0 For the ground state, we have c0 = 1 and cν = 0 for ν > 0. • 0 0 ¯ Also, c¯0 = 1 and c¯µ = tµ for µ > 0. Hence,

E = (c¯ Hˆ c ) = Λ Hˆ CC = E 0 0| | 0 # | | $ CC,var The EOM-CC eigenvalue problem

Differentiating the EOM-CC pseudo-expectation value w.r.t. • the ket and bra coefficients (assumed to be real), yields

Hck = Ekck T T c¯k H = c¯k Ek

Note that

(c¯ c ) = c¯ c = c¯Tc = δ i| j # i| j$ i j ij (c¯ Hˆ c ) = c¯ exp( Tˆ)Hˆ exp(Tˆ) c i| | j # i| − | j$ The EOM-CC states are obtained by diagonalizing the • unsymmetric matrix representation of the similarity- transformed Hamiltonian. The ground-state amplitudes are used in the similarity transformation.

Eigenvalues of the Jacobian

We shall level-shift the similarity-transformed Hamiltonian • by the ground-state energy E0. The eigenvalues will then correspond to the excitation energies,

0 ηT ∂E ∆H = H E 1 = , η = 0 = HF Hˆ τˆ CC − 0 0 Ω ν ∂t # | ν | $ $ % ν Thus, 0 ηT s ηTt s k = k = ∆E k 0 Ω t Ωt k t $ % $ k % $ k % $ k % The EOM-CC excitation energies correspond to the • eigenvalues of the CC Jacobian Ω. Since Ω is unsymmetric, there is no guarantee that the eigenvalues are real, but this is not a problem in practice. Some remarks on EOM-CC

The EOM-CC states are eigenvectors of the • similarity-transformed Hamiltonian (using ground-state amplitudes). The excitation energies are eigenvalues of the ground-state CC Jacobian. EOM-CC can be applied to the standard models CCSD, • CCSDT, etc. An EOM-CC calculation on two non-interacting systems A • and B will recover the excitation energies of A and B (size-intensivity), but simultaneous excitations in A and B are not size-intensive. For CCSD, CCSDT, etc., the EOM-CC excitation energies • are equal to those obtained from CC response theory.

Molecular gradients

So far, we have only considered orbital-unrelaxed molecular • properties. CC first-order properties can easily be computed from the pseudo-expectation value Λ Vˆ CC , that is, from the # | | $ corresponding variational density.

Next, consider a perturbation that changes the MOs (but not the • AOs). The orbital-relaxed approach is now required. Consider, for example, a static electric field that is switched on already in the Hartree–Fock step.

Matters become even more complicated when also the AO basis • is perturbed. This happens, for example, when derivatives are taken w.r.t. nuclear coordinates (molecular gradients), when the metric is changed by relativistic perturbations, or when GIAOs (London orbitals) are used for calculations of magnetic properties. Hartree–Fock orbitals The MOs are expanded in a basis of AOs, • P ϕP = cµ χµσ µ & Thus, the derivative w.r.t. a nuclear coordinate becomes

∂ϕ ∂cP ∂χ P = µ χ + cP µ σ ∂x ∂x u µ ∂x µ ) * & Changes occur in the MO coefficients and in the AOs. The • problem can be handled in a two-step procedure. At each geometry x, we write the orthonormal Hartree–Fock orbitals as

CHF(x) = COMO(x) U(x)

where U(x) is a unitary (or orthogonal) matrix and COMO(x) a basis of orthonormal molecular orbitals (OMOs).

OMOs and UMOs

At the reference geometry x , we choose U(x ) = 1 and • 0 0 COMO(x0) = CHF(x0).

If the geometry changes from x to x, the unmodified molecular • 0 orbitals (UMOs) are no longer orthonormal,

CUMO(x) = COMO(x0) S(x) = CT (x)S (x)C (x) = 1 UMO AO UMO ! We define orthonormalized molecular orbitals (OMOs), • 1/2 COMO(x) = CUMO(x)S− (x)

Of course,

T 1/2 1/2 S(x) = COMO(x)SAO(x)COMO(x) = S− (x)S(x)S− (x) = 1 An orthogonal orbital connection

The connection matrix S 1/2(x) connects orthonormal • − orbitals at neighbouring geometries. Rules that accomplish this are called orbital connections. We use the OMOs (not the Hartree–Fock orbitals) to define • a Fock space, in which we represent the Hamiltonian in second quantization,

ˆ ˜ 1 H(x) = hnuc(x) + hP Q(x)a˜P† a˜Q + 2 g˜P QRS(x)a˜P† a˜R† a˜Sa˜Q &P Q P&QRS with

1/2 a˜P† = aQ† S− (x) QP &Q + , 1/2 a˜P = aQ S− (x) P Q &Q + ,

Second quantization

We may ignore the geometry dependence of the creation • and annihilation operators!

ˆ ˜ 1 H(x) = hnuc(x) + hP Q(x)aP† aQ + 2 g˜P QRS(x)aP† aR† aSaQ &P Q P&QRS At each geometry, all matrix elements can be written as vacuum • expectation values of strings of operators. According to Wick’s theorem, only totally contracted terms contribute, depending only on the overlap between the orbitals. Since the OMOs are orthormal at all geometries, the vacuum expectation values are independent of the geometry. The geometry dependence of the Hamiltonian is isolated in the • integrals. First derivative of the one-electron Hamiltonian Consider • ∂ ˜ ∂ 1/2 1/2 hP Q(x)aP† aQ = S− (x)h(x)S− (x) aP† aQ ∂x ∂x P Q &P Q &P Q + , S(x) and h(x) are the overlap and Hamiltonian matrices at • the new geometry x0 + ∆x in the basis of the UMOs. When we expand around x , we get • 0 (0) (1) h(x0 + ∆x) = h (x0) + h (x0)∆x + . . . (1) S(x0 + ∆x) = 1 + S (x0)∆x + . . . 1/2 1 (1) S− (x + ∆x) = 1 S (x )∆x + . . . 0 − 2 0 (1) (1) where h (x0) and S (x0) are the first derivatives of h and S in the UMO basis, computed at the reference geometry x0.

One-index transformations Hence, • ˜ ∂hP Q(x) (1) 1 (1) (0) 1 (0) (1) = h (x0) 2 S (x0)h (x0) 2 h (x0)S (x0) ∂x " − − "x=x0 + , " " We may wr" ite this in a compact brace notation for one-index • transformations,

˜ ∂hP Q(x) ˜(1) (1) 1 (1) (0) = hP Q = hP Q 2 S , h ∂x " − P Q "x=x0 - . " where " "

A, B = (A B + A∗ B ) { }P Q P T T Q QT P T &T A, B = (A B + A∗ B { }P QRS P T T QRS QT P T RS &T + ART BP QT S + AS∗ T BP QRT ) The Hartree–Fock gradient With the AO-dependence isolated in the integrals of the • second-quantization Hamiltonian, we may write the Hartree–Fock gradient as

(1) (1) ˜(1) ¯ HF 1 (1) ¯HF EHF = Enuc + hP QDP Q + 2 g˜P QRS dP QRS &P Q P&QRS = E(1) + h˜(1) + 1 g˜(1) g˜(1) nuc II 2 IIJJ − IJJI &I &IJ / 0 = E(1) + h(1) h(0)S(1) nuc II − T I T I &I &IT + 1 g(1) g(1) g(0) g(0) S(1) 2 IIJJ − IJJI − T IJJ − T JJI T I &IJ / 0 I&JP / 0 = E(1) + h(1) + 1 g(1) g(1) f (0)S(1) nuc II 2 IIJJ − IJJI − T I T I &I &IJ / 0 &IT = E(1) + h(1) ε(0)S(1) + 1 g(1) g(1) nuc II − I II 2 IIJJ − IJJI &I / 0 &IJ / 0

Parametrization of the Hartree–Fock state The HF orbitals are obtained from the OMOs by a unitary • (or orhogonal) transformation,

CHF(x) = COMO(x) U(x), with U(x0) = 1

We can write U(x) = exp( κ), with κ = κ. − † − In second quantization, this translates into •

κˆ = κ a† a , κˆ† = κˆ P Q P Q − &P Q Spin and spatial restrictions may apply. In closed-shell HF • theory, one usually writes

κˆ = κ (a† a a† a ) = κ E− pq pσ qσ − qσ pσ pq pq p>q σ p>q & & & Orbitals at the displaced geometry

Geometry Orbital x a† vac = ϕHF 0 P | $ P x + ∆x a˜† vac = ϕOMO 0 P | $ P x + ∆x exp( κˆ)ϕOMO = exp( κˆ)a˜† vac = ϕHF 0 − P − P | $ P

HF At the new geometry x0 + ∆x, the HF orbital ϕP orbital is • HF replaced by ϕQ through

HF ϕ = exp( κˆ)a˜† a˜ a˜† vac Q − Q P P | $ = exp( κˆ)a˜† a˜ exp(κˆ) exp( κˆ)a˜† vac − Q P − P | $ Thus, the replacement operator is •

exp( κˆ)a˜† a˜ exp(κˆ) − Q P

CC energy at the displaced geometry

The CC energy at the displaced geometry is written as • E = OMO exp( Tˆ) exp(κˆ)Hˆ exp( κˆ) exp(Tˆ) OMO CC # | − − | $ The wavefunction parameters in κˆ and Tˆ do • depend on the geometry. The change of the AO basis is accounted for in Hˆ . • In the following, we shall write • E = 0 exp( Tˆ) exp(κˆ)Hˆ exp( κˆ) exp(Tˆ) 0 CC # | − − | $ with the Fermi vacuum 0 HF OMO at the | $ ≡ | $ ≡ $ reference geometry and expansion point x0. The closed-shell CC Lagrangian

We are now in the position to write the CC Lagrangian as • L = 0 exp( Tˆ) exp(κˆ)Hˆ exp( κˆ) exp(Tˆ) 0 CC # | − − | $ + t¯ µ exp( Tˆ) exp(κˆ)Hˆ exp( κˆ) exp(Tˆ) 0 µ# | − − | $ µ & + κ¯ (F δ ε ) pq pq − pq p p q &≥ where we have introduced the canonical condition, which helps to implement the frozen-core approximation. The orbital energies ε are treated as wavefuntion • p parameters. Derivatives of εp are not required according to the 2n+1 rule.

The closed-shell CC gradient

The CC gradient E(1) can be written as • CC (1) (1) ˜(1) eff 1 (1) eff ECC = Enuc + hpq Dpq + 2 g˜pqrs dpqrs pq pqrs & & (1) (1) eff 1 (1) eff = Enuc + hpq Dpq + 2 gpqrs dpqrs pq pqrs & & S(1)F eff − pq pq pq & eff eff where we have introduced effective densities Dpq and dpqrs and the effective Fock matrix

eff eff eff Fpq = Dpohoq + dporsgqors o ors & & Effective CC densities

The effective CC densities contain the CC Lagrangian • densities plus contributions from the orbital rotation multipliers κ¯pq,

eff D = Λ E CC + κ¯$ pq # | pq| $ pq eff HF HF d = Λ e CC + 2κ¯$ D κ¯$ D pqrs # | pqrs| $ pq rs − pr qs 1 with κ¯pq$ = 2 (1 + δpq)κ¯pq. The effective CC densities depend on the zeroth-order • wavefunction parameters and multipliers. The zeroth-order wavefunction parameters and multipliers • are obtained by making the Lagrangian stationary.

Coupled-perturbed Hartree–Fock (CPHF)

The diagonal zeroth-order orbital rotation multipliers are • obtained from requiring that ∂L/∂εp = 0. The off-diagonal zeroth-order orbital rotation multipliers are • obtained from the CPHF or Z-vector equations, which follow from ∂L/∂κrs = 0 for all r > s,

κ¯ A + Λ [Hˆ , E− ] CC = 0 pq pqrs # | rs | $ p q &≥ with A = 0 [a† , [a , [E− , Hˆ ]]] 0 pqrs − # | pσ qσ rs +| $ σ & Second derivatives

Wavefunction parameters follow the 2n+1 rule. • Multipliers follow the 2n+2 rule (since the Lagrangian is • linear in the multipliers). Hence, we need the first-order wavefunction parameters • (amplitudes and orbital rotation parameters) to compute second derivatives, but only zeroth-order multipliers,

ε(2) = E(20) + 2E(11)λ(1) + E(02) λ(1) 2 { } + λ¯(0) e(20) + 2e(11)λ(1) + e(02) λ(1) 2 { } + , ∂m+nE ∂m+ne E(mn) = , e(mn) = ∂xm∂λn ∂xm∂λn

The 2n+1 and 2n+2 rules

The Lagrangian is written as • L = E + λ¯e

where E is the energy and e the constraint e = 0. For the first derivative, we obtain • dL ∂L = = E(10) + λ¯(0)e(10) dx ∂x The second derivative is obtained from • d E(10) + E(01)λ(1) + λ¯(0)e(10) + λ¯(0)e(01)λ(1) + λ¯(1)e(00) dx - . Note that e(00) = 0 and E(01) + λ¯(0)e(01) = ∂L/∂λ = 0. The 2n+1 and 2n+2 rules The first-order response of the wavefunction parameters is • obtained from requiring that de/dx = 0. This yields

e(10) + e(01)λ(1) = 0 The second derivative yields • ε(2) = E(20) + E(11)λ(1) + λ¯(0)e(20) + λ¯(0)e(11)λ(1) + λ¯(1)e(10) + λ¯(2)e(00) + E(01)λ(2) + λ¯(0)e(10)λ(2) 2 + E(11)λ(1) + E(02) λ(1) + λ¯(0)e(11)λ(1) 2 - . + λ¯(0)e(02) λ(1) + λ¯(1)e(01)λ(1) - . 2 = E(20) + 2E(11)λ(1) + E(02) λ(1)

- . 2 + λ¯(0) e(20) + 2e(11)λ(1) + e(02) λ(1) $ - . %

The symmetric approach

Thus far, we have used the symmetric formula for second • derivatives w.r.t. 2 perturbations x and y

∂2E ∂λ ∂2E ∂λ 2E(11)λ(1) + ≡ ∂x∂λ ∂y ∂y∂λ ∂x $ % $ % In order to compute the second derivatives (e.g., the • molecular Hessian) of the CC energy, we need to solve

E(01) + λ¯(0)e(01) = 0 e(10) + e(01)λ(1) = 0 The zeroth-order multipliers equation is independent of the • perturbation, whereas the first-order wavefunction parameters are determined by a set of equations that involve the perturbation-dependent e(10) (w.r.t. x and y). Dalgarno’s interchange theorem

The asymmetric formula is obtained by considering the • total derivative of the gradient,

d ∂L d ε(2) = = E(010) + λ¯(00)e(010) dx ∂y dx 1 2 - . = E(110) + E(011)λ(10) + λ¯(10)e(010) + λ¯(00)e(110) + λ¯(00)e(011)λ(10) with ∂k+l+mE ∂m+nλ E(klm) = , λ(mn) = , etc. ∂xk∂yl∂λm ∂xm∂yn

For mixed second derivatives (NMR chemical shifts, IR • intensities) it is sufficient to consider only the responses λ(10) = ∂λ/∂x and λ¯(10) = ∂λ¯/∂x.

Time-dependent perturbations

In the following, we shall investigate a time-dependent • Hamiltonian of the form

Hˆ (t, #) = Hˆ (0) + Vˆ (t, #)

where Hˆ (0) is the unperturbed and Vˆ (t, #) the time-dependent perturbation, written as sum of Fourier components

N Vˆ (t, #) = Xˆ - (ω ) exp( iω t) j j j − j j= N &− The Xˆ are time-independent Hermitian operators, • j ω j = ωj, and -j(ω j) = -j(ωj)∗. Thus, Vˆ (t, #) is − − − Hermitian. Frequency-dependent response functions The time evolution of the observable A can be expressed • by means of response functions,

iω t Aˆ (t) = Aˆ + Aˆ; Xˆ e− j - (ω ) # $ # $0 ## j$$ωj j j j & 1 i(ω +ω )t + Aˆ; Xˆ , Xˆ e− j k - (ω )- (ω ) + . . . 2 ## j k$$ωj ,ωk j j k k &jk Examples include the (frequency-dependent) polarizability • µˆ ; µˆ and the first hyperpolarizability µˆ ; µˆ , µˆ . ## x y$$ω ## x y z$$ω1,ω2

Important symmetries: • Aˆ; Bˆ, Cˆ, . . . = Aˆ; Cˆ, Bˆ, . . . ## $$ωB ,ωC ,... ## $$ωC ,ωB ,... = Bˆ; A,ˆ Cˆ, . . . (ω +ω +... ),ω ,... ## $$− B C C ˆ ˆ ˆ ˆ ˆ ˆ A; B, C, . . . ωB ,ωC ,... = A; B, C, . . . ∗ ω , ω ,... ## $$ ## $$− B − C

Time-dependent Schrödinger equation

We write the time-dependent wavefunction 0¯ in the • | $ phase-isolated form

iF (t) 0¯ = e− 0˜ | $ | $ Note that also 0˜ is time-dependent. | $ The time-dependent Schrödinger equation becomes • Hˆ (t) i∂/∂t ∂F (t)/∂t 0˜ = 0 − − | $ - . Projection onto 0˜ yields • # | ∂F (t) Q(t) = 0˜ Hˆ (t) i∂/∂t 0˜ ∂t ≡ # | − | $ - . Time-dependent quasi-energy

We term Q(t) the time-dependent quasi-energy. Note that • in the time-independent limit, F (t) = Et and Q(t) = E. In CC response theory, we write 0˜ as • | $ 0˜ CC(t, #) = exp Tˆ(t, #) HF | $ ≡ | $ { }| $ All time-dependence is contained in the cluster operator • 3 Tˆ(t, #) (cf. orbital-unrelaxed properties). The CC Lagrangian is • L(t, #) = Λ˜(t, #) Hˆ (t) i∂/∂t CC(t, #) − 4 "- ." 5 Λ˜(t, #) = HF + " t¯µ(t, #) µ exp " Tˆ(t, #) # | # | " # | {"−3 } µ &

The Frenkel– variational principle

In the spirit of the Frenkel–Dirac variational principle • δΨ Hˆ (t) i∂/∂t Ψ = 0 # | − | $ we project the time-dependent Schrödinger equation onto HF and the excitation manifold µ exp Tˆ(t, #) , # | # | {− } Q(t) = HF Hˆ (t) CC(t, #) # | | $ 0 = µ exp Tˆ(t, #) Hˆ (t) i∂/∂t CC(t, #) # | {− 3 } − | $ - . The CC equations can be written as 3 • ∂t (t, #) Ω (t, #) i µ = 0 µ − ∂t Ω (t, #) = µ exp Tˆ(t, #) Hˆ (t) CC(t, #) µ # | {− } | $ 3 Response functions

The response functions are defined as the nth derivative of • the CC Lagrangian,

n ˆ ˆ ˆ 1 ˆ ω d L(t, #) X1; X2, . . . , Xn ω2,...,ωn = 2 C± ## $$ d-1(ω1)d-2(ω2) . . . d-n(ωn) with

ω Cˆ± f(ω , ω , . . . , ω ) = f(ω , ω , . . . , ω )+f( ω , ω , . . . , ω )∗ 1 2 n 1 2 n − 1 − 2 − n The cluster amplitudes are expanded in the Fourier • components of the perturbation(s),

(0) Xˆj iωj t tµ(t, #) = tµ + tµ (ωj)-j(ωj)e− j & 1 Xˆj Xˆk i(ωj +ωk)t + 2 tµ (ωj, ωk)-j(ωj)-k(ωk)e− + . . . &jk

The CC Jacobian

The amplitude responses are obtained from • ˆ ˆ ˆ ˆ (Ω ω1) tX1...Xn (ω , . . . , ω ) = ξX1...Xn (ω , . . . , ω ) − 1 n − 1 n with ω = ω + + ω and 1 · · · n n+1 Xˆ1...Xˆn ∂ L(t, #) ξµ (ω1, . . . , ωn) = ∂t¯µ∂-1(ω1)∂-2(ω2) . . . ∂-n(ωn) Ω is the Jacobian of the unperturbed system. • Similar equations determine the multiplier responses, • ˆ ˆ Xˆ ...Xˆ ¯tX1...Xn (ω , . . . , ω ) (Ω + ω1) = ξ¯ 1 n (ω , . . . , ω ) 1 n − 1 n Linear-response equations

Consider the linear response function Xˆ; Yˆ . • ## $$ω The amplitude and multiplier responses are obtained from • the equations

(Ω ω1) tY (ω) = ξY (ω) − − Y ¯tY (ω) (Ω + ω1) = ξ¯ (ω) = ηY (ω) + FtY (ω) − − with 6 7

2 2 2 Y ∂ L(t, #) Y ∂ L(t, #) ∂ L(t, #) ξµ (ω) = , ηµ (ω) = , Fµν (ω) = ∂t¯µ∂-Y (ω) ∂tµ∂-Y (ω) ∂tµ∂tν The same equations are obtained for orbital-unrelaxed • second derivatives except for the ω1 level shifts, which ± are due to the term i∂t (t, #)/∂t in the amplitudes − µ equation.

Linear-response functions Recall the (symmetric) expression for the static 2nd • derivative,

ε(2) = E(20) + 2E(11)λ(1) + E(02) λ(1) 2 { } + λ¯(0) e(20) + 2e(11)λ(1) + e(02) λ(1) 2 { } = L(20)++ 2L(11)λ(1) + L(02) λ(1) 2 , { } The frequency-dependent polarizability becomes • 1 ω x y α ( ω, ω) = x; y = Cˆ± t ( ω)η (ω) xy − ## $$ω 2 { − + ty(ω)ηx( ω) + tx( ω)Fty(ω) 1 ω y− x − x } y = Cˆ± t (ω)η ( ω) + t ( ω)ξ¯ (ω) 2 − − The asymmetric formula is8 9 • 1 ω y x y x x; y = Cˆ± t (ω)η ( ω) + ¯t (ω)ξ ( ω) ## $$ω 2 { − − } Poles and residues

(Ω ω1) tY (ω) = ξY (ω) − − Y ¯tY (ω) (Ω + ω1) = ξ¯ (ω) = ηY (ω) + FtY (ω) − − 6 7 The response equations become singular when ω is • ± equal to an eigenvalue of the CC Jacobian. Thus, these eigenvalues refer to an excitation energy. The residues are related to transition moments. • The 2n+1 and 2n+2 rules apply. • x; y is the orbital-unrelaxed static polarizability. • ## $$0 Frequency-dependent properties are obtained by • level-shifting the Jacobian in the response equations that determine the perturbed amplitudes and multipliers.