general framework for calculating spin orbit couplings...

14
J. Chem. Phys. 151, 034106 (2019); https://doi.org/10.1063/1.5108762 151, 034106 © 2019 Author(s). General framework for calculating spin– orbit couplings using spinless one-particle density matrices: Theory and application to the equation-of-motion coupled-cluster wave functions Cite as: J. Chem. Phys. 151, 034106 (2019); https://doi.org/10.1063/1.5108762 Submitted: 02 May 2019 . Accepted: 20 June 2019 . Published Online: 15 July 2019 Pavel Pokhilko , Evgeny Epifanovsky, and Anna I. Krylov

Upload: others

Post on 03-Aug-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

J. Chem. Phys. 151, 034106 (2019); https://doi.org/10.1063/1.5108762 151, 034106

© 2019 Author(s).

General framework for calculating spin–orbit couplings using spinless one-particledensity matrices: Theory and applicationto the equation-of-motion coupled-clusterwave functionsCite as: J. Chem. Phys. 151, 034106 (2019); https://doi.org/10.1063/1.5108762Submitted: 02 May 2019 . Accepted: 20 June 2019 . Published Online: 15 July 2019

Pavel Pokhilko , Evgeny Epifanovsky, and Anna I. Krylov

Page 2: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

General framework for calculating spin–orbitcouplings using spinless one-particledensity matrices: Theory and applicationto the equation-of-motion coupled-clusterwave functions

Cite as: J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762Submitted: 2 May 2019 • Accepted: 20 June 2019 •Published Online: 15 July 2019

Pavel Pokhilko,1 Evgeny Epifanovsky,2 and Anna I. Krylov3,a)

AFFILIATIONS1Department of Chemistry, University of Southern California, Los Angeles, California 90089-0482, USA2Q-Chem, Inc., 6601 Owens Drive, Suite 105, Pleasanton, California 94588, USA3Department of Chemistry, University of Southern California, Los Angeles, California 90089-0482, USAand Institut für Physikalische Chemie, Johannes Gutenberg-Universität Mainz, Duesbergweg 10-14,D-55128 Mainz, Germany

a)Electronic mail: [email protected]

ABSTRACTStandard implementations of nonrelativistic excited-state calculations compute only one component of spin multiplets (i.e., Ms = 0 triplets);however, matrix elements for all components are necessary for deriving spin-dependent experimental observables. Wigner–Eckart’s theoremallows one to circumvent explicit calculations of all multiplet components. We generate all other spin–orbit matrix elements by applyingWigner–Eckart’s theorem to a reduced one-particle transition density matrix computed for a single multiplet component. In addition tocomputational efficiency, this approach also resolves the phase issue arising within Born–Oppenheimer’s separation of nuclear and electronicdegrees of freedom. A general formalism and its application to the calculation of spin–orbit couplings using equation-of-motion coupled-cluster wave functions are presented. The two-electron contributions are included via the mean-field spin–orbit treatment. Intrinsic issuesof constructing spin–orbit mean-field operators for open-shell references are discussed, and a resolution is proposed. The method is bench-marked by using several radicals and diradicals. The merits of the approach are illustrated by a calculation of the barrier for spin inversion ina high-spin tris(pyrrolylmethyl)amine Fe(II) complex.

Published under license by AIP Publishing. https://doi.org/10.1063/1.5108762., s

I. INTRODUCTION

Relativistic effects are important in chemistry and spec-troscopy.1,2 Related to the intrinsic magnetic moment of theelectron associated with its spin, they give rise to the orbitalcontraction, an increase in electron binding energies, and spin–orbitinteractions. The latter effect, which arises due to the coupling ofthe angular momentum of an electron and its spin, is of specialimportance. Although small in magnitude for molecules composed

of light atoms, spin–orbit coupling (SOC) leads to the mixing ofotherwise noninteracting states, e.g., singlets and triplets, and splitselectronic degeneracies, e.g., between degenerate Π or Δ states orbetween different Ms components of spin multiplets. Consequently,SOCs open new reaction channels (via intersystem crossing, ISC)and change the spectroscopic behavior by lighting up dark states viaintensity borrowing. SOCs lead to noticeable changes in the wavefunctions and electronic properties and are particularly importantin open-shell systems.

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-1

Published under license by AIP Publishing

Page 3: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

Spin-related properties, such as SOCs and spin–spin inter-actions, determine the magnetic behavior of molecules, which iscentral for understanding spectroscopic signatures of unpaired elec-trons (e.g., EPR spectroscopy), their macroscopic magnetic proper-ties (magnetozabilities), and magnetic relaxation times. The abilityto model these properties computationally is a key prerequisite forthe design of novel materials. For example, single-molecule magnets(SMMs), molecules with several unpaired electrons that have a high-spin ground state in the absence of an applied magnetic field,3–5 canbe used as building blocks to create novel, light-weight, and tunablemagnetic materials or as building blocks for quantum informationstorage and quantum computations.6–9 The key challenge in realiz-ing the full potential of SMMs is the ability to tune and control theirmagnetic behavior.

Magnetic properties are usually modeled using phenomeno-logical Hamiltonians of varying complexity, parameterized eitherempirically or by using first-principle calculations.5 In the contextof SMMs’ use in information storage, the key quantities needed forthe parameterization of these Hamiltonians are energy levels corre-sponding to different spin states and properties such as zero-fieldsplitting (ZFS), hyperfine (HF) couplings, g-tensors, and asymmet-ric Dzyaloshinskii–Moriya (DM) interactions. All these quantitiesdepend on SOCs, either directly or indirectly.

For example, SOC significantly contributes to ZFS, which canbe estimated perturbatively from the SOC between a selected statewith other closely lying states.10–13 ZFS can be (approximately)decomposed into single-ion anisotropy and exchange anisotropytensors, giving rise to effective Hamiltonians.5 These quantities ulti-mately determine the barrier for the reversal of magnetization andmagnetic relaxation times.

Fully relativistic schemes describe SOC in a natural manner.For light atoms, relativistic effects can be treated perturbativelyby using the Breit–Pauli (BP) Hamiltonian.14,15 Here, we consideronly such a perturbative scheme in which SOCs are computedas matrix elements of the BP Hamiltonian using nonrelativisticwave functions.2,16 Different implementations of full and approx-imate SOC treatments have been reported for different typesof wave functions1,17–31 including those obtained by completeactive-space self-consistent field (CASSCF),16–22 restricted activespace self-consistent field (RASSCF),29,30 multireference configura-tion interaction (MRCI),1,23 density-matrix renormalization group(DMRG),31 coupled-cluster (CC) response,25 equation-of-motionCC (EOM-CC),25,26,28 and multireference CC (MRCC) within theMk-MRCCSD formulation.27,32 Density functional theory (DFT)implementations have also been reported.24,33–36

While the working expressions for calculating the matrix ele-ments of the BP Hamiltonian between the pairs of nonrelativisticstates have been thoroughly described in previous works, one impor-tant aspect of the theory deserves further discussion. To constructexperimentally relevant quantities, such as rates of ISC, oscillatorstrengths, or magnetic anisotropies, one needs SOCs computed forall multiplet components. For example, the expression for the SOCconstant (SOCC), the quantity that enters the Fermi golden ruleexpression of the ISC rate, is given by the following expression:2

SOCC =√

∑M,M′∣⟨ΨSM ∣HSO∣Ψ′S′M′⟩∣

2, (1)

where the sum runs over all spin projections of both multiplets. Like-wise, calculations of the parameters for effective Hamiltonians, suchas magnetic anisotropy or DM tensors, also require matrix elementsbetween all multiplet components.

Previously reported implementations of SOC calculationswithin the EOM-CCSD framework25,26,28 enable calculationsbetween the selected target EOM states. In a typical nonrelativisticEOM-CC calculation (and in many other excited-state calculations),one computes only one component of the target multiplet becausethe other components are exactly degenerate. For example, tradi-tionally only Ms = 0 components of triplet states are computed.This is obviously insufficient for computing the full SOCC matrix.One possible strategy is to explicitly compute two other components(Ms = ±1), which requires additional coding efforts and increasesthe cost of the calculations. In addition, special care needs to betaken to synchronize the phases of all computed states.37 Alterna-tively, one can apply coordinate rotations (which is equivalent tochanging the quantization axis) and extract the SOC matrix ele-ments in the original coordinate frame from the matrix elements inthe rotated frames. However, practical application of this approachis also complicated by the phase problem mentioned above: thephases of the wave functions computed in different coordinateframes are arbitrary. In addition, frame rotations may require dis-abling point-group symmetry, thus further increasing the cost ofcalculations.

Here, we describe how to generate all required SOC matrixelements with a minimal amount of calculations. The theory isformulated in terms of reduced one-particle density matrices, inspin-orbital representation, such that it is ansatz-agnostic and canbe applied to any electronic structure method that can furnishthe required transition densities. We illustrate this approach byusing the EOM-CC method with single and double excitations(EOM-CCSD) as an example.

The formalism allows one to compute an entire SOC matrixbetween all multiplet components at once, without additional cal-culations in rotated frames or explicit calculations of the othermultiplet components, and without the need to address the phaseproblem. In essence, the approach is grounded in Wigner–Eckart’stheorem38–40 that outlines the relationships between the matrix ele-ments involving multiplet components. In the context of SOC, theutility of Wigner–Eckart’s theorem has been recognized in manyprevious studies;19,22,23,29–31,33,36,41–43 however, the concrete strategiesof its application vary considerably. Reported implementations dif-fer by whether Wigner–Eckart’s theorem is applied to the states ortransition densities, which multiplet components are used as gener-ators, and whether spinless or spin-orbital expressions are used toconstruct the working equations. For example, within the configu-ration interaction formalism, an algorithm for generating the entireSOC matrix from only three reduced matrix elements has beendescribed by Fedorov.20,41 However, as explained below, Fedorov’sapproach20,41 requires access to several multiplet components(Ms = 0, ±1), a feature which is not commonly available in excited-state codes; it is also affected by the phase issue unless specialcare is taken. A similar strategy has been used in the Orca pack-age, where a highest-spin multiplet component is used as a gen-erator.12,33,36,44 An alternative approach, based on using transitiondensities for the spinless states, was used in the Molcas implemen-tation within the restricted active space state interaction (RAS-SI)

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-2

Published under license by AIP Publishing

Page 4: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

framework29,30 and in a recent DMRG implementation;31 in thesestudies, the required transition densities were obtained from spinlessstates.

Although not always acknowledged, the most general strategybased on the application of Wigner–Eckart’s theorem to transi-tion density was originally outlined by McWeeny in 1965.45,46 Here,we follow McWeeny’s proposal and develop second quantized for-mulation of this approach. The essential components of our formal-ism are very similar to the Molcas and DMRG implementations.29,31

In our approach, one needs to compute only one multiplet com-ponent (for example, the Ms = 0 state) from which the reducedspinless density matrix is generated. In contrast to Refs. 29 and 31,we employ a spin-orbital formalism, which allows us to obtain therequisite reduced transition densities for the transitions betweenany types of open-shell states. This density matrix is then used tocompute a reduced matrix element, and the entire SOC matrix isobtained by the application of Wigner–Eckart’s theorem. The phaseproblem is avoided by construction because the density matrix cor-responding to the reduced matrix element is computed from onlyone transition between the pair of target states (or between the ref-erence and the target EOM state). The theory is illustrated by appli-cation to the EOM-CCSD wave functions. As numerical examples,we report calculations for a set of small open-shell molecules, con-sidered as candidates for laser cooling experiments, a set of dirad-icals used as benchmark systems in previous studies,25,28 severaltransition-metal ions, and a Fe(II) SMM with a large spin-reversalbarrier (magnetic anisotropy) for which a synthetic analog has beensynthesized.47

The structure of the paper is as follows. In Sec. II, we outlineessential features of the EOM-CCSD method and the calculationof the state and transition properties within EOM-CC. We thendescribe the spin-tensor formalism and apply it to derive the keyequations for calculation of the SOC matrix elements via reduceddensity matrices and Wigner–Eckart’s theorem. We also discuss theimplementation of the algorithm within the EOM-CC suite of meth-ods in the Q-Chem electronic structure package.48,49 We discussadditional aspects of the theory in the case of open-shell references.Section III presents illustrative calculations.

II. THEORYA. Equation-of-motion coupled-cluster methodswith single and double excitations

In the EOM-CC framework,50–54 the target states are parame-terized as

∣ΨR⟩ = ReT ∣Φ0⟩, (2)

where Φ0 is a reference determinant, which defines the many-bodyvacuum state and, consequently, the occupied and virtual orbitalspaces,55 eT |Φ0⟩ is a CC wave function, and R is a general exci-tation operator.56 The explicit form of R depends on the level ofcorrelation included in the model (e.g., it generates up to doublyexcited configurations at the EOM-CCSD level) and on the type ofthe target states. For example, in EOM-CCSD for excitation ener-gies,57 R is spin- and particle-conserving, i.e., of 1h1p and 2h2ptypes at the CCSD level, with an additional constraint that in eachoperator, the number of α holes equals the number of α particles

and the number of β holes equals the number of β particles.In EOM-CCSD for ionized58,59 (EOM-IP-CCSD) or electron-attached60 (EOM-EA-CCSD) states, R’s are not particle conserv-ing. In EOM-IP-CCSD, R is of 1h and 2h1p types, and in EOM-EA-CCSD, R is of 1p and 2p1h types. In the spin-flip variant61,62

(EOM-SF-CCSD), operators R change the spin-projection by flip-ping the spin of an electron. Other variants include double-ionization potential (DIP),63,64 double electron-attachment (DEA),and double spin-flip (DSF)65 ansätze. R amplitudes are foundthrough the diagonalization of the similarity-transformed Hamilto-nian,H = e−THe−T , in the space of the target set of determinants,

HRK = EKRK , (3)

L†KH = EKL

†K . (4)

Because H is non-Hermitian, its right and left eigenvectors arenot Hermitian conjugates of each other but form a biorthogonalset,

⟨ΨLI ∣Ψ

RJ ⟩ = δIJ . (5)

The normalization and phases within the right and left sets are arbi-trary, but the biorthogonality condition [Eq. (5)] fixes the relativenorms and phases, i.e., if one changes the sign and norm of RK ,the sign and norm of LK changes accordingly. In many implementa-tions, the diagonalization is carried out independently for the rightand left vectors and the biorthogonality condition, which synchro-nizes the phases, is applied a posteriori (in the case of degenerateeigenstates, one needs to rotate the degenerate eigenstates such thatthe overlap matrix between the left and right degenerate manifoldsbecomes a unit matrix).

In many-body theories, the state and transition propertiesare computed as contractions of the corresponding integrals withreduced density matrices.66 One-electron operators require a one-electron density matrix,67

⟨ΨI ∣A∣ΨI′⟩ = Tr[γ(I, I′)A] =∑pq

γpqApq, (6)

and two-electron operators require a two-electron density matrix.(Here, A denotes an operator describing a specific property.) Forthe exact states or within the expectation-value formulation ofproperties,68,69 the one-particle density matrix γ used for propertycalculation is simply

γpq(I, I′) = ⟨ΨI ∣a†paq∣ΨI′⟩. (7)

Within the response formulation of properties, the relaxed one-particle density matrices are used; they include additional termsdescribing relaxation of nonvariational wave function parame-ters (e.g., CC amplitudes). Here, we follow the expectation-valueapproach, as in our previous work.28,70

Because of a non-Hermitian nature of EOM-CC, γpq(I, I′)≠ γqp(I′, I), which leads to different numerical results for ⟨ΨI |A|ΨI ′⟩

and ⟨ΨI ′ |A|ΨI⟩. A common resolution of this problem is to takea geometric average.57 However, this treatment should be mod-ified for complex-valued tensor properties, in order to preservephase consistency and to generalize it for a matrix. In Sec. II C, weintroduce a spinless density matrix, which enables generation of all

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-3

Published under license by AIP Publishing

Page 5: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

phase-consistent SOCs between the selected pairs of states. We alsopropose a different resolution of the averaging problem.

Finally, we would like to comment on spin contamination ofEOM-CC states.71 Although the standard formulation of EOM-CC51

does not involve explicit spin-adaptation, the CC/EOM-CC statesare spin pure when closed-shell reference is used. For example,EOM-IP/EA wave functions of doublet states are naturally spin-adapted, even when the spin symmetry is not explicitly enforced.For closed-shell references, even in spin-orbital implementations,the spin-symmetry can be exploited in the same fashion as per-mutational or point-group symmetries, at the level of tensorlibrary.72 Only when open-shell references are used in EOM cal-culations, for example, in spin-flip calculations or in EOM-EE cal-culations using doublet references, spin contamination becomesan issue.71 Recently, we discussed the impact of spin contam-ination on SOCs between the singlet and triplet states of theCvetanovic diradicals.73 By using an approximate a posteriori spin-projection technique, we have shown73 that in these species, theimpact of spin contamination on SOCCs is small, i.e., less than1 cm−1.

B. Spin properties of operatorsIn wave function approaches, spin is usually described as a

property of the wave function. However, one can consider proper-ties of a spin operator S alone. It has three Cartesian componentsand transforms under rotations as a vector. Operators involving spinmay obey other transformation rules. For example, the S2 operator,formed as a dot product of the two spin operators, is a scalar oper-ator: it has only one component, and it does not change upon therotation of the coordinate frame. The spin–orbit operator of the BPHamiltonian is

HSOBP =

12c2

⎝∑ihSO(i) ⋅ s(i) −∑

i≠jhSOO(i, j) ⋅ ( s(i) + 2s( j)

⎠, (8)

hSO =∑K

ZK(ri − RK) × pi∣ri − RK ∣3

=∑K

Zk

r3iK(riK × pi), (9)

hSOO(i, j) =(ri − rj) × pi∣ri − rj∣3

=∑i≠j

1r3ij(rij × pi), (10)

where the coordinates and momenta of electron i are denoted by riand pi, the charge and coordinates of nucleus K are denoted by ZKand RK , and the relative coordinates of electron i and electron j ornucleus K are denoted by rij and riK . Because of the dot product, thespin–orbit operator is a scalar operator. Therefore, its componentstransform as L⋅S upon the rotation of the coordinate frame. Here, wedo not consider spatial rotations but focus on the transformations inthe spin space.

Representation theory tells us that the objects (wave functionsor operators) transform according to some representations of thetransformation group. These representations are irreducible if theycannot be decomposed into other representations. This leads toirreducible tensor operators, which are especially useful for second-quantized formulations. It is convenient to define irreducible spin-tensor operators ÔS ,M through commutation relations,69,74

[S±, OS,M] =√S(S + 1) −M(M ± 1)OS,M±1, (11)

[Sz , OS,M] =MOS,M . (12)

Here, S denotes the spin value and M denotes the spin projection,which goes from −S, −S + 1, . . ., S − 1, S. Any set of operators sat-isfying these relations are called irreducible spin-tensor operators.For example, the S = 0 case corresponds to singlet operators, whichcommute with Sz , S+, and S−. Any spin-independent operator, suchas a dipole moment operator, is a singlet operator. The spin–orbitoperator depends on spin, and, as shown below, it is composed oftriplet spin-tensor operators. In particular, McWeeny writes downboth angular momentum and spin in an irreducible tensor form forthe spin–orbit operator in Eqs. (2.1) and (2.2) in Ref. 45 (see alsoexercise 2.10 in Ref. 69).

Because we are interested in second-quantized formulation, itis useful to express spin in second quantization and consider severalimportant examples. If α and β spin-orbitals have the same spatialpart,

S± = Sx ± iSy, (13)

S+ =∑pa†pαapβ, (14)

S− =∑pa†pβapα, (15)

Sz =12∑p(a†

pαapα − a†pβapβ). (16)

One important example is excitation operators. A one-particle sin-glet excitation operator can be written as

T0,0pq =

1√

2(a†

pαaqα + a†pβaqβ), (17)

and a set of triplet operators can be written as

T1,1pq = −a

†pαaqβ, (18)

T1,0pq =

1√

2(a†

pαaqα − a†pβaqβ), (19)

T1,−1pq = a

†pβaqα. (20)

Note that S± and Sz are not expressed in the irreducible spin-tensorform. If one wishes to use the irreducible form, these operatorsshould be multiplied by factors from Eqs. (18)–(20).

The spin parts of all terms in the spin–orbit operator are tripletoperators. The one-electron part can be written as

H1eSO=

12⎛

⎝∑pq

h1eSOL+ ,pqa

†pβaqα +∑

pqh1eSOz,pq (a

†pαaqα − a

†pβaqβ)

+∑pq

h1eSOL− ,pqa

†pαaqβ

⎠, (21)

where

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-4

Published under license by AIP Publishing

Page 6: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

h1eSOL+ = h1eSO

x + ih1eSOy , (22)

h1eSOL− = h1eSO

x − ih1eSOy . (23)

The one-electron part requires only one-electron transition den-sity matrix for the final SOC matrix elements. The two-electronpart requires the spin–orbit two-electron integrals and two-electrontransition densities, which makes the full calculation expensive.However, the two-electron contribution can be effectively approx-imated by considering only the separable part of the two-electrontransition density such that the two-electron spin–orbit integralsare first contracted with the density of the reference determinant(usually Hartree–Fock determinant), and the resulting mean-field-like one-electron operator is then folded into the one-electronpart. This spin–orbit mean-field (SOMF)2,75 approximation is com-monly used in calculations of SOCs;2,23–26,28,76–79 it is describedin Sec. II D.

One of the applications of irreducible tensor operators isWigner–Eckart’s theorem. The theorem allows one to computematrix elements through a reduced matrix element, which does notdepend on spin projection,

⟨S′M′∣OS,M∣S′′M′′⟩ = ⟨S′′M′′; SM∣S′M′⟩⟨S′∣∣OS,⋅

∣∣S′′⟩. (24)

Here, the matrix element between the bra state with spin S′ and spinprojection M′ and the ket state with spin S′′ and spin projection M′′

is expressed through a Clebsh–Gordan coefficient ⟨S′′M′′; SM|S′M′⟩and a reduced matrix element ⟨S′||OS,⋅||S′′⟩. Here, superscript S, ⋅signifies the reduced matrix element, which does not depend on spinprojection.

C. Evaluation of matrix elementsTo compute all required matrix elements between the two

multiplets, we apply Wigner–Eckart’s theorem in the spin space,

⟨ISS′z − 1∣HL+S− ∣I′S′S′z⟩ = ⟨S

′S′z1 − 1∣SSz − 1⟩⟨IS∣∣HL+ ∣∣L′S′⟩, (25)

⟨ISS′z ∣HLzSz ∣I′S′S′z⟩ = ⟨S

′S′z10∣SSz⟩⟨IS∣∣HLz ∣∣L′S′⟩, (26)

⟨ISS′z + 1∣HL−S+ ∣I′S′S′z⟩ = ⟨S

′S′z11∣SS′z + 1⟩⟨IS∣∣HL− ∣∣L′S′⟩, (27)

where I and I′ enumerate multiplets. This strategy has been used,for example, within the CI framework,41 where the full spin–orbitmatrix between I and I′ spin manifolds was constructed from three(or two, if symmetry is used) reduced matrix elements. An appli-cation of this approach41 to EOM-CCSD would require calcula-tion of target states of different spin projections (with consistentphases).

Rather than applying Wigner–Eckart’s theorem to the spin–orbit operator directly, we apply it to general triplet excitationoperators, as was done in Refs. 29, 31, 45, and 46,

⟨I′S′M′∣T1,Mpq ∣I

′′S′′M′′⟩ = ⟨S′′M′′; 1M∣S′M′⟩⟨I′S′∣∣T1,⋅pq ∣∣I

′′S′′⟩. (28)

These reduced matrix elements form a new density, which we denoteas u,

upq ≡ ⟨I′S′∣∣T1,⋅pq ∣∣I

′′S′′⟩. (29)

Being a reduced matrix element, upq does not depend on the spinprojections of the states or the excitation operator. To emphasizethis property, we drop the corresponding index in the excitationoperator and replace it with a dot: T

1,⋅pq. If the transition between

the target states (or between the reference and target states) is aspin-flip transition, the spin-tensor form of the one-particle tran-sition density matrix is the transition density matrix γ up to asign; see Eqs. (18) and (20). If the transition is spin-conserving, thetransition density matrix always can be decomposed into the sin-glet and triplet components in any spin-adapted basis (i.e., atomicorbitals),

γΔMs=0= (

γΔMs=0αα 00 γΔMs=0

ββ)

=12((

γΔMs=0αα 00 γΔMs=0

ββ) + (

γΔMs=0ββ 00 γΔMs=0

αα))

+12((

γΔMs=0αα 00 γΔMs=0

ββ) − (

γΔMs=0ββ 00 γΔMs=0

αα))

= γΔMs=0singlet + γΔMs=0

triplet . (30)

Here, the ΔMs = 0 label indicates that the density matrix γΔMs=0

is computed for the spin-conserving excitations. Importantly, thisapplies not only to the EOM-EE ansatz but also to the EOM-SF,EOM-IP, and EOM-EA variants for the transitions between thestates with the same spin projections. Nonzero spin blocks aredenoted as γΔMs=0

αα and γΔMs=0ββ , where the spin label indicates the α

or β spin-orbitals in Eq. (7). It is easy to see that γΔMs=0triplet matrix is

constructed through a†pαaqα − a

†pβaqβ excitations between the bra and

ket states. Although it is a triplet excitation operator, for the con-sistency with other triplet excitation operators, the γΔMs=0

triplet matrix

should be further divided by√

2, as per the definition of T inEq. (19). If the spin projections of bra and ket differ by ±1, the cor-responding excitation operators are spin-flip triplet operators, andno decomposition is needed. However, for the spin-tensor form,they should be multiplied by the appropriate phase factors fromEqs. (18) and (20). We note that for these spin-flipping excita-tions, right-to-left and left-to-right density matrices gain oppositephase factors. Once the density matrix between the states with thesame spin projection is computed, one can easily calculate u fromEq. (29) as

u =1√

2(γΔMs=0

αα − γΔMs=0ββ )/⟨S′M; 10∣SM⟩, (31)

where γ is computed for the transition from the state with spin S′ andspin projection M to the state with spin S and the same spin projec-tion M. Equation (31) gives a recipe for computing u from the Ms = 0transition densities obtained for general wave functions expressedin spin-orbital basis, rather than spinless states29,31 (u can be alsocomputed from the Ms = ±1 transition densities). The contraction ofu with the spin–orbit integrals gives the spin–orbit reduced matrixelements (see the note66),

⟨IS∣∣HL− ∣∣I′S′⟩ = −

12∑p,q

h1eSOL− ;pqu

1,⋅pq, (32)

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-5

Published under license by AIP Publishing

Page 7: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

⟨IS∣∣HL0 ∣∣I′S′⟩ =

√2

2 ∑pqh1eSOLz ;pqu

1,⋅pq, (33)

⟨IS∣∣HL+ ∣∣I′S′⟩ = +

12∑pq

h1eSOL+ ;pqu

1,⋅pq, (34)

and the full spin–orbit matrix is a product of Clebsh–Gordan’s coef-ficients to these reduced matrix elements. The prefactors in theexpressions above arise because the spin part is taken in the spin-tensor form, whereas the orbital momentum is taken in the L+/L−, Lzform, which is not a spin-tensor form. The overall workflow of thealgorithm is shown in Fig. 1.

D. SOMF approximation for closed- and open-shellreferences

SOMF approximation2,75 considerably reduces the cost of thecalculation of the two-electron part of SOC while introducing neg-ligible errors. Originally introduced in 1996,75 SOMF (and somemore aggressive approximations) have been used with great successfor computing SOCs within wave function approaches and den-sity functional theory.2,23–26,28,76–79 Briefly, SOMF entails28 consid-ering only the contributions from the separable part of the two-particle density matrix. In CC/EOM-CC, the separable part of thetwo-particle density matrix, Γ, has the following form:80

Γpqrs = P+(pr, qs)P−(p, q)ρrpγqs = ρrpγqs − ρrqγps + ρsqγpr − ρspγqr ,(35)

ρrp, =⎧⎪⎪⎨⎪⎪⎩

δrp r, p ∈ occupied in Φ0,

0, otherwise,(36)

where the operators P+(pr, qs) and P−(p, q) generate sym-metrized/antisymmetrized expressions, respectively. Here, ρ is thedensity of the reference determinant, which is diagonal in the case of

FIG. 1. The workflow of the present algorithm (shown by green arrows), Fedorov’sscheme41 (red arrows), and the approach implemented in Molcas29 (blue arrows).Here, the WE label denotes the application of Wigner–Eckart’s theorem. In ourscheme, we start from one transition density matrix, extract the triplet part, con-struct the spinless density, and then form three reduced matrix elements. In theMolcas scheme, the spinless transition density matrix is built directly from spinmanifolds ||IS⟩ and ||I′S′⟩, also known as “tensor” states. In Fedorov’s scheme,the reduced matrix elements are computed from three matrix elements.

canonical Hartree–Fock molecular orbitals. When contracted with atwo-electron operator A = 1

2 ∑pqrs Apqrsp†q†sr, the separable part ofthe two-particle density matrix yields

α2e=

12∑pq

γpq[∑rsρrs(Aprqs + Arpsq − Aprsq − Arpqs)] =∑

pqγpqASOMF

pq ,

(37)

where ASOMFpq is defined by the above equation. Further simplifica-

tions are possible depending on the permutational symmetry of ten-sor A. When applied to the two-electron part of the BP Hamiltonian,hSOO(1, 2), the mean-field matrices are

hSOMF,2epq =

12∑rs

ρrs[Jprqs − Jprsq − Jrpqs], (38)

where J denotes the two-electron spin–orbit integrals,

Jprqs = ⟨ϕp(1)ϕr(2)∣hsoo(1, 2)(s(1) + 2s(2))∣ϕq(1)ϕs(2)⟩. (39)

A detailed derivation of SOMF, the expressions for the spin–orbit integrals, and the spin-integrated expressions (expressionsfor hSOMF,2e

pq in spatial molecular orbital set) can be found inRef. 28.

An efficient evaluation of hSOMF,2epq defined by Eq. (38) is car-

ried out in atomic orbitals. Since it contains Coulomb-like and twoexchangelike terms, the implementation is using the same tech-niques as in the Fock matrix build within the Hartree–Fock proce-dure. Computation of one- and two-electron spin–orbit integrals isperformed with the King–Furlani algorithm.18

As often is the case with mean-field treatments, constructionof mean-field effective operators for open-shell references mightlead to artifacts. Below, we discuss intrinsic issues of SOMF in thecase of open-shell references. Note that these problems arise onlywhen the reference, not the target state, has open-shell character. Forexample, EOM-IP/EA calculations of doublet states using closed-shell references are not affected by this issue; however, EOM-SFcalculations using high-spin references are. The analysis is basedon considering the second-quantized form of the mean-field spin–orbit operator in a spin-restricted spin-orbital basis (meaning thatthe spatial parts of spin-orbitals ϕpα and ϕpβ are the same, e.g.,AOs, restricted HF orbitals for a closed-shell reference, etc.). Thetwo-electron mean-field operator can be written as

HSOMF,2e=

12(∑

pqhSOMF,2eL+ ,pq a†

pβaqα

+∑pq(hSOMF,2e

z,pαqα a†pαaqα + hSOMF,2e

z,pβqβ a†pβaqβ)

+∑pq

hSOMF,2eL− ,pq a†

pαaqβ), (40)

wherehSOMF,2eL+

= hSOMF,2ex,βα + ihSOMF,2e

y,βα , (41)

hSOMF,2eL− = hSOMF,2e

x,αβ − ihSOMF,2ey,αβ . (42)

The indices p and q run over spin-restricted orbitals (e.g., AOs), andthe spin labels for hSOMF indicate the parts that should be contracted

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-6

Published under license by AIP Publishing

Page 8: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

with the corresponding parts of the transition density matrix to yieldSOCs (see the note66).

In the case of closed-shell references, the SOMF approxima-tion of the two-electron part of SOC can be treated in the sameway as the one-electron part. To illustrate the problem of open-shellreferences, let us first discuss the contributions from the hSOMF,2e

zterms. In the case of a closed-shell reference, α and β parts of theHF density (or, in general, the density of the reference determi-nant) are the same, which makes hSOMF,2e

z,pαqα equal to −hSOMF,2ez,pβqβ . In this

case, the SOMF expression reduces to the form of one-electron part[Eq. (21)]. The minus sign comes from the dependence of spin–orbit operators on spin, as per Eq. (8), and from the fact that αand β electrons have the opposite signs of Sz . However, in gen-eral, hSOMF,2e

z,pαqα ≠ −hSOMF,2ez,pβqβ for an open-shell reference determinant in

which the number of α and β electrons is different. In this case, α andβ parts of the HF density are not the same; therefore, the mean-fieldFock-like matrices from Eq. (38) also do not have a proper sym-metry of the different spin blocks. The hSOMF,2e

z terms acquire addi-tional unphysical contributions from the singlet part of the transi-tion density matrix, which leads to artifacts illustrated numerically inSec. III D.

As shown below, the singlet contribution violates the symmetryof the spin–orbit operator. One possible solution, which we adopt inthis work, is to antisymmetrize the hSOMF,2e

z integrals with respectto the αα and ββ parts and take a trace triplet transition densitymatrix for SOCs. This operation eliminates the unphysical contri-bution from the singlet transition densities. By doing so, one cantake the hSOMF,2e

z integrals out of parentheses, making the form ofthe SOMF expression [Eq. (40)] the same as for the one-electron part[Eq. (21)].

The contribution of the two remaining terms to SOC in thecase of the open-shell references requires the formation of hSOMF

L± notfrom the full x and y matrices, as in Eqs. (22) and (23), but fromtheir parts that will be contracted with transition density matricesof a†

pβaqα and a†pαaqβ types, respectively (see Appendix B in Ref. 28).

This leads to the partitioning of the terms as in Eqs. (41) and (42).Different α and β parts of the HF density deteriorate the symmetryof the resulting matrix elements as well. The impact of this violationand a possible fix are discussed in Sec. III D.

E. Averaging scheme for interstate matrix elementswithin a non-Hermitian framework

To obtain a Hermitian SOC matrix, the SOC couplings shouldbe averaged in some way. The averaging procedure should satisfy thefollowing requirements:

1. The averaging procedure should be applicable to complex-valued quantities.

2. The averaging procedure should produce meaningful phases.3. The averaged matrix should transform under the rotations

in the same manner as the unaveraged matrices, making theSOCC and splittings invariant with respect to the choice of thecoordinate system.A common strategy for computing interstate properties within

EOM-CC is to take a geometric average of individual matrix ele-ments.57 Numerical examples illustrate that, as expected, the phasesof A → B and B → A matrix elements are not exactly conjugated.

This leads to an ambiguity in geometric averaging: one can eitheraverage absolute values of the matrix elements and assign phasesin some way, or average the Cartesian components of the matrix.We observe that in the absence of point group symmetry geomet-ric averaging of the absolute values violates the rotational invarianceof SOCC. By contrast, the arithmetic average satisfies the require-ments above while producing the same value regardless of whetherthe spherical or Cartesian components were averaged.

We can consider the transformation properties of the arith-metic average,

H′ar.av,A→B = H′A→B + (H′B→A)

†= D†

SHA→BDS′ + (D†S′HB→ADS)

= D†SHA→BDS′ + (D†

SH†B→ADS′) = D†

SHar.av,A→BDS′ .(43)

Here, D are Wigner’s D-matrices, written in the basis of states withS or S′ spins, and H denotes spin–orbit Hamiltonian matrix, com-puted between electronic states with spin projections −S, −S + 1, . . .,S − 1, S and −S′, −S′ + 1, . . ., S′ − 1, S′. Although weighted arithmeticaverage is more theoretically justified,81 a simple arithmetic averageis often numerically close to the weighted average.82

III. RESULTS AND DISCUSSIONA. Computational details

To illustrate the methodology and to provide reference datafor future developments, we considered several groups of moleculeswith various electronic structure patterns:

1. A set of isoelectronic diradicals, which were investigated inprevious studies,28,83 CH2, NH+

2 , SiH2, and PH+2 .

2. A set of doublet radicals, which are of interest in laser coolingexperiments: AsN+, GeO+, CaF, and CaOCH3.

3. A set of transition-metal ions: Fe3+, Fe2+, and Mn+.4. A single-center iron-containing SMM derived from the struc-

ture from Ref. 47.Figure 2 illustrates essential features of the electronic structure

patterns of the molecules from the four sets and explains whichvariants of EOM-CC were used to access the target states.

The diradical set was treated with the EOM-EE-CCSD andEOM-SF-CCSD methods. The latter provides a balanced treatmentof the leading configurations in the target-states wave functions—the determinants shown in Fig. 2(a) appear at the same level ofexcitations from the high-spin reference determinant.

In set 2, the target states have either one electron or one holein the two degenerate orbitals (π or E). To achieve a balancedtreatment of the corresponding configurations, we used EOM-IP-CCSD and EOM-EA-CCSD. EOM-IP-CCSD removes an electronfrom the fully occupied degenerate orbitals of the closed-shell neu-tral reference to describe diatomic cations [Fig. 2(b)], whereasEOM-EA-CCSD adds an electron to empty degenerate orbitalsof the cationic reference to describe neutral calcium derivatives[Fig. 2(c)].

The electronic degeneracies are even more extensive for tran-sition metal ions and the SMM. To treat these systems, we usedthe EOM-SF-CCSD and EOM-EA-CCSD methods with high-spinhextet references, which afford a balanced treatment of the quartetd5 and quintet d6 determinants [Figs. 2(d) and 2(e)].

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-7

Published under license by AIP Publishing

Page 9: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

FIG. 2. Electronic configurations for reference and target states of the consideredsystems: (a) Methylenelike diradicals: CH2, SiH2, NH+

2 , and PH+2 ; (b) diatomic

cations: AsN+ and GeO+; (c) CaF and CaOCH3; (d) and (e) Hextet reference(left), one of the quartet d5 configurations of Fe3+ (top), and one of the quintetd6 configurations of Fe2+, Mn+, and (tpa)Fe− (bottom).

We used Dunning’s cc-pVDZ and cc-pVTZ basis sets.84–87 Coreelectrons were frozen in all calculations. Unrestricted HF refer-ences were used in the calculations using open-shell references. Fordiradicals, we used the same geometries as in Refs. 28 and 83. Thecalculations for set 2 were carried out at their ground-state dou-blet geometries. The structures of AsN+ and GeO+ were optimizedwith EOM-IP-CCSD/cc-pVDZ from the neutral reference; CaF andCaOCH3 were optimized with EOM-EA-CCSD/cc-pVDZ from thecationic reference.

The structure of a model system representing SMM was derivedfrom the original one47 by replacing the mesitylene groups in thechelating ligand, tpaMes, by hydrogen. We refer to the resulting lig-and as tris(pyrrolylmethyl)amine (tpa). The ground state of the neu-tral (tpa)Fe has 5 unpaired electrons, as in Fe3+ ion. The geometry ofthe neutral (tpa)Fe complex was optimized with ωB97X-D/cc-pVDZfor the hextet state. We used this geometry for the calculations of theanion as well.

All relevant Cartesian geometries are given in supplementarymaterial.

We used tight convergence criteria for the SOC calculationsfor all examples except Fe(II) complex: SCF (10−12), EOM (10−10),CC energy (10−10), and T and Λ-amplitudes (10−9). For (tpa)Fe−,we used the default Q-Chem’s convergence criteria: SCF (10−8) andEOM (10−5).

B. Numerical examplesTables I–III show the SOCCs computed with the selected

EOM-CC methods.Table I shows the results for diradicals.88 As discussed in previ-

ous work,28 the SOCCs computed with EOM-EE and EOM-SF agreewell in the systems with small diradical character (SiH2 and PH2),but the differences increase when diradical character increases. As

TABLE I. Spin–orbit coupling constants (cm–1) between the 3B1 and 1A1 states inselected diradicals.

EOM-SF-CCSD/cc-pVTZ EOM-EE-CCSD/cc-pVTZ

System 1e SOMF 1e SOMF

CH2 21.79 10.86 26.55 13.22SiH2 73.79 56.74 77.86 59.87NH+

2 31.65 18.26 61.20 35.26PH+

2 152.26 119.97 160.92 126.78

one can see, the discrepancy is dominated by the one-electronpart; therefore, it is not related to the artifacts due to open-shellreferences.

The results in Table II show anticipated trends: the magnitudeof SOCC increases for heavy elements. Moreover, for these systems,the contribution of the two-electron part is smaller than for typicalorganic molecules. The magnitude of SOCCs between the degener-ate pairs of states (2Π or 2E) is slightly larger than between π- andσ manifolds, which probably can be rationalized by considering theshapes of the respective MOs and El-Sayed’s rule.89

Table III shows atomic SOCCs. For atoms, Eq. (1) no longergives a rotationally invariant constant. Therefore, here we summedthe matrix elements not only by spin projections but also by theprojections of orbital angular momentum. This also leads to largeatomic SOCCs. These values can be compared with experimentaldata, as explained below. To compare these values with experiment,one should divide the SOCC by

√L(L + 1)

√S(S + 1). The result-

ing quantity can be compared with the Russell–Sauders LS couplingconstant (ζ) extracted from splittings through the Landé intervalrule.90 For example, SOMF gives an estimate for 5Dg levels of Mn+ ofζ = 68.29 cm−1, while the values extracted from experiment arewithin 58.66–66.99 cm−1 range.91 The averaged experimental valueof ζ for Fe2+ is 100 cm−1,92 while the SOMF estimate from Table IIIis 108.54 cm−1.

TABLE II. Doublet radicals relevant for laser cooling experiments. SOCCs (cm–1)between the lowest degenerate and nondegenerate doublet states are com-puted with EOM-EA-CCSD/cc-pVTZ from closed-shell cationic references (CaF,CaOCH3) and with EOM-IP-CCSD/cc-pVTZ from neutral closed-shell references(AsN+ and GeO+).

2Σ+/2Π or 2A1/2E 2Π/2Π or 2E/2E

System 1e SOMF 1e SOMF

CaFa 28.86 24.95 49.60 39.09CaOCH3

b 26.57 22.89 43.24 34.42

AsN+a 373.49 349.31 436.81 375.19GeO+a 97.56 110.07 228.81 166.90

aTransitions between 2Π components and 2Σ+ were considered.bTransitions between 2E components and 2A1 were considered.

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-8

Published under license by AIP Publishing

Page 10: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

TABLE III. Atomic SOCCs (cm–1) computed for several ions. EOM-SF-CCSD/cc-pVTZ from the hextet 6S reference was used for Fe3+ states; EOM-EA-CCSD/cc-pVTZ with β electron attachment was used for isoelectronic Fe2+ and Mn+ statesfrom the 6S Fe3+ and Mn2+ references.

System 1e SOMF

Fe3+, 4Fg 135.90 80.95Fe3+, 4Pg 32.64 19.32Fe2+, 5Dg 2159.21 1189.01Mn+, 5Dg 1413.78 748.14

C. ZFS in Fe(II) SMMIron complexes, which feature rich electronic properties related

to the malleability of their spin states, have been extensively stud-ied.93–95 Here, we consider a single-center Fe(II) SMM, which wasshown to have a large spin-reversal barrier (magnetic anisotropy).47

The structure of the model system representing this SMM is shownin Fig. 3. The SMM has d6 electronic configuration, correspond-ing to the overall negative charge.47 As explained below, this elec-tronic configuration gives rise to electronic degeneracy and Jahn–Teller distortions. EOM-CC methods are capable of tackling Jahn–Teller and pseudo-Jahn–Teller effects very well;59,71,96–99 however,this particular system is different and its description is affected bythe non-Hermitian nature of EOM-CC theory.100

The high-spin hextet neutral (tpa)Fe state has electronic con-figuration d5 and, therefore, can be well described by a single Slaterdeterminant, similarly to Fe3+ and Mn2+. This state is not degen-erate, and it has a geometry with C3 point group of symmetry, asshown in Fig. 3. Although this group is Abelian, it has one- andtwo-dimensional real irreducible representations (irreps), which cangive rise to the Jahn–Teller effect for certain electronic configura-tions of the two-dimensional irrep.101 For example, the quintet anion(tpa)Fe− has doubly degenerate states in C3 structures. Unlike C3vgroup, C3 group does not have planes of symmetry, which wouldsplit a two-dimensional irrep into a symmetric and antisymmet-ric irrep with respect to the plane. Therefore, the two Jahn–Tellerstates would not fall into two different irreps as it happens in themajority of symmetry-imposed degenerate states.59,71,96–99 The onlyAbelian subgroup of C3 is C1; therefore, the two degenerate statesbelong to the same irrep, giving rise to general conical intersec-tion problem. As documented numerically102 and theoretically,100

the description of true conical intersections is problematic in EOM-CC due to the non-Hermitian nature of the theory. In our case, thisleads to issues with finding left vectors at the symmetric geometry.To circumvent this problem, we carried out calculations at slightlyasymmetric geometry, coming from DFT optimization (given in thesupplementary material). Although this geometry is asymmetriconly within the magnitude of a symmetry threshold, it leads to asmall artificial energy splitting (∼0.01 eV in the EOM-EA calcula-tion from a neutral reference) of the states that should be degenerate.This artifact is small enough to be neglected in typical photochemicalapplications, but it affects the magnitude of spin–orbit splitting. Tomitigate this issue, we average the energies of the states that shouldbe degenerate by symmetry.

FIG. 3. (a) Structure of (tpa)Fe (C15N4H15Fe) in the hextet neutral state. Iron isshown in red, nitrogen in blue, carbon in gray, and hydrogen in white. (b) Thelowest spin-split levels of quintet anionic (tpa)Fe−, showing a spin-reversal barrierU (the energy gap between the lowest and the highest spin-split states within theground state multiplet).

Table IV clearly shows that the major contribution to ZFS of theground state comes from degenerate state by point group symmetry.ZFS splits not only the spin components, it also splits the degen-erate irreps (the symmetry of the system is no longer described bypoint group symmetry; it is now described by double group sym-metry), and the lowest of them looks very similar to the structureshown in Fig. 3. We followed the state-interaction procedure in

TABLE IV. Spin-reversal barrier (spin-splitting gap in the multiplet) computed withSOMF EOM-EA-MP2/cc-pVDZ from the hextet reference. Spin–orbit splittings werecomputed in the state-interaction approach with the indicated number of electronicstates. The experimental estimate for the barrier is 158 cm–1.

No. electronic multiplets Spin-reversal barrier (cm−1)

2 1733 1585 157

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-9

Published under license by AIP Publishing

Page 11: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

computation of the ZFS splittings: the arithmetically averaged SOCblocks were plugged in the matrix Hamiltonian, and the unper-turbed state energies were on the diagonal. Then, the entire matrixwas diagonalized to yield energy-split multiplet states. To study con-vergence with respect to the number of states included in the calcula-tion, we computed the splittings between 2, 3, and 5 multiplets (10,15, and 25 electronic states with different spin-projections, respec-tively). This sequence was chosen to include the degenerate irrepsof the point group. The values in Table IV are very close to theexperimentally derived result for (tpaMes)Fe− of 158 cm−1.

The computed spin-reversal barriers U are related to the axialmagnetic anisotropy D through the following relation:

U = S2∣D∣. (44)

To estimate a transverse magnetic anisotropy, one should considerJahn–Teller distortions, allowing mixing of ±Ms states, which leadsto the nonzero probabilities of tunneling under the barrier.

D. Impact of violation of L +/L − symmetry and tripletform by canonical SOMF

As explained in Sec. II D, the direct application of SOMF inthe case of open-shell references may violate L+/L−-symmetry andthe triplet structure of the spin–orbit operator. L+/L−-symmetrybecomes evident from the following observation. Equations (8)–(10)from Ref. 41 give

⟨IS∣∣HL+ ∣∣I′S + 1⟩ = ⟨IS∣∣HL− ∣∣I

′S + 1⟩∗. (45)

Because the requirement of having a Hermitian method does notappear in the proof, the statement is applicable to EOM-CC as well.Similarly, Eqs. (11)–(13) from Ref. 41 lead to the condition for thesame multiplicity,

⟨IS∣∣HL+ ∣∣I′S⟩ = ⟨IS∣∣HL− ∣∣I

′S⟩∗. (46)

Because the α part of the Hartree–Fock density ρ is not equal tothe β part for open-shells, the symmetry between hSOMF,2e

± is bro-ken. Table V shows numerical consequences of such violation. For-tunately, regardless of the size of the system or of the number ofunpaired electrons in the reference, the extent of such violation doesnot exceed 0.05 cm−1 for the reduced matrix elements. To eliminateits influence on SOCC and splittings, we used arithmetic averagingfor ⟨S∣∣HL− ∣∣S

′⟩ and ⟨S∣∣HL+ ∣∣S

′⟩, similar to Eq. (43).

TABLE V. Extent of violation of L+/L−-symmetry of SOMF with open-shell references,cm–1. Only A(lowest energy)→ B(higher energy) transition is shown.a

Element/system CH2 NH+2 SiH2 PH+

2

⟨S∣∣HL− ∣∣S′⟩ −13.241i 22.269i 40.489i −146.113i

⟨S∣∣HL+ ∣∣S′⟩ 13.215i −22.264i −40.492i 146.111i

(tpa)Fe–, 1→ 2 (tpa)Fe–, 1→ 3

⟨S∣∣HL− ∣∣S′⟩ 0.293 − 0.096i −32.067 + 228.232i

⟨S∣∣HL+ ∣∣S′⟩ 0.290 + 0.095i −32.073 − 228.276i

aSame computational setup as in Tables I and IV.

TABLE VI. Unphysical singlet reduced matrix elements for the considered systems,cm–1. The hydrides were oriented in the way that ⟨S∣∣HLz ∣∣S

′⟩ elements are notzero.

Element/system CH2 NH+2 SiH2 PH+

2

⟨S∣∣HsingletLz ∣∣S′⟩ 0.004i −0.001i −0.002i 0.001i

(tpa)Fe–, 1→ 2 (tpa)Fe–, 1→ 3

⟨S∣∣HsingletLz ∣∣S′⟩ 13.083i 0.004i

Table VI shows the impact of the singlet component of thetransition density matrix. The singlet part is small for transitionsbetween states with different multiplicities. It is not zero because ofsmall spin contamination of the electronic states. The magnitude ofsinglet part can be significant for the transitions between states ofthe same multiplicities. Using the triplet component of the transi-tion density matrix for calculation of matrix elements circumventsthis issue.

To conclude, the arithmetic averaging between ⟨S∣∣HL− ∣∣S′⟩

and ⟨S∣∣HL+ ∣∣S′⟩∗ restores the correct symmetry of these elements.

Usage of the triplet transition density does not lead to contam-ination of ⟨S∣∣HLz ∣∣S

′⟩ matrix element with the unphysical singlet

contribution.

IV. CONCLUSIONWe presented a theory for calculating matrix elements of the

BP SOC operator based on the application of Wigner–Eckart’s the-orem to reduced one-particle density matrices. The key equationsare given in the second-quantized spin-orbital form. The algorithmis ansatz-agnostic and can be used with any electronic structuremethod for which state and transition density matrices are available.The current implementation is based on the EOM-CCSD suite ofmethods. The approach allows one to compute the SOC matrix forthe entire multiplet from just one transition density matrix, whichsolves the problem of accessibility of states of different spin pro-jections. It also addresses, by constriction, the phase problem dueto Born–Oppenheimer’s separation of the nuclear and electronicdegrees of freedom. We also highlighted an important aspect of thecalculation of SOCCs within a non-Hermitian theory and proposedusing arithmetic averaging of the EOM-CC matrix elements, whichleads to rotationally invariant SOCCs.

The current implementation treats the two-electron part of theBP Hamiltonian via the SOMF approximation. We discuss specialaspects of the application of SOMF to open-shell references and pro-pose practical solutions. In particular, L+ and L− reduced matrixelements are slightly different. This issue can be solved by averaging.Contribution of the (unphysical) singlet part of the transition den-sity can be important. This issue does not occur if only the tripletcomponent is used.

SUPPLEMENTARY MATERIAL

See supplementary material for the example of SOC matrix andrelevant Cartesian coordinates.

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-10

Published under license by AIP Publishing

Page 12: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

ACKNOWLEDGMENTSThis work was supported by the Department of Energy through

Grant No. DE-SC0018910 (A.I.K.). A.I.K. is also a grateful recipientof the Simons Fellowship in Theoretical Physics and a visiting pro-fessorship of the MAINZ graduate school of excellence. We thankDr. Maxim Ivanov for his guidance in calculations of moleculesrelevant in laser cooling experiments.

The authors declare the following competing financial inter-est(s): A.I.K. is a member of the Board of Directors and a part-ownerof Q-Chem, Inc.

APPENDIX: EVALUATION OF CLEBSH–GORDANCOEFFICIENTS

The Racah recursions74 formally give a definite set of Clebsh–Gordan’s coefficients, but their implementation requires a certaineffort to avoid division by zero and possible phase mismatch.Instead, we follow the Complete Set of Commuting Observables(CSCO) approach103 in which matrices of commuting operatorsare constructed and diagonalized. The following algorithm pro-duces Clebsh–Gordan’s coefficients in the Condon–Shortley nota-tion. Here, J denotes angular momentum operators, indices 1 and 2refer to the particles 1 and 2, the absence of an index refers to theresult of addition of angular momentum, j denotes angular momen-tum quantum number, and m denotes the quantum number ofthe projection of angular momentum. The algorithm proceeds asfollows:

1. Take j1, j2, j from the input.

2. Make ⟨ j1m1| J1,±| j1m′1⟩, ⟨ j2m2| J2,±| j2m′2⟩.

3. Make ⟨ j1m1; j2m2| J2| j1m′1; j2m′2⟩ in the tensor space of| j1m′1; j2m′2⟩.

4. Diagonalize ⟨ j1m1; j2m2| J2| j1m′1; j2m′2⟩ and save only eigen-vectors for j( j + 1) eigenvalue.

5. From these eigenvectors, build a projector P onto theirsubspace.

6. Diagonalize B = P( Jz + C)P, where the shift constant C islarge enough to move eigenvalues above zero.

7. Synchronize the phases, multiply the eigenvector vector withthe largest eigenvalue of B by sign of ⟨ j1 j1; j2( j − j2)| jj⟩.

8. Fix the phases of other eigenvectors recursively: for m = j,j − 1, . . ., − j + 1 multiply the |jm − 1⟩ vector by sign of⟨jm − 1| J−|jm⟩.

9. Return ⟨ j1m1; j2m2|jm⟩ coefficients.

This algorithm produces a cube of coefficients with the m1, m2,and m indices, which are needed for calculation of the SOC matrix.

REFERENCES1D. R. Yarkony, Int. Rev. Phys. Chem. 11, 195 (1992).2C. M. Marian, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2, 187 (2012).3Magnetic Properties of Organic Materials, edited by P. M. Lahti (Marcel Dekker,Inc., 1999).

4M. R. Pederson and T. Baruah, “Molecular magnets: Phenomenology and the-ory,” in Handbook of Magnetism and Magnetic Materials, edited by S. Parkin(J. Wiley and Sons, London, 2007), pp. 1–22.5J. P. Malrieu, R. Caballol, C. J. Calzado, C. de Graaf, and N. Guihéry, Chem. Rev.114, 429 (2013).6M. N. Leuenberger and D. Loss, Nature 410, 789 (2001).7C. Schlegel, J. van Slageren, M. Manoli, E. K. Brechin, and M. Dressel, Phys. Rev.Lett. 101, 147203 (2008).8M. J. Martínez-Pérez, S. Cardona-Serra, C. Schlegel, F. Moro, P. J. Alonso,H. Prima-García, J. M. Clemente-Juan, M. Evangelisti, A. Gaita-Arino, J. Sesé,J. van Slageren, E. Coronado, and F. Luis, Phys. Rev. Lett. 108, 247213(2012).9M. D. Jenkins, D. Zueco, O. Roubeau, G. Aromi, J. Majere, and F. Luis, DaltonTrans. 45, 16682 (2016).10M. R. Pederson and S. N. Khanna, Phys. Rev. B 60, 9566 (1999).11O. Vahtras, O. Loboda, B. Minaev, H. Årgen, and K. Ruud, Chem. Phys. 279,133 (2002).12D. Ganyushin and F. Neese, J. Chem. Phys. 125, 024103 (2006).13M. Atanasov, D. Aravena, E. Suturina, E. Bill, D. Maganas, and F. Neese, Coord.Chem. Rev. 289-290, 177 (2015).14H. A. Bethe and E. E. Salpeter, Quantum Mechanics of One and Two ElectronAtoms (Plenum, New York, 1977).15Relativistic Electronic Structure Theory, edited by P. Schwerdtfeger (Elsevier,Amsterdam, 2002).16D. G. Fedorov, S. Koseki, M. W. Schmidt, and M. S. Gordon, Int. Rev. Phys.Chem. 22, 551 (2003).17T. R. Furlani and H. F. King, J. Chem. Phys. 82, 5577 (1985).18H. F. King and T. R. Furlani, J. Comput. Chem. 9, 771 (1988).19O. Vahtras, H. Ågren, P. Jørgensen, H. J. A. Jensen, T. Helgaker, and J. Olsen,J. Chem. Phys. 96, 2118 (1992).20D. G. Fedorov and M. S. Gordon, J. Chem. Phys. 112, 5611 (2000).21D. G. Fedorov and J. P. Finley, Phys. Rev. A 64, 042502 (2001).22H. Ågren, O. Vahtras, and B. Minaev, “Response theory and calculations of spin-orbit coupling phenomena in molecules,” in Advances in Quantum Chemistry(Academic Press, 1996), Vol. 27, pp. 71–162.23A. Berning, M. Schweizer, H.-J. Werner, P. Knowles, and P. Palmieri, Mol. Phys.98, 1823 (2000).24F. Neese, J. Chem. Phys. 122, 034107 (2005).25O. Christiansen, J. Gauss, and B. Schimmelpfennig, Phys. Chem. Chem. Phys.2, 965 (2000).26K. Klein and J. Gauss, J. Chem. Phys. 129, 194106 (2008).27L. A. Mück, “Highly accurate quantum chemistry: Spin-orbit splittings via mul-tireference coupled-cluster methods and applications in heavy-atom main-groupchemistry,” Ph.D. thesis, Johannes-Gutenberg Universität Mainz, 2013.28E. Epifanovsky, K. Klein, S. Stopkowicz, J. Gauss, and A. I. Krylov, J. Chem.Phys. 143, 064102 (2015).29P. Å. Malmqvist, B. O. Roos, and B. Schimmelphennig, Chem. Phys. Lett. 357,230 (2002).30B. O. Roos and P.-Å. Malmqvist, Phys. Chem. Chem. Phys. 6, 2919 (2004).31M. Roemelt, J. Chem. Phys. 143, 044112 (2015).32L. A. Mück and J. Gauss, J. Chem. Phys. 136, 111103 (2012).33F. Neese, T. Petrenko, D. Ganyushin, and G. Olbrich, Coord. Chem. Rev. 251,288 (2007).34S. G. Chiodo and N. Russo, Chem. Phys. Lett. 490, 90 (2010).35Z. Li, B. Suo, Y. Zhang, Y. Xiao, and W. Liu, Mol. Phys. 111, 3741(2013).36M. Roemelt, D. Maganas, S. DeBeer, and F. Neese, J. Chem. Phys. 138, 204101(2013).37Within Born–Oppenheimer separation of nuclear and electronic degrees offreedom, the phase of the electronic wave function is not defined: it changes atrandom at different nuclear geometries or even from calculation to calculation atthe same geometry. To ensure the consistency of the signs of the matrix elements,the phase should be fixed by an additional condition.38C. Eckart, Rev. Mod. Phys. 2, 305 (1930).

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-11

Published under license by AIP Publishing

Page 13: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

39E. Wigner, Z. Phys. 43, 624 (1927).40R. N. Zare, Angular Momentum: Understanding Spatial Aspects in Chemistryand Physics (J. Wiley & Sons, 1988).41D. G. Fedorov, “Theoretical study of spin-orbit coupling in molecules,” Ph.D.thesis, Iowa State University, 1999.42S. Schmitt, P. Jost, and C. van Wüllen, J. Chem. Phys. 134, 194113 (2011).43B. Cahier, R. Maurice, H. Bolvin, T. Mallah, and N. Guihéry, Magnetochemistry2, 31 (2016).44D. Ganyushin and F. Neese, J. Chem. Phys. 128, 114117 (2008).45R. McWeeny, J. Chem. Phys. 42, 1717 (1965).46R. McWeeny, Spins in Chemistry (Academic Press, New York, 1970).47D. E. Freedman, W. H. Harman, T. D. Harris, G. J. Long, C. J. Chang, andJ. R. Long, J. Am. Chem. Soc. 132, 1224 (2010).48Y. Shao, Z. Gan, E. Epifanovsky, A. T. B. Gilbert, M. Wormit, J. Kussmann,A. W. Lange, A. Behn, J. Deng, X. Feng, D. Ghosh, M. Goldey, P. R. Horn,L. D. Jacobson, I. Kaliman, R. Z. Khaliullin, T. Kus, A. Landau, J. Liu, E. I. Proynov,Y. M. Rhee, R. M. Richard, M. A. Rohrdanz, R. P. Steele, E. J. Sundstrom,H. L. Woodcock III, P. M. Zimmerman, D. Zuev, B. Albrecht, E. Alguires,B. Austin, G. J. O. Beran, Y. A. Bernard, E. Berquist, K. Brandhorst, K. B. Bravaya,S. T. Brown, D. Casanova, C.-M. Chang, Y. Chen, S. H. Chien, K. D. Closser,D. L. Crittenden, M. Diedenhofen, R. A. DiStasio, Jr., H. Do, A. D. Dutoi, R. G.Edgar, S. Fatehi, L. Fusti-Molnar, A. Ghysels, A. Golubeva-Zadorozhnaya,J. Gomes, M. W. D. Hanson-Heine, P. H. P. Harbach, A. W. Hauser, E. G. Hohen-stein, Z. C. Holden, T.-C. Jagau, H. Ji, B. Kaduk, K. Khistyaev, J. Kim, J. Kim, R. A.King, P. Klunzinger, D. Kosenkov, T. Kowalczyk, C. M. Krauter, K. U. Laog,A. Laurent, K. V. Lawler, S. V. Levchenko, C. Y. Lin, F. Liu, E. Livshits, R. C.Lochan, A. Luenser, P. Manohar, S. F. Manzer, S.-P. Mao, N. Mardirossian,A. V. Marenich, S. A. Maurer, N. J. Mayhall, C. M. Oana, R. Olivares-Amaya, D. P.O’Neill, J. A. Parkhill, T. M. Perrine, R. Peverati, P. A. Pieniazek, A. Prociuk,D. R. Rehn, E. Rosta, N. J. Russ, N. Sergueev, S. M. Sharada, S. Sharmaa, D. W.Small, A. Sodt, T. Stein, D. Stuck, Y.-C. Su, A. J. W. Thom, T. Tsuchimochi,L. Vogt, O. Vydrov, T. Wang, M. A. Watson, J. Wenzel, A. White, C. F. Williams,V. Vanovschi, S. Yeganeh, S. R. Yost, Z.-Q. You, I. Y. Zhang, X. Zhang, Y. Zhou,B. R. Brooks, G. K. L. Chan, D. M. Chipman, C. J. Cramer, W. A. Goddard III,M. S. Gordon, W. J. Hehre, A. Klamt, H. F. Schaefer III, M. W. Schmidt, C. D.Sherrill, D. G. Truhlar, A. Warshel, X. Xu, A. Aspuru-Guzik, R. Baer, A. T. Bell,N. A. Besley, J.-D. Chai, A. Dreuw, B. D. Dunietz, T. R. Furlani, S. R. Gwaltney,C.-P. Hsu, Y. Jung, J. Kong, D. S. Lambrecht, W. Z. Liang, C. Ochsenfeld,V. A. Rassolov, L. V. Slipchenko, J. E. Subotnik, T. Van Voorhis, J. M. Herbert,A. I. Krylov, P. M. W. Gill, and M. Head-Gordon, Mol. Phys. 113, 184 (2015).49A. I. Krylov and P. M. W. Gill, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 3, 317(2013).50K. Emrich, Nucl. Phys. A 351, 379 (1981).51A. I. Krylov, Annu. Rev. Phys. Chem. 59, 433 (2008).52R. J. Bartlett, Mol. Phys. 108, 2905 (2010).53K. Sneskov and O. Christiansen, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2,566 (2012).54R. J. Bartlett, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2, 126 (2012).55Traditionally, orbitals occupied in Φ0 are denoted by i, j, k, . . . and virtualorbitals are denoted by a, b, c, . . .. Orbitals that can be either occupied or unoccu-pied are denoted by p, q, r, . . .. Atomic orbitals are denoted by Greek letters: μ, ν,λ, . . ..56Excitation/de-excitation operators are defined with respect to the vacuum state,Φ0. An excitation operator can only contain creation operators corresponding tothe virtual orbitals, referred to as particle (p) operators, and annihilation oper-ators corresponding to the occupied orbitals, referred to as hole (h) operators.EOM operators R and CC operators T are excitation operators. At the CCSD level,cluster operators T are of the 1h1p and 2h2p types.57J. F. Stanton and R. J. Bartlett, J. Chem. Phys. 98, 7029 (1993).58J. F. Stanton and J. Gauss, J. Chem. Phys. 101, 8938 (1994).59P. A. Pieniazek, S. E. Bradforth, and A. I. Krylov, J. Chem. Phys. 129, 074104(2008).60M. Nooijen and R. J. Bartlett, J. Chem. Phys. 102, 3629 (1995).61A. I. Krylov, Chem. Phys. Lett. 338, 375 (2001).62A. I. Krylov, Acc. Chem. Res. 39, 83 (2006).

63K. W. Sattelmeyer, H. F. Schaefer, and J. F. Stanton, Chem. Phys. Lett. 378, 42(2003).64M. Musiał, A. Perera, and R. J. Bartlett, J. Chem. Phys. 134, 114108 (2011).65D. Casanova, L. V. Slipchenko, A. I. Krylov, and M. Head-Gordon, J. Chem.Phys. 130, 044103 (2009).66In the case of nonorthogonal orbitals (e.g., atomic orbitals), it is important todistinguish contravariant and covariant indices. Herein, the integrals are assumedto be contravariant and the densities are assumed to be covariant. Since the orbitalspaces have a metric, the overlap matrix S, the relation between the covariantand contravariant density matrices is γcov = S−1γcontraS−1. Since the workingexpressions of EOM-CC calculate the density matrices in the MO basis, covari-ant densities can be obtained without a need for the inverse overlap: γcov(AO)= Cγ(MO)C†. Covariance or contravariance does not change the spin-tensorproperties of the density matrix, such as spin projection M or singlet or triplettype.67For I = I′, γ is simply one-particle state density matrix, and for I ≠ I′, γ is one-particle transition density matrix.68T. Helgaker, S. Coriani, P. Jørgensen, K. Kristensen, J. Olsen, and K. Ruud,Chem. Rev. 112, 543 (2012).69T. Helgaker, P. Jørgensen, and J. Olsen, Molecular Electronic Structure Theory(Wiley & Sons, 2000).70K. D. Nanda and A. I. Krylov, J. Chem. Phys. 142, 064118 (2015).71A. I. Krylov, “The quantum chemistry of open-shell species,” in Reviews in Com-putational Chemistry, edited by A. L. Parrill and K. B. Lipkowitz (J. Wiley & Sons,2017), Vol. 30, Chap. 4, pp. 151–224.72E. Epifanovsky, M. Wormit, T. Kus, A. Landau, D. Zuev, K. Khistyaev,P. U. Manohar, I. Kaliman, A. Dreuw, and A. I. Krylov, J. Comput. Chem. 34,2293 (2013).73P. Pokhilko, R. Shannon, D. Glowacki, H. Wang, and A. I. Krylov, J. Phys.Chem. A 123, 482 (2019).74G. Racah, Phys. Rev. 62, 438 (1942).75B. A. Hess, C. M. Marian, U. Wahlgren, and O. Gropen, Chem. Phys. Lett. 251,365 (1996).76K. Ruud, B. Schimmelpfennig, and H. Ågren, Chem. Phys. Lett. 310, 215(1999).77M. Kaupp, R. Reviakine, O. L. Malkina, A. Arbuznikov, B. Schimmelpfenning,and V. G. Malkin, J. Comput. Chem. 23, 794 (2001).78J. Liu, Y. Shen, A. Asthana, and L. Cheng, J. Chem. Phys. 148, 034106(2018).79J. Liu and L. Cheng, J. Chem. Phys. 148, 144108 (2018).80S. V. Levchenko, T. Wang, and A. I. Krylov, J. Chem. Phys. 122, 224106(2005).81A. Tajti and P. G. Szalay, J. Chem. Phys. 131, 124104 (2009).82S. Faraji, S. Matsika, and A. I. Krylov, J. Chem. Phys. 148, 044103 (2018).83Y. A. Bernard, Y. Shao, and A. I. Krylov, J. Chem. Phys. 136, 204103 (2012).84T. H. Dunning, Jr., J. Chem. Phys. 90, 1007 (1989).85N. B. Balabanov and K. A. Peterson, J. Chem. Phys. 123, 064107 (2005).86J. Koput and K. A. Peterson, J. Phys. Chem. A 106, 9595 (2002).87A. K. Wilson, D. E. Woon, K. A. Peterson, and T. H. Dunning, J. Chem. Phys.110, 7667 (1999).88The SOCCs for CH2, SiH2, PH+

2 , and NH+2 reported in Ref. 28 are off by a factor

of√

2. The reason for this unfortunate mistake becomes clear by considering therespective Cartesian geometries listed in the supplementary material of Ref. 28.89M. A. El-Sayed, Acc. Chem. Res. 1, 8 (1968).90A. Landé, Z. Phys. 15, 189 (1923).91A. Kramida, Yu. Ralchenko, J. Reader, and NIST ASD Team, NIST atomicspectra database, v. 5.6.1, National Institute of Standards and Technology,Gaithersburg, MD, 2018, https://physics.nist.gov/asd.92T. M. Dunn, Trans. Faraday Soc. 57, 1441 (1961).93P. Gütlich, “Spin crossover in iron(II)-complexes,” in Metal Complexes(Springer Berlin Heidelberg, Berlin, Heidelberg, 1981), pp. 83–195.94P. Gütich and H. A. Goodwin, Top. Curr. Chem. 233, 1 (2004).95T. J. Penfold, E. Gindensperger, C. Daniel, and C. M. Marian, Chem. Rev. 118,6975 (2018).

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-12

Published under license by AIP Publishing

Page 14: General framework for calculating spin orbit couplings ...iopenshell.usc.edu/pubs/pdf/jcp-151-034106.pdf · contraction, an increase in electron binding energies, and spin–orbit

The Journalof Chemical Physics ARTICLE scitation.org/journal/jcp

96V. A. Mozhayskiy, D. Babikov, and A. I. Krylov, J. Chem. Phys. 124, 224309(2006).97V. A. Mozhayskiy and A. I. Krylov, Mol. Phys. 107, 929 (2009).98L. V. Slipchenko and A. I. Krylov, J. Chem. Phys. 118, 6874 (2003).99L. V. Slipchenko and A. I. Krylov, J. Chem. Phys. 118, 9614 (2003).

100E. F. Kjønstad and H. Koch, J. Phys. Chem. Lett. 8, 4801 (2017).101H. A. Jahn and E. Teller, Proc. R. Soc. London, Ser. A 161, 220 (1937).102A. Köhn and A. Tajti, J. Chem. Phys. 127, 044105 (2007).103J.-Q. Chen, J. Ping, and F. Wang, Group Representation Theory for Physicists,2nd ed. (World Scientific, 2002).

J. Chem. Phys. 151, 034106 (2019); doi: 10.1063/1.5108762 151, 034106-13

Published under license by AIP Publishing