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

Similar documents
Time-independent molecular properties

Electric properties of molecules

Wave function methods for the electronic Schrödinger equation

Lecture 4: Hartree-Fock Theory

Introduction to Electronic Structure Theory

Introduction to multiconfigurational quantum chemistry. Emmanuel Fromager

Gradient Theory. unfamiliar with this particular formalism, the essential elements of second quantization are described in the Appendix (Section 9).

A Whirlwind Introduction to Coupled Cluster Response Theory. 1 Time-Independent Coupled Cluster Theory

Introduction to the calculation of molecular properties by response theory

Second quantization. Emmanuel Fromager

We also deduced that transformations between Slater determinants are always of the form

Chapter 2 Approximation Methods Can be Used When Exact Solutions to the Schrödinger Equation Can Not be Found.

MP463 QUANTUM MECHANICS

Attempts at relativistic QM

Time-dependent linear response theory: exact and approximate formulations. Emmanuel Fromager

Lecture 3: Quantum Satis*

The Klein-Gordon equation

I. CSFs Are Used to Express the Full N-Electron Wavefunction

7.1 Creation and annihilation operators

1 Mathematical preliminaries

Response Theory A (hopefully) gentle introduction

Lecture 5: More about one- Final words about the Hartree-Fock theory. First step above it by the Møller-Plesset perturbation theory.

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

Answers Quantum Chemistry NWI-MOL406 G. C. Groenenboom and G. A. de Wijs, HG00.307, 8:30-11:30, 21 jan 2014

AN INTRODUCTION TO QUANTUM CHEMISTRY. Mark S. Gordon Iowa State University

Lecture 5. Hartree-Fock Theory. WS2010/11: Introduction to Nuclear and Particle Physics

Recent Advances in the MRexpT approach

Introduction to Computational Chemistry

Pure and zero-point vibrational corrections to molecular properties

1 Rayleigh-Schrödinger Perturbation Theory

Derivatives and Properties

Time-dependent linear-response variational Monte Carlo.

DFT calculations of NMR indirect spin spin coupling constants

Introduction to DFTB. Marcus Elstner. July 28, 2006

Mean-Field Theory: HF and BCS

where P a is a projector to the eigenspace of A corresponding to a. 4. Time evolution of states is governed by the Schrödinger equation

Density Functional Theory: from theory to Applications

Density functional calculation of nuclear magnetic resonance chemical shifts

PHY 396 K. Problem set #5. Due October 9, 2008.

Department of Physics, Chemistry and Biology

Internally Contracted Multireference Coupled Cluster Theory. Computer Aided Implementation of Advanced Electronic Structure Methods.

The quantum state as a vector

Non-degenerate Perturbation Theory. and where one knows the eigenfunctions and eigenvalues of

2 Quantization of the Electromagnetic Field

Electron Correlation - Methods beyond Hartree-Fock

Introduction to Second-quantization I

Jack Simons, Henry Eyring Scientist and Professor Chemistry Department University of Utah

( r) = 1 Z. e Zr/a 0. + n +1δ n', n+1 ). dt ' e i ( ε n ε i )t'/! a n ( t) = n ψ t = 1 i! e iε n t/! n' x n = Physics 624, Quantum II -- Exam 1

1 The Quantum Anharmonic Oscillator

Second Quantization: Quantum Fields

Molecular Magnetic Properties

Explicitly correlated wave functions

INTRODUCTION TO QUANTUM ELECTRODYNAMICS by Lawrence R. Mead, Prof. Physics, USM

1 Infinite-Dimensional Vector Spaces

Ch 125a Problem Set 1

0 belonging to the unperturbed Hamiltonian H 0 are known

Second quantization (the occupation-number representation)

An Introduction to Quantum Chemistry and Potential Energy Surfaces. Benjamin G. Levine

Pseudo-Hermitian eigenvalue equations in linear-response electronic-structure theory

Time dependent perturbation theory 1 D. E. Soper 2 University of Oregon 11 May 2012

Theoretical Photochemistry WiSe 2017/18

Approximation Methods in QM

B2.III Revision notes: quantum physics

2. The Schrödinger equation for one-particle problems. 5. Atoms and the periodic table of chemical elements

0 + E (1) and the first correction to the ground state energy is given by

I. Perturbation Theory and the Problem of Degeneracy[?,?,?]

QFT PS1: Bosonic Annihilation and Creation Operators (11/10/17) 1

NANOSCALE SCIENCE & TECHNOLOGY

Systems of Identical Particles

Density matrix functional theory vis-á-vis density functional theory

Multiconfigurational Quantum Chemistry. Björn O. Roos as told by RL Department of Theoretical Chemistry Chemical Center Lund University Sweden

Physics 221A Fall 1996 Notes 13 Spins in Magnetic Fields

OVERVIEW OF QUANTUM CHEMISTRY METHODS

Special 5: Wavefunction Analysis and Visualization

THE DIRAC EQUATION (A REVIEW) We will try to find the relativistic wave equation for a particle.

Classical field theory 2012 (NS-364B) Feynman propagator

Quantum mechanics can be used to calculate any property of a molecule. The energy E of a wavefunction Ψ evaluated for the Hamiltonian H is,

P. W. Atkins and R. S. Friedman. Molecular Quantum Mechanics THIRD EDITION

. D CR Nomenclature D 1

Statistical Interpretation

Linear response time-dependent density functional theory

Nucleons in Nuclei: Interactions, Geometry, Symmetries

SCF calculation on HeH +

Semi-Classical Theory of Radiative Transitions

Ket space as a vector space over the complex numbers

( R)Ψ el ( r;r) = E el ( R)Ψ el ( r;r)

Going with the flow: A study of Lagrangian derivatives

Chapter 2. Linear Algebra. rather simple and learning them will eventually allow us to explain the strange results of

Hartree-Fock. It can also be viewed as the lowest order term in perturbative expansion (linear in U) with respect to interaction between electrons.

Green Functions in Many Body Quantum Mechanics

2 Electronic structure theory

Time Independent Perturbation Theory Contd.

Representation of the quantum and classical states of light carrying orbital angular momentum

1 Time-Dependent Two-State Systems: Rabi Oscillations

16.1. PROBLEM SET I 197

Other methods to consider electron correlation: Coupled-Cluster and Perturbation Theory

Consequently, the exact eigenfunctions of the Hamiltonian are also eigenfunctions of the two spin operators

Atomic electric dipole moment calculation with perturbed coupled cluster method

The Klein-Gordon Equation Meets the Cauchy Horizon

Advanced Electronic Structure Theory Density functional theory. Dr Fred Manby

Transcription:

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) C 4 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) λ = 0 for all x 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 HF (x, κ) κ = 0 for all x 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 MCSCF (x, κ, p) κ = 0, E MCSCF (x, κ, p) p = 0 for all x CI wavefunctions are not variational since the variational condition is not fulfilled for the orbital rotation parameters κ, E CI (x, κ, p) κ 0, E CI (x, κ, p) p = 0 for all x 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 If dε(x) dx E(x, λ) λ = E(x, λ) + E(x, λ) λ = 0, then dε(x) dx λ = E(x, λ) 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 Ĥ(x) λ We then obtain dε(x) dx = λ Ĥ(x) λ 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 molecule in a static electric field E, Ĥ(E) = Ĥ(0) µ E Thus, in Hartree Fock theory, the z-component of the molecule s dipole can be computed as dε(x) Ĥ(E) = λ de z E z λ = λ µ z λ Concerning Hartree Fock calculations in a finite basis, note that the Hellmann Feynman theorem does not hold for a geometrical distortion (A x ), dε(x) da x Ĥ(x) λ A x λ

Hellmann Feynman force and corrections In Hartree Fock theory, a geometrical distortion (A x ) yields the Hellmann Feynman force plus corrections, dε(x) Ĥ(x) = λ da x A x λ + corrections N = Z A λ x i A x ra 3 λ +... i=1 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, d 2 ε(x) dx 2 = 2 E 2 + 2 2 E λ ( ) λ + 2 E λ 2 ( ) λ 2 The term E/ λ( λ 2 / 2 ) is eliminated by the variational condition. We need the first-order response ( λ/) 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/. 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 = 2 E dx λ λ + 2 E λ 2 ( ) λ = 0 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 CI (x, κ, p) p = 0, E CI (x, κ, p) κ 0 Therefore, if we differentiate the CI energy function w.r.t. x, we do not obtain the simplifications of the 2n+1 rule, dε CI (x) dx = E CI(x, κ, p) + E CI(x, κ, p) κ κ It appears that we need the first-order response of the orbitals, κ/.

First-order response of the orbitals The orbital rotation parameters κ are determined by the variational Hartree Fock condition E HF (x, κ) κ = 0 for all x Thus, the first-order response of the orbitals κ/ can be determined by differentiating the Hartree Fock condition with respect to x, 2 E HF (x, κ) κ 2 ( ) κ = 2 E HF (x, κ) κ 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 L CI (x, κ, κ, p) = E CI (x, κ, p) + κ E HF(x, κ) κ κ are the Lagrange multipliers. The form of L CI is different from E CI, but it gives the same energy when the Hartree Fock condition is fulfilled. We adjust the multipliers so that L CI becomes variational 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 CI (x, κ, κ, p) p L CI (x, κ, κ, p) κ L CI (x, κ, κ, p) κ = E CI(x, κ, p) p = E HF(x, κ) κ = E CI(x, κ, p) κ = 0 = 0 + κ 2 E HF (x, κ) κ 2 = 0 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 CI (x) dx = dl CI(x, κ, κ, p) dx = E CI(x, κ, p) = L CI(x, κ, κ, p) + κ 2 E HF (x, κ) κ The multipliers κ are obtained from the equation 2 E HF (x, κ) κ 2 κ = E CI(x, κ, p) 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 t µ. Thus, we must introduce Langrange multipliers κ and t µ, L CC (x, κ, κ, t µ, t µ ) = E CC (x, κ, t µ )+ κ E HF(x, κ) + t µ Ω µ (x, κ, t µ ) κ where Ω µ (x, κ, t µ ) is the coupled-cluster vector function Ω µ (x, κ, t µ ) = µ Ĥ T (x) HF = µ exp( ˆT )Ĥ(x) exp( ˆT ) HF The coupled-cluster amplitudes equations are Ω µ (x, κ, t µ ) = 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 L CC,unrelaxed (x, κ, t µ, t µ ) = E CC (x, κ, t µ ) + t µ Ω µ (x, κ, t µ ) The property can be obtained from dl CC,unrelaxed dx = L CC,unrelaxed = E CC + Ω µ t µ (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, Ĥ(E) = Ĥ(0) µ E The (orbital-unrelaxed) z-component of the molecule s dipole can be computed as µ z = E CC E z + t µ Ω µ E z Note that the z-component of the molecule s dipole can also be computed by means of finite perturbation theory by adding the operator µ z E z after the Hartree Fock 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, L CC,unrelaxed t ν = E CC t ν + t µ Ω µ t ν = E CC t ν + t µ Ω µν = 0 where Ω µν is the coupled-cluster Jacobian, Furthermore, Ω µν = µ exp( ˆT ) [Ĥ(x), ˆτ ν ] exp( ˆT ) HF E CC t ν = HF Ĥ(x)ˆτ ν CC

The coupled-cluster Hellmann Feynman theorem Consider the following partial derivatives: E CC Ĥ(x) = HF CC Ω µ t µ = t µ µ exp( ˆT ) Ĥ(x) CC Thus, if we define a bra state Λ = HF + t µ µ exp( ˆT ) we can write the total derivative of the Lagrangian as dl CC,unrelaxed Ĥ(x) = Λ dx CC A variational coupled-cluster energy The usual expression for the coupled-cluster energy is (now omitting the x-dependence of Ĥ(x)) E CC = HF Ĥ CC = HF Ĥ T HF Alternatively, we may compute the energy from E CC,var = Λ Ĥ CC = HF Ĥ T HF + t µ µ Ĥ T HF HF + t µ µ is the left eigenvector and HF is the right eigenvector of the similarity-transformed Hamiltonian Ĥ T. Of course, E CC,var is nothing but the CC Lagrangian. The CC energy is less sensitive to numerical errors in the amplitudes and multipliers when evaluated from E CC,var.

Coupled-cluster density matrices Recall that the Hamiltonian in second quantization is Ĥ = h nuc + h P Q a P a Q + 1 2 g P QRS a P a R a Sa Q P Q P QRS Hence, the energy E CC = HF Ĥ CC can be written as E CC = D P Q h P Q + 1 2 P Q d P QRS g P QRS P QRS D P Q = HF a P a Q CC, d P QRS = HF a P a R a Sa Q CC 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 CC,var = Λ Ĥ CC can be written as E CC = P Q D Λ P Qh P Q + 1 2 d Λ P QRSg P QRS P QRS D Λ P Q = Λ a P a Q CC, d Λ P QRS = Λ a P a R a Sa Q CC 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( ˆT ) = HF ˆτ µ exp( ˆT ) ν) = exp( ˆT ) ν = exp( ˆT )ˆτ ν HF These states form a biorthogonal set, (µ ν) = δ µν For convenience, we identify ˆτ 0 as the identity operator, (0 = HF ˆτ 0 exp( ˆT ) = HF exp( ˆT ) = HF = (HF 0) = exp( ˆT )ˆτ 0 HF = exp( ˆT ) HF = CC = HF) Matrix representation of the Hamiltonian We consider the matrix representation of the molecular electronic Hamiltonian Ĥ in the biorthogonal basis, ( ) H00 H H = 0ν with µ, ν > 0 H µ0 H µν Ĥ is an unsymmetric real matrix. It follows that H 00 = (0 Ĥ 0) = HF exp( ˆT )Ĥ exp( ˆT ) HF = E CC H µ0 = (µ Ĥ 0) = µ exp( ˆT )Ĥ exp( ˆT ) HF = Ω µ = 0 H 0ν = (0 Ĥ ν) = HF exp( ˆT )Ĥ exp( ˆT ) ν = E CC / t ν

Left and right eigenvectors ( ECC E Apparently, H = CC / t ν 0 H µν E CC is an eigenvalue of H with right eigenvector The left eigenvector (1 t µ ) must fulfill ) with µ, ν > 0 ( 1 0 ). E CC t ν + t µ (H µν δ µν E CC ) = 0 Recall the multipliers equation, E CC t ν + t µ Ω µν = 0 Ω µν = H µν δ µν E CC The coupled-cluster Jacobian Earlier, we have encountered the Jacobian Ω µν = Ω µ = µ exp( ˆT ) [Ĥ, ˆτ ν ] exp( ˆT ) HF t ν = (µ [Ĥ, ˆτ ν ] HF) = (µ Ĥ ν) (µ ˆτ ν Ĥ HF) = H µν (µ ˆτ ν Ĥ HF) We invoke the resolution of the identity to show that (µ ˆτ ν Ĥ HF) = µ ˆτ ν Ĥ T HF = ρ µ ˆτ ν ρ ρ Ĥ T HF = µ ˆτ ν HF HF Ĥ T HF = δ µν E 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 k ) = µ c k µ µ) = µ c k µ exp( ˆT ) µ = µ c k µ exp( ˆT )ˆτ µ HF = µ c k µˆτ µ exp( ˆT ) HF = µ c k µˆτ µ CC = exp( ˆT ) µ c k µˆτ µ HF The EOM-CC excited state 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 k ) = ν c k ν ν), ( c k = µ c k µ (µ and express the energy as a pseudo-expectation value E k = ( c k Ĥ c k ), with c T i c j = δ ij For the ground state, we have c 0 0 = 1 and c0 ν = 0 for ν > 0. Also, c 0 0 = 1 and c0 µ = t µ for µ > 0. Hence, E 0 = ( c 0 Ĥ c 0 ) = Λ Ĥ CC = E 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 Note that Hc k = E k c k c T k H = ct k E k ( c i c j ) = c i c j = c T i c j = δ ij ( c i Ĥ c j ) = c i exp( ˆT )Ĥ exp( ˆT ) c j The EOM-CC states are obtained by diagonalizing the unsymmetric matrix representation of the similaritytransformed 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 E 0. The eigenvalues will then correspond to the excitation energies, ( ) 0 η T H = H E 0 1 =, η 0 Ω ν = E 0 = HF Ĥ ˆτ ν CC t ν Thus, ( 0 η T 0 Ω ) ( ) ( sk η = T t k t k Ωt k ) = E k ( sk t 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 ˆV 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, ϕ P = µ c P µ χ µ σ Thus, the derivative w.r.t. a nuclear coordinate becomes ϕ P = { } c P µ χ u + c P χ µ µ σ 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 C HF (x) = C OMO (x) U(x) where U(x) is a unitary (or orthogonal) matrix and C OMO (x) a basis of orthonormal molecular orbitals (OMOs). OMOs and UMOs At the reference geometry x 0, we choose U(x 0 ) = 1 and C OMO (x 0 ) = C HF (x 0 ). If the geometry changes from x 0 to x, the unmodified molecular orbitals (UMOs) are no longer orthonormal, C UMO (x) = C OMO (x 0 ) S(x) = C T UMO(x)S AO (x)c UMO (x) 1 We define orthonormalized molecular orbitals (OMOs), Of course, C OMO (x) = C UMO (x)s 1/2 (x) S(x) = C T OMO(x)S AO (x)c OMO (x) = S 1/2 (x)s(x)s 1/2 (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, Ĥ(x) = h nuc (x) + h P Q (x)ã P ãq + 1 2 g P QRS (x)ã P ã RãSã Q P Q P QRS with ã P = Q ã P = Q [ ] a Q S 1/2 (x) QP ] a Q [S 1/2 (x) P Q Second quantization We may ignore the geometry dependence of the creation and annihilation operators! Ĥ(x) = h nuc (x) + h P Q (x)a P a Q + 1 2 g P QRS (x)a P a R a Sa Q 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 P Q h P Q (x)a P a Q = P Q [ ] S 1/2 (x)h(x)s 1/2 (x) P Q a P a Q S(x) and h(x) are the overlap and Hamiltonian matrices at the new geometry x 0 + x in the basis of the UMOs. When we expand around x 0, we get h(x 0 + x) = h (0) (x 0 ) + h (1) (x 0 ) x +... S(x 0 + x) = 1 + S (1) (x 0 ) x +... S 1/2 (x 0 + x) = 1 1 2 S(1) (x 0 ) x +... where h (1) (x 0 ) and S (1) (x 0 ) are the first derivatives of h and S in the UMO basis, computed at the reference geometry x 0. Hence, h P Q (x) One-index transformations = x=x0 [ ] h (1) (x 0 ) 1 2 S(1) (x 0 )h (0) (x 0 ) 1 2 h(0) (x 0 )S (1) (x 0 ) We may write this in a compact brace notation for one-index transformations, h P Q (x) { = h (1) P Q = h(1) P Q 1 2 S (1), h (0)} P Q x=x0 where {A, B} P Q = T (A P T B T Q + A QT B P T ) {A, B} P QRS = T (A P T B T QRS + A QT B P T RS + A RT B P QT S + A ST B P 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 E (1) HF = E (1) nuc + h (1) HF P Q D P Q + 1 2 P Q = E (1) nuc + I = E (1) nuc + I + 1 2 IJ = E (1) nuc + I = E (1) nuc + I h (1) II + 1 2 h (1) II IT IJ P QRS g (1) P QRS ( g (1) IIJJ g(1) IJJI h (0) T I S(1) T I d HF P QRS ( ) g (1) IIJJ g(1) IJJI ( ) g (0) T IJJ g(0) T JJI S (1) T I IJP h (1) II + 1 2 ( h (1) IJ II ε(0) I ) ( ) g (1) IIJJ g(1) IJJI IT ) S(1) II + 1 2 IJ f (0) T I S(1) T I ( ) g (1) IIJJ g(1) IJJI Parametrization of the Hartree Fock state The HF orbitals are obtained from the OMOs by a unitary (or orhogonal) transformation, C HF (x) = C OMO (x) U(x), with U(x 0 ) = 1 We can write U(x) = exp( κ), with κ = κ. In second quantization, this translates into ˆκ = P Q κ P Q a P a Q, ˆκ = ˆκ Spin and spatial restrictions may apply. In closed-shell HF theory, one usually writes ˆκ = κ pq (a pσa qσ a qσa pσ ) = κ pq Epq p>q σ p>q

Orbitals at the displaced geometry Geometry x 0 x 0 + x x 0 + x Orbital a P vac = ϕhf P ã P vac = ϕomo P exp( ˆκ)ϕ OMO P = exp( ˆκ)ã P vac = ϕhf P At the new geometry x 0 + x, the HF orbital ϕ HF P replaced by ϕ HF Q through ϕ HF Q = exp( ˆκ)ã QãP ã P vac = exp( ˆκ)ã QãP exp(ˆκ) exp( ˆκ)ã P vac Thus, the replacement operator is exp( ˆκ)ã QãP exp(ˆκ) orbital is CC energy at the displaced geometry The CC energy at the displaced geometry is written as E CC = OMO exp( ˆT ) exp(ˆκ)ĥ exp( ˆκ) exp( ˆT ) OMO The wavefunction parameters in ˆκ and ˆT do depend on the geometry. The change of the AO basis is accounted for in Ĥ. In the following, we shall write E CC = 0 exp( ˆT ) exp(ˆκ)ĥ exp( ˆκ) exp( ˆT ) 0 with the Fermi vacuum 0 HF OMO at the reference geometry and expansion point x 0.

The closed-shell CC Lagrangian We are now in the position to write the CC Lagrangian as L CC = 0 exp( ˆT ) exp(ˆκ)ĥ exp( ˆκ) exp( ˆT ) 0 + t µ µ exp( ˆT ) exp(ˆκ)ĥ exp( ˆκ) exp( ˆT ) 0 µ + p q κ pq (F pq δ pq ε p ) where we have introduced the canonical condition, which helps to implement the frozen-core approximation. The orbital energies ε p are treated as wavefuntion parameters. Derivatives of ε p are not required according to the 2n+1 rule. The closed-shell CC gradient The CC gradient E (1) CC E (1) CC = E (1) nuc + pq = E (1) nuc pq can be written as + pq S pq (1) Fpq eff h (1) pq D eff pq + 1 2 h (1) pq D eff pq + 1 2 pqrs pqrs g (1) pqrs d eff pqrs g (1) pqrs d eff pqrs where we have introduced effective densities D eff pq and d eff pqrs and the effective Fock matrix F eff pq = o D eff poh oq + ors d eff porsg qors

Effective CC densities The effective CC densities contain the CC Lagrangian densities plus contributions from the orbital rotation multipliers κ pq, Dpq eff = Λ E pq CC + κ pq dpqrs eff = Λ e pqrs CC + 2 κ pqdrs HF κ prdqs HF with κ pq = 1 2 (1 + δ pq) 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/ ε p = 0. The off-diagonal zeroth-order orbital rotation multipliers are obtained from the CPHF or Z-vector equations, which follow from L/ κ rs = 0 for all r > s, κ pq A pqrs + Λ [Ĥ, Ers] CC = 0 p q with A pqrs = σ 0 [a pσ, [a qσ, [E rs, Ĥ]]] + 0

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] E (mn) = m+n E m λ n, e(mn) = m+n e m λ n The 2n+1 and 2n+2 rules The Lagrangian is written as L = E + λe where E is the energy and e the constraint e = 0. For the first derivative, we obtain dl dx = L = E(10) + λ (0) e (10) 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 The second derivative yields e (10) + e (01) λ (1) = 0 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) + E (11) λ (1) + E (02) { λ (1)} 2 + λ(0) e (11) λ (1) + λ (0) e (02) { λ (1)} 2 + λ(1) e (01) λ (1) = E (20) + 2E (11) λ (1) + E (02) { λ (1)} 2 + λ (0) ( e (20) + 2e (11) λ (1) + e (02) { λ (1)} 2 ) The symmetric approach Thus far, we have used the symmetric formula for second derivatives w.r.t. 2 perturbations x and y 2E (11) λ (1) 2 E λ ( ) λ + 2 E y y λ ( ) λ 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, ε (2) = d { } L = d {E (010) + λ (00) e (010)} dx y dx with = E (110) + E (011) λ (10) + λ (10) e (010) + λ (00) e (110) + λ (00) e (011) λ (10) E (klm) = k+l+m E k y l λ m, λ(mn) = m+n λ m y n, etc. For mixed second derivatives (NMR chemical shifts, IR intensities) it is sufficient to consider only the responses λ (10) = λ/ and λ (10) = λ/. Time-dependent perturbations In the following, we shall investigate a time-dependent Hamiltonian of the form Ĥ(t, ɛ) = Ĥ (0) + ˆV (t, ɛ) where Ĥ (0) is the unperturbed molecular Hamiltonian and ˆV (t, ɛ) the time-dependent perturbation, written as sum of Fourier components ˆV (t, ɛ) = N j= N ˆX j ɛ j (ω j ) exp( iω j t) The ˆX j are time-independent Hermitian operators, ω j = ω j, and ɛ j (ω j ) = ɛ j (ω j ). Thus, ˆV (t, ɛ) is Hermitian.

Frequency-dependent response functions The time evolution of the observable A can be expressed by means of response functions,  (t) =  0 + j Â; ˆX j ωj e iω jt ɛ j (ω j ) + 1 2 Â; ˆX j, ˆX k ωj,ω k e i(ω j+ω k )t ɛ j (ω j )ɛ k (ω k ) +... jk Examples include the (frequency-dependent) polarizability ˆµ x ; ˆµ y ω and the first hyperpolarizability ˆµ x ; ˆµ y, ˆµ z ω1,ω 2. Important symmetries: Â; ˆB, Ĉ,... ωb,ω C,... = Â; Ĉ, ˆB,... ωc,ω B,... = ˆB; Â, Ĉ,... (ωb +ω C +... ),ω C,... Â; ˆB, Ĉ,... ωb,ω C,... = Â; ˆB, Ĉ,... ω B, ω C,... Time-dependent Schrödinger equation We write the time-dependent wavefunction 0 in the phase-isolated form 0 = e if (t) 0 Note that also 0 is time-dependent. The time-dependent Schrödinger equation becomes {Ĥ(t) i / t F (t)/ t } 0 = 0 Projection onto 0 yields F (t) t } Q(t) = 0 {Ĥ(t) i / t 0

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, ɛ) = exp{ ˆT (t, ɛ)} HF All time-dependence is contained in the cluster operator ˆT (t, ɛ) (cf. orbital-unrelaxed properties). The CC Lagrangian is } L(t, ɛ) = Λ(t, ɛ) {Ĥ(t) i / t CC(t, ɛ) Λ(t, ɛ) = HF + µ t µ (t, ɛ) µ exp{ ˆT (t, ɛ)} The Frenkel Dirac variational principle In the spirit of the Frenkel Dirac variational principle δψ Ĥ(t) i / t Ψ = 0 we project the time-dependent Schrödinger equation onto HF and the excitation manifold µ exp{ ˆT (t, ɛ)}, Q(t) = HF Ĥ(t) CC(t, ɛ) } 0 = µ exp{ ˆT (t, ɛ)} {Ĥ(t) i / t CC(t, ɛ) The CC equations can be written as Ω µ (t, ɛ) i t µ(t, ɛ) = 0 t Ω µ (t, ɛ) = µ exp{ ˆT (t, ɛ)}ĥ(t) CC(t, ɛ)

Response functions The response functions are defined as the n th derivative of the CC Lagrangian, ˆX 1 ; ˆX 2,..., ˆX n ω2,...,ω n = 1 2Ĉ±ω d n L(t, ɛ) dɛ 1 (ω 1 )dɛ 2 (ω 2 )... dɛ n (ω n ) with Ĉ ±ω f(ω 1, ω 2,..., ω n ) = f(ω 1, ω 2,..., ω n )+f( ω 1, ω 2,..., ω n ) The cluster amplitudes are expanded in the Fourier components of the perturbation(s), t µ (t, ɛ) = t (0) µ + j + 1 2 jk t ˆX j µ (ω j )ɛ j (ω j )e iω jt t ˆX j ˆX k µ (ω j, ω k )ɛ j (ω j )ɛ k (ω k )e i(ω j+ω k )t +... The CC Jacobian The amplitude responses are obtained from (Ω ω1) t ˆX 1... ˆX n (ω 1,..., ω n ) = ξ ˆX 1... ˆX n (ω 1,..., ω n ) with ω = ω 1 + + ω n and ξ ˆX 1... ˆX n µ (ω 1,..., ω n ) = n+1 L(t, ɛ) t µ ɛ 1 (ω 1 ) ɛ 2 (ω 2 )... ɛ n (ω n ) Ω is the Jacobian of the unperturbed system. Similar equations determine the multiplier responses, t ˆX 1... ˆX n (ω 1,..., ω n ) (Ω + ω1) = ξ ˆX 1... ˆX n(ω1,..., ω n )

Linear-response equations Consider the linear response function ˆX; Ŷ ω. The amplitude and multiplier responses are obtained from the equations with (Ω ω1) t Y (ω) = ξ Y (ω) t Y (ω) (Ω + ω1) = ξ Y (ω) = ( η Y (ω) + Ft Y (ω) ) ξ Y µ (ω) = 2 L(t, ɛ) t µ ɛ Y (ω), ηy µ (ω) = 2 L(t, ɛ) t µ ɛ Y (ω), F µν(ω) = 2 L(t, ɛ) t µ t ν The same equations are obtained for orbital-unrelaxed second derivatives except for the ±ω1 level shifts, which are due to the term i t µ (t, ɛ)/ t in the amplitudes equation. Linear-response functions Recall the (symmetric) expression for the static 2 nd 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 α xy ( ω, ω) = x; y ω = 1 2Ĉ±ω {t x ( ω)η y (ω) The asymmetric formula is + t y (ω)η x ( ω) + t x ( ω)ft y (ω)} = 1 2Ĉ±ω { t y (ω)η x ( ω) + t x ( ω) ξ y (ω) } x; y ω = 1 2Ĉ±ω {t y (ω)η x ( ω) + t y (ω)ξ x ( ω)}

Poles and residues (Ω ω1) t Y (ω) = ξ Y (ω) t Y (ω) (Ω + ω1) = ξ Y (ω) = ( η Y (ω) + Ft Y (ω) ) 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 0 is the orbital-unrelaxed static polarizability. Frequency-dependent properties are obtained by level-shifting the Jacobian in the response equations that determine the perturbed amplitudes and multipliers.