<<

Quantum System Partitioning at the Single-Particle Level Quantum System Partitioning at the Single-Particle Level

Adrian H. M¨uhlbach and Markus Reihera) ETH Z¨urich, Laboratorium f¨ur Physikalische Chemie, Vladimir-Prelog-Weg 2, CH-8093 Z¨urich,Switzerland

(Dated: 17 October 2018) We discuss the partitioning of a quantum system by subsystem separation through unitary block- diagonalization (SSUB) applied to a Fock operator. For a one-particle Hilbert space, this separation can be formulated in a very general way. Therefore, it can be applied to very different partitionings ranging from those driven by features in the molecular structure (such as a solute surrounded by solvent molecules or an active site in an enzyme) to those that aim at an orbital separation (such as core-valence separation). Our framework embraces recent developments of Manby and Miller as well as older ones of Huzinaga and Cantu. Projector-based embedding is simplified and accelerated by SSUB. Moreover, it directly relates to decoupling approaches for relativistic four-component many-electron theory. For a Fock operator based on the Dirac one-electron Hamiltonian, one would like to separate the so-called positronic (negative-energy) states from the electronic bound and continuum states. The exact two-component (X2C) approach developed for this purpose becomes a special case of the general SSUB framework and may therefore be viewed as a system- environment decoupling approach. Moreover, for SSUB there exists no restriction with respect to the number of subsystems that are generated — in the limit, decoupling of all single-particle states is recovered, which represents exact diagonalization of the problem. The fact that a Fock operator depends on its eigenvectors poses challenges to all system-environment decoupling approaches and is discussed in terms of the SSUB framework. Apart from improved conceptual understanding, these relations bring about technical advances as developments in different fields can immediately cross-fertilize one another. As an important example we discuss the atomic decomposition of the unitary block-diagonalization in X2C-type approaches that can inspire approaches for the efficient partitioning of large total systems based on SSUB.

I. INTRODUCTION including its sophisticated variants such as polarizable embedding theories14–16. Various fragmentation and em- bedding approaches were conceived to enhance computa- The quantum mechanical study of isolated molecular tional efficiency by reducing the number of one-particle systems has been an important endeavor. Examples basis functions or by fragmenting the system, which also range from scrutinizing our understanding of fundamen- make calculations amenable to massive parallelization; tal physical theory (as highlighted, for instance, by the examples can be found in Refs. 17–24. high resolution results available for the dihydrogen bind- ing energy1–3) to analyzing vast amounts of experimental From the more formal point of view of quantum the- (gas-phase) data in great detail (examples can be found ory, nesting a subsystem into an environment of one or in astrochemistry4,5 as well as in atmospheric and com- more subsystems requires the adoption of open-system bustion chemistry6). However, the majority of experi- ,25,26 which in principle can cope with ments in chemistry considers molecules in some specific any such situation. For an open quantum system, many- environment (in solution, on surfaces, in solid bulk, in particle basis states defined on a subsystem may not nec- enzymes and so forth), which poses huge challenges for essarily conserve particle number as they can be com- their theoretical description. bined with states from the environment to produce a to- tal state of, in most practical cases, fixed particle number. Naturally, a plethora of approximations has been de- The total state may then be expanded in terms of a (ten- veloped to cope with situations in which a local phe- sor) product basis where the double sum runs over indices nomenon, i.e., one that can be described by studying only that refer to subsystem (sub)states and to environment a subsystem, is embedded into some environment that

arXiv:1809.04476v2 [physics.chem-ph] 18 Oct 2018 (sub)states. Such a partitioning of a system can be di- more or less strongly interacts with the subsystem. Some rectly exploited to optimize basis states on a subsystem of these embedding approaches were driven by chemical in numerical procedures. The renormal- and physical insights resting on rather ad hoc theoretical ization group algorithm27,28 is an example, where in each bases of which quantum-mechanics molecular-mechanics 7–13 iteration step a total many-particle state may be viewed (QMMM) coupling is the most prominent example as being decomposed into a product basis of substates defined on a system and an environment of orbitals.

A very special decomposition is the Schmidt a)Electronic mail: [email protected] (correspond- decomposition29,30, which restricts the double sum ing author) over product states to a single sum by connecting each Quantum System Partitioning at the Single-Particle Level 2 state on a system to exactly one (specially prepared, correlated orbitals considered for exact diagonalization e.g., contracted) many-particle state of the environment. into the complementary space of all other less corre- It is this decomposition that has prompted Knizia and lated orbitals, as recently exploited by Shiozaki and co- Chan to define an efficient embedding model called workers in what they call the active space decomposition density matrix embedding theory (DMET)31,32. DMET method83–85. exploits the fact that a potentially small number of Whereas all general open-quantum-system methods op- relevant system states couples, by virtue of the Schmidt erate, as they should, on the many-particle state level, decomposition, to only the same number of states in such separations of a system into subsystems can be the environment, no matter how large the latter is. leveraged by a properly prepared one-particle basis from Obviously, the optimization of such environment states which the many-particle states are then constructed (ei- might be considered as complicated as solving the ther in the usual way by tensor products or in a mean- full quantum problem for the total system (i.e., for field sense for KS-DFT). This was recently demonstrated subsystem and environment). To arrive at a practical in the work of Manby, Miller, and co-workers on different DMET approach, Knizia and Chan proposed a mean- variants of embedding52,61,63 and we come back to their field approximation for the environment states.31,32 The embedding approaches later in this work. mean-field approximation to the general DMET has been studied in detail by them,31,32 by Scuseria and In this work, we consider the separation of a quantum co-workers33,34, and by van Voorhis and co-workers35,36. system by a suitable linear combination of one-electron basis states that allows us to provide a separation ac- Mean-field environments had been considered for system- cording to any desired target, which may be defined in environment partitioning before the introduction of terms of an underlying nuclear framework or by exploit- DMET. The motivation for this has always been the ob- ing a separation of one-particle states based on some en- servation that a part of a total system may be subject to ergy criterion (producing, for instance, core-valence sep- strong quantum correlations whereas for the rest a mean- aration within an atomic or molecular structure). Our field approach can be chosen, which is usually taken to approach is designed to be efficient and generally appli- be Kohn–Sham density functional theory (KS-DFT)37,38. cable. It even relates to exact two-component relativistic Within DFT, it is possible to define density-based for- theories. However, the fact that a Fock operator usually mulations of a system-environment embedding.39–54 The depends on the solution of a one-particle equation (as, in strongly correlated part of a molecule, i.e., the system, general, the 4-current or the electron density (matrix) is may also be described by an accurate wave-function- required to represent the interaction) will pose difficul- theory approach55–66 if deemed necessary to allow for ties for all such embedding approaches, but at the same better error control. time, facilitates the proposition of suitable approxima- Mean-field approximations lend themselves to studying tions. Within our general framework, we will show that quantum system partitioning at the single-particle level, approximations developed for exact two-component ap- i.e., at the level of the one-particle equations of motion proaches may cross-fertilize developments in embedding that describe the dynamics of an electron in a mean- theories by Miller and Manby. field potential. Obviously, Hartree–Fock and Kohn– Sham equations are the most popular targets for such a decomposition. In this work, we will present a gen- II. ONE-ELECTRON HILBERT SPACE eral unitary-transformation-based partitioning approach for single-particle equations. We would like to empha- size, however, that these single-particle equations do not We discuss the partitioning of a system at the level of a need to be of the mean-field type. Our unitary decou- one-electron equation and assume that one can construct pling approach will apply to any single-particle equation, a Fock matrix for the total system and eventually diag- which could, for instance, be of a multi-configuration self- onalize it. Clearly, if the system is very large, this will consistent-field type, in which configuration-interaction become a problem. For such cases, it will be necessary to state parameters enter the electron-electron interaction introduce focused methods to construct and diagonalize at the one-particle level. Fock matrices, which are, for instance, known in plane- wave calculations. QMMM may be viewed as a radical We briefly mention that other embedding theories ex- solution that treats part of a system classical so that it ploit different formulations of the quantum mechan- does not at all contribute any basis states for the rep- ical equations of motion. Examples are the self- resentation of the Fock operator, whereas semiempirical energy embedding theory of Zgid that starts from a methods86–98 allow one to approximate fairly large Fock Green’s function formalism67–69, the dynamical mean- matrices. field theory of Georges and Kotliar for the description of impurities70–73, and work that allows one to nest dif- For the sake of brevity, we focus on mean-field equations ferent quantum formalisms into one another74–76. Also and consider the restricted formalism only. An exten- active-orbital space methods77–82 can be viewed as em- sion to an unrestricted formalism is straightforward. We bedding approaches nesting a set of strongly statically first give a concise overview on the general formalism to Quantum System Partitioning at the Single-Particle Level 3 introduce a unified notation and keep our account self- Each component of V depends on the one-electron den- contained. sity matrix P. The elements of the Coulomb matrix J[P] and the K[P] are calculated from the The Fock matrix F is the representation of a Fock oper- two-electron repulsion integrals evaluated in the chosen ator Fˆ within a basis B consisting of N basis functions, B basis as

B = {φi : i ∈ [1; NB]}. (1) N XB 1 The Fock matrix F depends on the density matrix P, Jij[P] = Pklhφi(1)φk(2)| |φj(1)φl(2)i (7) r12 F[P]. It is the sum of the one-electron matrix H, which kl does not depend on the density matrix P, and the two- and electron matrix V = V[P],

NB F[P] = H + V[P], (2) 1 X 1 Kij[P] = − Pklhφi(1)φk(2)| |φl(1)φj(2)i, (8) 2 r12 both of which are hermitian. The (closed-shell) density kl matrix P, respectively. Here, r12 denotes the Euclidean distance T between the two integration coordinates. In Kohn–Sham P = 2 Cocc (Cocc) , (3) density functional theory37,38, the exchange-correlation occ is calculated from the molecular orbitals ψi , which oc- matrix Vxc[P] is calculated from the exchange-correlation cupy the Hartree–Fock (HF) Slater determinant, in the potential vxc[ρ[P]], chosen basis, Vxc,ij[P] = hφi|vxc[ρ[P]]|φji, (9) NB occ X (k) ψ = c φk (4) i iocc with the electron density ρ[P], k=1

(k) NB where the c are elements of a vector ci represent- X iocc occ ρ[P] = P φ φ , (10) ing the i-th occupied orbital in this basis. All vectors ij i j ij ciocc enter Cocc as column vectors. Their determination requires diagonalization of the Fock matrix F. where we assume real orbitals. The one-electron matrix H consists of contributions from the kinetic energy of an electron and the potential energy The electronic energy Eel[P] is the sum of the Coulomb arising from the attractive Coulomb interaction between energy of all nuclei Enuc, the one-electron energy EH [P], the Coulomb and exchange energies EJK [α,P], and the an electron and the Nnuc nuclei of the system. We write exchange-correlation functional Exc[P], its matrix elements Hij in Hartree atomic units (used throughout) as Eel[P] = Enuc + EH [P] + EJK [α,P] + βExc[P]. (11) N 1 Xnuc 1 Hij = − hφi|∆|φji − ZI hφi| |φji, (5) The Coulomb energy of all nuclei is calculated as 2 rI I Nnuc X ZiZj with the Laplace operator in three dimensions ∆, the E = . (12) nuc r nuclear charge number of the I-th nucleus ZI , and the i

V[P] = J[P] + αK[P] + βVxc[P], (6) The energy contribution EJK [α,P] is evaluated from the Coulomb and exchange matrices J and K, respectively, where α mixes in exact (Hartree–Fock) exchange that needs to be corrected for in V (not shown). In Hartree– NB xc 1 X 1 Fock theory, α=1 and β=0 in Eq. (6). In Kohn–Sham E [P] = (J + αK ) P = Tr((J + αK) P). JK 2 ij ij ji 2 density functional theory, β=1 and α controls the amount ij of exact exchange admixture. (14) Quantum System Partitioning at the Single-Particle Level 4

III. THE SELF-CONSISTENT-FIELD PROCEDURE a new density matrix P(n),

T (n) (n)  (n) To obtain the ground-state energy, the electronic energy P = 2 Cocc Cocc , (20) is minimized in a self-consistent manner through the self- consistent field (SCF) procedure. The following descrip- is calculated. tion will only consider a basic formulation that is required As convergence criteria can serve the Frobenius norm of to later discuss all options of self-consistent and approx- the difference between two consecutive density matrices, imate embedding schemes.

(n) (n−1) (n) Quantities that do not depend on the density matrix P δP = P − P , (21) are evaluated before the iterative part of the SCF proce- dure starts. Apart from the one-electron matrix H whose and between total electronic energies calculated from elements were defined in Eq. (5), this is the overlap ma- them, trix S, with elements (n) δ = |E [P(n−1)] − E [P(n)]| , (22) Eel el el Sij = hφi|φji (15) to be below a predefined threshold. If the procedure has Depending on hardware constraints, the two-electron re- not yet converged, a new iteration step will begin where pulsion integrals in the atomic-orbital basis, which are the new density matrix P(n) is injected into the calcula- required for the evaluation of the Coulomb and exchange (n+1) tion of the next Fock matrix F = F[P(n)]. matrices in Eqs. (7) and (8), may be precalculated as well. As a starting point for the minimization, an initial We implemented all procedures discussed in this paper density matrix P(0) is to be determined, for which vari- into a local version of PySCF 1.5b101. The Def2-SVP102 ous options exist (e.g., superposition of atomic densities, basis set was chosen for all calculations carried out in this extended H¨uckel theory guess, basis set projection, or the work, which are all carried out in the Hartree–Fock ap- diagonalization of the one-electron matrix H). Then, in proximation (all molecular structures are provided in the (n) the n-th iteration step, the Fock matrix F = F[P(n−1)] supporting information). At the example of formalde- is calculated from the density matrix of the previous step, hyde, Fig. 1 shows the structure of the Fock matrix F, P(n−1). eigenvector matrix C, density matrix P, and overlap ma- trix S, to which we later compare transformed matrices The generalized eigenvalue problem of the Roothaan– emerging in the embedding approaches. Hall equation99,100 is solved to obtain the n-th approxi- mation to the eigenvector matrix C(n) and to the diago- nal matrix of eigenvalues (n), IV. FOCK MATRIX BLOCK-DIAGONALIZATION F(n)C(n) = SC(n)(n). (16)

This can be achieved by converting it to the ordinary In this section, we introduce a general decoupling ap- eigenvalue problem proach that eliminates interactions between subsystems of a molecular system at the matrix level. This then S−1F(n)C(n) = C(n)(n). (17) allows for separate treatments of subsystems. Decou- pling is achieved through two subsequent matrix trans- The molecular orbitals are defined by the eigenvectors in formations: first an orthogonalization of the basis, then a the atomic-orbital basis B chosen, block-diagonalization of the orthogonalized Fock matrix. In comparison with orbital localization schemes, which NB NB prepare the one-electron basis and then evaluate the op- (n) X (n) X (n) ξi = ci,j φj = Cij φj. (18) erator matrices in this tailored basis, we manipulate the j=1 j=1 operator matrices to achieve exact block-diagonalization and then obtain basis states that are localized on either (n) subsystem, but not necessarily localized within this sub- The orbital energy of the ξi corre- (n) system. sponds to the diagonal entry ii in the eigenvalue ma- (n) trix  . The Nocc molecular orbitals with the lowest orbital energies enter the Hartree–Fock determinant (in the restricted formalism, Nocc is equal to half the num- A. Transforming the basis (n) ber of electrons Nel). From the matrix Cocc of occupied eigenvectors, We first elaborate on the implications of a modified basis (n)  (n) (n) (n)  for the SCF procedure. As we will consider cases where Cocc = ci ci ··· ci , (19) 1 2 Nocc the transformation would be required in each iteration Quantum System Partitioning at the Single-Particle Level 5

O1 C2 H3 H4 OCH2 O1 C2 H3 H4 O1 C2 H3 H4

FIG. 1. Fock matrix F, eigenvector matrix C, density matrix P, and overlap matrix S (from left to right) of a converged Hartree–Fock calculation for formaldehyde. The color code is such that red indicates negative values, whereas blue stands for positive ones. Then, the matrix-element values covered by the color range from red to blue for the different matrices are as follows: F −2 to +2, C −1 to +1, P −0.4 to +0.4, and S from −1 to +1. Values beyond these ranges are marked in deep red or deep blue.

(n) step, we include the superscript ’(n)’ for the SCF itera- The occupied eigenvector matrix C˜ occ is constructed as tion steps. before and the transformed density matrix P˜ (n) is calcu- lated from the transformed eigenvector matrix C˜ (n), A W transforms the initial basis set B = {φ } into the new basis set B˜ = {φ˜ }, T i i ˜ (n) ˜ (n)  ˜ (n) P = 2 Cocc Cocc (28) NB ˜ X (n) φi = Wijφj. (23) The density matrix P in the original basis B can be (n) j=1 recovered by a back-transformation of P˜ , (n) T ˜ (n) A matrix representation A of an operator Aˆ in the new P = W P W. (29) ˜ basis B can then be expressed through the congruent This back-transformation is necessary because it would transformation, be inefficient to evaluate a new transformed Fock ma- ˜ (n+1) ˜ (n) ˜ T trix F from the transformed density matrix P A = WAW . (24) directly because of the 4-index transformation required for Eqs. (7) and (8). Hence, the calculation of the new It might be not convenient to evaluate the expressions for ˜ (n+1) the transformed Fock matrix in terms of the transformed transformed Fock matrix F may be more efficiently density matrix P˜ in the basis B˜ as it is necessary to achieved by a sequence of backward and forward trans- introduce additional transformation steps into the SCF formations, (n+1) T iterations: F˜ = WF[WTP˜ (n)W]W . (30) With a Fock matrix F˜ (n) in basis B˜ that is evaluated from the transformation of the original Fock matrix F(n), B. Partitioning of the system F˜ (n) = WF(n)WT, (25) the Roothaan–Hall equation becomes In most of the following, we consider a subdivision of a molecular system into two parts, denoted subsystem (n) T −T (n) WF W W C = (S) and environment (E). However, an extension to an (26) WSWTW−TC(n)(n), arbitrary number of subsystems (including hierarchical subsystem nesting) is also discussed. or expressed in terms of the transformed matrix quanti- In this work, subsystem and environment are chosen ac- ties, cording to a partitioning of the atom-centered basis set B into the subsets B and B , such that F˜ (n)C˜ (n) = S˜C˜ (n)(n), (27) S E BS ∪ BE = B, (31) with C˜ (n) = W−TC(n). Solving this new generalized BS ∩ BE = ∅. (32) eigenvalue equation yields the transformed molecular or- (n) bital coefficient matrix C˜ and the eigenvalue matrix The subsets BS and BE consist of nS and nE basis func- (n). tions, respectively. It is convenient to order the basis Quantum System Partitioning at the Single-Particle Level 6 functions φi, assigning the index i such that they consist D. Block-diagonalization of the Fock matrix of two contiguous groups, pertaining to the subsystem and the environment. This leads to the following defini- At the heart of our decoupling method is a transforma- tion for the subsets BS and BE , tion matrix Q separating subsystem and environment at BS = {φi : i ∈ [1; NS ]} , (33) the matrix level. To this end, we seek to transform the ˘ ˜ BE = {φi : i ∈ [NS + 1; NB]} . (34) Fock matrix F to a basis B in which it assumes a block- diagonal form, This ordering ensures that every matrix representation A of an operator Aˆ can be split into distinct subblocks, ˜  ˜ ˘ T FS 0 F = QFQ = ˜ . (41) A A  0 FE A = 11 12 . (35) A21 A22 Here, the block-diagonalization matrix Q needs to be uni-

The subblocks A11, A12, A21, and A22 are of size tary. Otherwise, the eigenvalues will not be invariant and NS ×NS , NS ×NE , NE ×NS , and NE ×NE , respectively. the resulting block-diagonal Fock matrix will not be her- Whenever we encounter block-diagonal matrices where mitian. it is possible to assign each block to either the subsystem Such a block-diagonalization is not unique. Given that or the environment, we will denote the diagonal blocks a matrix Q block-diagonalizes the matrix F˘, then any as AS and AE , matrix DQ, with a unitary block- D, also A 0  transforms F˘ into a block-diagonal form, A = S . (36) 0 A E DQF˘(DQ)T = DQFQ˘ TDT (42) It should be noted that whilst we pursue an atom-wise   ˜   T  D11 0 F11 0 D11 0 partitioning in most of this work, it is possible to have = T (43) 0 D22 0 F˜ 0 D any kind of partition of the basis set B – even basis func- 22 22  ˜ T  tions centered on the same atom can be partitioned into D11F11D11 0 = ˜ T . (44) both BS and BE if deemed useful (consider, for exam- 0 D22F22D22 ple, core-valence separations). If the ordering of the ba- sis functions differs from the specification above, it will This means that the transformation matrix Q can, in be trivial to construct permutation matrices which will principle, only be determined up to a unitary block- produce a block structure according to this partitioning diagonal matrix. scheme. In addition to the non-uniqueness of the block- diagonalization matrix Q, there also exist multiple al- gorithms to determine such a matrix. For example, one C. Orthogonalization of the overlap matrix may subject the Fock matrix F˘ to subsequent Givens rotations104 to eliminate those off-diagonal matrix ele- ments which represent the system-environment coupling. We now transform the generalized eigenvalue problem in This is essentially equivalent to the Jacobi eigenvalue Eq. (16) into an ordinary eigenvalue equation. This can algorithm105, targeting only elements in the off-diagonal be done with L¨owdin’ssymmetric orthogonalization103, blocks. However, this iterative procedure is tedious and a congruent transformation, may have difficulties to achieve off-diagonal blocks to be XSXT = I, (37) sufficiently close to zero. Another example is the House- holder block-diagonalization algorithm in which block re- with flectors are applied in an iterative fashion to eliminate − 1 106 X = S 2 . (38) off-diagonal blocks. However, it suffers from the same shortcomings as the Jacobi algorithm. Also, all of these As it is well known, we arrive at a transformed eigenvalue algorithms do not offer any kind of control and physi- equation, cal insight into how the subsystem and environment are XFXTX−TC = XSXTX−TC, (39) separated from one another. which can be rewritten as, In an attempt to obtain a well-defined block- diagonalization matrix Q we require it to fulfill additional F˘C˘ = C˘ , (40) constraints. For practical purposes it may be most con- venient to have a transformed basis B˜ that is as close with the definition of the transformed Fock matrix F˘ = ˘ T 1 as possible to the initial basis B. This means that the XFX and its eigenvector matrix C˘ = S 2 C. The trans- ˘ block-diagonalization matrix Q is as close as possible to formed Fock matrix F is hermitian. The diagonal matrix the ,  containing the eigenvalues is invariant under this trans- formation. The structure of the matrices in Fig. 1 after Q = argmin (kQ − Ik) . (45) orthogonalization is depicted in Fig. 2. Q Quantum System Partitioning at the Single-Particle Level 7

O1 C2 H3 H4 OCH2 O1 C2 H3 H4 O1 C2 H3 H4

FIG. 2. Orthogonalized Fock matrix F˘, eigenvector matrix C˘ , density matrix P˘ , and orthogonalization matrix X (from left to right) of a Hartree–Fock calculation for formaldehyde. Color coding as in Fig. 1 and for X we have a range of −1 to 1.

Under these circumstances, matrices will be subject to It also offers an additional layer of control considering minimal changes upon such a transformation, leaving the composition of subsystem and environment. Since their matrix structure mostly intact. Subsequent approx- the order of eigenvectors in an eigenvector matrix is ar- imations that are based on such a minimal-effect separa- bitrary, we can freely choose which eigenvectors of C˘ are tion scheme can be expected to feature smallest errors. projected into the subsystem or the environment basis. In the following, we discuss how such a minimal-effect To proceed with a unique construction of the block- block-diagonalization matrix Q can be constructed for diagonalization matrix Q, we need to specify how eigen- which the condition in Eq. (45) has been proven to be vectors from C˘ are assigned to either the subsystem C˘ S 107 fulfilled by Cederbaum et al. or the environment C˘ E . In this work, eigenvectors are assigned according to a simple localization scheme. For First, we express the block-diagonalization in Eq. (41) in each eigenvector c there exists an associated localization terms of the eigenvectors C˘ and C˜ , i function fi describing by how much the first NS basis F˜ = C˜ C˜ T = QC˘ C˘ TQT, (46) functions of the orthogonalized basis contribute to the corresponding molecular orbital. The localization func- with C˜ = QC˘ . Since the transformed Fock matrix F˜ is tion employed in this work is given by, block-diagonal, we may also write the eigenvector matrix NS C˜ in a block-diagonal form, X 2 fi = c˘i,j. (51)  ˜  j=1 ˜ ˘ CS 0 C = QC = ˜ . (47) 0 CE It is, of course, possible to construct other localization functions that could be used to assign eigenvectors to ei- This means that the first N eigenvectors of C˘ are trans- S ther the subsystem or the environment. Then, the eigen- formed into a basis B˜ in which they are located entirely vector matrix C˘ is constructed as, on the transformed subsystem basis functions of B˜S . Ac- cordingly, the last N eigenvectors of C˘ are transformed  E C˘ = c˘1 c˘2 ··· c˘N . (52) such that they are located entirely on the transformed B ˜ environment basis functions of BE . This assignment of where the eigenvectors are sorted in a descending order eigenvectors implies that the eigenvector matrix C˘ may according to their localization function, such that be written in terms of C˘ S and C˘ E , such that f ≥ f ≥ ... ≥ f . (53) ˘  1 2 NB C = C˘ S C˘ E . (48) This eigenvector ordering ensures that the N eigenvec- ˜ ˜ S The partitioning of the basis B into the subsets BS and tors with the highest localization function are present in B˜ is defined analogously to Eq. (34), E C˘ S and the remaining eigenvectors are present in C˘ E . n ˜ o Now that the eigenvector assignment has been discussed, B˜S = φi : i ∈ [1; NS ] , (49) we can continue with the construction of the unitary n ˜ o B˜E = φi : i ∈ [NS + 1; NB] . (50) block-diagonalization matrix Q. In the approach by Cederbaum et al.107, it is constructed as a product of Rewriting the block-diagonalization of the Fock matrix two matrices QR and QBD, in terms of the eigenvectors greatly simplifies the prob- lem of determining the block-diagonalization matrix Q. Q = QRQBD. (54) Quantum System Partitioning at the Single-Particle Level 8

This splits the procedure into two steps. First, matrix Note that this transformation matrix W is, in general, QBD block-diagonalizes the eigenvector matrix C˘ . Then, not unitary. The diagonal blocks F˜ S and F˜ E are given the matrix QR renormalizes the transformation, guaran- by teeing that the total block-diagonalization matrix Q is T T unitary. The matrix Q takes on the form, F˜ S =W11F11W + W12F21W BD 11 11 (62)  T T T I −U + W11F12W12 + W12F22W12, Q = . (55) BD UI T T F˜ E =W21F11W + W22F21W 21 21 (63) T T It block-diagonalizes the eigenvector matrix C˘ , + W21F12W22 + W22F22W22. −C˘ + UTC˘ −C˘ + UTC˘  ˘ 11 21 12 22 The one-electron matrix H˜ and the two-electron matrix QBDC = ˘ ˘ ˘ ˘ . (56) UC11 + C21 UC12 + C22 V˜ are transformed in the same way as the Fock matrix ˜ Inspection of the off-diagonal blocks which are supposed F. However, these matrices are in general not block- ˜ to vanish, yields the following solution for U, diagonal, only their sum F is. Fig. 3 provides a graphical representation of the transformed matrices from Fig. 1. T ˘ ˘ −1  ˘ ˘ −1 U = −C21C11 = C12C22 . (57) We already note at this stage that the approach by Cederbaum et al. closely relates to relativistic exact two- Note that this matrix U can be constructed from either ˘ ˘ ˘ component methods, which we discuss later in this work. the subsystem eigenvectors CS (as C11 and C21) or the We will also show how the system-environment sepa- ˘ ˘ ˘ environment eigenvectors CE (as C12 and C22). A proof ration according to this scheme can be generalized to- of this equality can be found in Appendix A. The renor- ward an arbitrary number of subsystems. Because of malization matrix QR is then given by, these facts, we underline the general applicability of the − 1 unitary block-diagonalization approach by assigning the T  2 QR = QBDQBD (58) general term ’subsystem separation by unitary block- − 1 ! I + UTU 2 0 diagonalization’ or SSUB (to be pronounced ’sub’) to this = − 1 (59) approach. 0 I + UUT 2 Finally, the block-diagonalization matrix Q reads,

 T − 1 T − 1 T E. Partitioning into an arbitrary number of subsystems (I + U U) 2 −(I + U U) 2 U Q = T − 1 T − 1 . (60) (I + UU ) 2 U (I + UU ) 2 In the literature concerning this kind of block- In the previous section, we considered a separation of a diagonalization,107–110 there exist multiple notations for Fock matrix into two subsystems. Starting from one of the actual form of the block-diagonalization matrix Q. these subsystems, we may apply another unitary block- In Appendix B we show that all of these different repre- diagonalization step, which separates the subsystem into sentations are in fact identical. another two subsystems. Clearly, this allows one to split a total system into any number of subsystems. This pro- Whereas L¨owdin’stransformation is a minimal orthogo- cess is only limited by the finite number of basis func- nalization, the Cederbaum scheme represents a minimal tions, which, in the limit, corresponds to the exact di- transformation in the sense that the new basis is as sim- agonalization of the Fock matrix, which has been the ilar as possible to the old one. It is therefore no surprise starting point for the construction of the unitary trans- that orbital localization schemes have been considered formation matrices. in the literature111,112 that are based on a Cederbaum- type transformation, which relates to our approach here, In the following, we discuss the technicalities of the sep- but focuses on the preparation of particular basis states aration of the system into multiple subsystems, following rather than on the block-diagonalization of the Fock ma- the formalism elaborated above. The system shall be trix. Interestingly, Ref. 112 notes that the Cederbaum separated into k subsystems, denoted by Si each. Each scheme can be efficiently evaluated with techniques de- subsystem consists of NSi basis functions. This gener- veloped for the exact decoupling of the Dirac Hamilto- alization of the bipartition in section IV B leads to the nian, however without considering further implications following partitioning of the basis, on the relation of relativistic exact decoupling to a gen-  i−1 i  eral embedding approach such as SSUB introduced here.  X X  BSi = φi : i ∈ [1 + NSj ; NSj ] . (64) The total transformation matrix W transforming the  j=1 j=1  original Fock matrix F into the block-diagonal form F˜ then takes the form ˘ − 1 The eigenvectors in eigenvector matrix C are localized W = QS 2 . (61) such that they are located on the respective parts of the Quantum System Partitioning at the Single-Particle Level 9

O1 C2 H3 H4 O CH2 O1 C2 H3 H4 O1 C2 H3 H4

FIG. 3. The transformed Fock matrix F˜, eigenvector matrix C˜ , density matrix P˜ , and transformation matrix W (from left to right) resulting from the matrices depicted in Fig. 1. The block-diagonalization procedure was applied to decouple the oxygen atom from the other atoms of formaldehyde. Color coding as in Fig. 1 and for W we have a range of −1 to +1. system. Unfortunately, it is not trivial to assign each Q(i) as, eigenvector to a single subsystem when partitioning the system into multiple subsystems. A detailed description  (i)  of how this was done in this work can be found in Ap- I 0 0 (i)  (i) (i) pendix C. Analogously to the partitioning in Eq. (48), Q = 0 Q11 Q12 , (69)  (i) (i) this means that the eigenvector matrix can be written 0 Q21 Q22 as,

˘ ˘ ˘ ˘  with the matrix subblocks Q(i), Q(i), Q(i) and Q(i) being C = CS1 CS2 ··· CSn , (65) 11 12 21 22 constructed exactly as in Eq. (60). The dimension d(i) of ˘ the identity matrix subblock I(i) is given by, where each of these eigenvector matrices CSi consists of

NSi eigenvectors. i−1 X This ordered eigenvector matrix is then block- d(i) = N . (70) diagonalized into k subsystems by a sequential applica- Sj j=1 tion of a block-diagonalization matrix. The total block- diagonalization matrix Q can be written as the product of each of these individual transformations, The matrix U(i), which is required for the construction of the transformation matrix Q(i) can be expressed (in x k−1 analogy to Eq. (57)) in terms of the intermediary subsys- Y (i) (k−1) (k−2) (2) (1) tem eigenvector matrix Cˇ (i−1), Q = Q = Q Q ··· Q Q (66) Si i=1    −1 (i) (i) ˇ (i−1) ˇ (i−1) Here, each block-diagonalization matrix Q produces U = − CS CS , (71) ˜ i L i D the diagonal block CSi . For the intermediary eigenvector matrices Cˇ (i), we can write a recursion relation, where the subscripts ’D’ and ’L’ denote the NSi ×NSi Cˇ (i+1) = Q(i+1)Cˇ (i), (67) diagonal block and the remaining lower off-diagonal block of the matrix Cˇ (i−1), respectively. Si with Cˇ (0) = C˘ and Cˇ (k−1) = C˜ . Each intermediary eigenvector matrix Cˇ (i) can also be written as, With this recursive scheme, it is now possible to con- struct a block-diagonalization matrix Q which block-   diagonalizes the Fock matrix into k subsystems. C˜ S 0 0 0 1 Two such sequential steps were carried out to arrive at  ..  (i)  .  C˜ =   . (68) a decomposition of formaldehyde into three subsystems:  0 C˜ 0 0  the oxygen atom, the carbon atom, and the two hydrogen  Si  ˇ (i) ˇ (i) atoms (see Fig. 4). As another example, we considered 0 0 CS ··· CS i+1 k a core-valence separation. The subsystems consist of the eigenvectors that describe the two 1s-like molecular or- From this representation of the intermediary eigenvec- bitals of the carbon and oxygen atoms and of all other tor matrices, we can construct the transformation matrix orbitals (shown in Fig. 5). Quantum System Partitioning at the Single-Particle Level 10

O1 C2 H3 H4 O C H2 O1 C2 H3 H4 O1 C2 H3 H4

FIG. 4. The transformed Fock matrix F˜, eigenvector matrix C˜ , density matrix P˜ , and transformation matrix W (from left to right) of formaldehyde, in which three subsystems were separated from each other: the oxygen atom, the carbon atom, and the hydrogen atoms. Note that the hydrogen subblock of the density matrix P˜ vanished. This is because the eigenvectors which were assigned to this subsystem during the localization procedure are not occupied. Instead, the occupied eigenvectors which contributed to the hydrogen atoms are now entirely located on the carbon-atom subsystem. Note that this does not correspond to a charge transfer as it is simply a matter of representation in the transformed basis. The color code is as in Fig. 3.

1s O1val C2val H3 H4 1s val 1s O1val C2val H3 H4 1s O1val C2val H3 H4

FIG. 5. The transformed Fock matrix F˜, eigenvector matrix C˜ , density matrix P˜ , and transformation matrix W (from left to right) of formaldehyde, in which the core eigenvectors of the carbon and oxygen atoms were separated from the rest of the system. This is an example of explicit core-valence separation. The color code is as in Fig. 3.

V. EXPLOITING THE BLOCK-DIAGONAL than for the total system. STRUCTURE IN AN APPROXIMATE SCF PROCEDURE In principle, this SCF procedure follows the same steps as the SCF procedure with a transformed basis in sec- tion IV A. However, the computational disadvantages We now proceed to introduce an approximate SCF arising from the additional forward and backward trans- scheme that avoids repeatedly solving the eigenvalue formations in the exact formulation are alleviated by an problem for the whole system. The scheme is based on advantageous representation of the eigenvalue problem, the block-diagonalization of the Fock matrix F˜, leading in which the dimension of the problem is significantly to an advantageous representation of the eigenvalue prob- reduced: lem encountered in the Roothaan–Hall equations.

With the initial density matrix P(0) an initial Fock ma- (1) trix F can be constructed. Following the steps laid out A. An approximate SCF procedure in a previous section concerning block-diagonalization, − 1 we obtain the total transformation matrix W = QS 2 from this Fock matrix F(1). In general, this step will The transformed Fock matrix F˜ (n) is obtained by trans- be computationally expensive, since it involves (i) solv- forming the Fock matrix F(n) with the transformation ing the eigenvalue problem for the whole system and (ii) matrix W. Since the Fock matrix varies throughout diagonalizing the overlap matrix for the calculation of the SCF procedure, exact block-diagonalization would re- the inverse of its square-root required for Q in Eq. (60). quire the construction of a block-diagonalization matrix Clearly, computational advantages will only emerge for Q in each iteration step. This requires the solution of subsequent steps performed for a subsystem only rather the full eigenvalue problem and subsequent matrix inver- Quantum System Partitioning at the Single-Particle Level 11 sions and therefore the repeated evaluation of the block- eigenvalues and eigenvectors of the block-diagonal Fock diagonalization matrix Q would be computationally in- matrix F˜ (n) can be assigned to either the subsystem or efficient. However, if we can construct a sufficiently good the environment, it is possible to construct the occupied initial density matrix (such as one taken from a calcula- ˜ (n) ˜ (n) eigenvector matrices CS,occ and CE,occ from the eigen- tion on a very similar molecular structure as, for instance, vector matrix blocks C˜ (n) and C˜ (n), respectively. encountered in structure optimizations or first-principles S E molecular dynamics trajectories), we may expect changes The transformed density matrix P˜ (n) is calculated from in the Fock matrix over the course of the SCF procedure the occupied eigenvectors according to Eq. (28). This to be comparatively small. Then, the initial transforma- matrix is block-diagonal since the eigenvector matrix tion matrix W can be kept over the course of a calcu- (n) C˜ occ takes a block-diagonal form. In terms of the occu- lation and still block-diagonalize the resulting Fock ma- (n) (n) (0) ˜ ˜ ˜ (n) trices sufficiently well. If the initial density matrix P pied eigenvector matrices CS,occ and CE,occ, P reads deviates too much from the converged one, the proposed scheme will fail to find the self-consistent solution.   T  C˜ (n) C˜ (n) 0 ˜ (n) ˜ (n) ˜ (n) S,occ S,occ The diagonal blocks F and F can be evaluated as in P = 2  T . S E  ˜ (n)  ˜ (n)   Eqs. (62) and (63), respectively. However, if we assume 0 CE,occ CE,occ the environment to change much less than the subsystem, (77) ˜ (n) we may leave FE constant, ˜ (n) ˜ (1) FE = FE . (72) The density matrix P(n) is obtained through back- transformation of the transformed density matrix P˜ (n). Note that such an assumption may later be lifted in al- Since the transformed density matrix P˜ (n) is block- ternating freeze and thaw cycles. diagonal, the back-transformation in Eq. (29) can be sim- Transforming the Roothaan–Hall equation with the plified to yield the density matrix P(n), − 1 transformation matrix W = QS 2 according to Eq. (26) requires us to solve an ordinary eigenvalue problem,  T ˜ (n) T ˜ (n)  W11PS W11+ W11PS W12+ T (n) T (n) ˜ (n) ˜ (n) ˜ (n) (n)  W P˜ W21 W P˜ W22  F C = C  . (73) (n)  21 E 21 E  P =   . (78)  T ˜ (n) T ˜ (n)  Since the transformed Fock matrix F˜ (n) is block-diagonal, W12PS W11+ W12PS W12+  T ˜ (n) T ˜ (n) the whole eigenvalue problem assumes a block-diagonal W22PE W21 W22PE W22 form, ! ! Note that it will be possible to precompute the terms F˜ (n) 0 C˜ (n) 0 ˜ (n) S S = arising from PE , if this part of the transformed density ˜ (n) ˜ (n) 0 FE 0 CE matrix is kept constant throughout the SCF procedure ! ! (74) (for a frozen environment). C˜ (n) 0 (n) 0 S S . ˜ (n) (n) As an example, we chose a simple structure that rep- 0 CE 0 E resents a typical embedding situation, i.e., a solute This allows us to separate the eigenvalue problem into surrounded by solvent molecules; here, an acetonitrile two smaller eigenvalue problems molecule surrounded by seven water molecules. A stan- dard HF calculation is carried out on the initial equi- ˜ (n) ˜ (n) ˜ (n) (n) FS CS = CS S (75) librium structure to obtain a converged density matrix. Then, one of the C–H bonds is elongated. For the sub- and sequent structures with the elongated C–H bond, both F˜ (n)C˜ (n) = C˜ (n)(n), (76) a standard SCF calculation (as a reference for the exact E E E E electronic energy) and a calculation in the approximate both of which can be solved separately. SCF procedure exploiting the block-diagonal form of the ˜ (n) ˜ (1) transformed Fock matrix with a frozen environment of However, if we set FE = FE , the second eigenvalue problem need not be solved as its solution has already seven water molecules are performed. For the approxi- been obtained for the construction of W. The eigenvec- mate calculation, the initial density matrix is taken from the converged calculation on the equilibrium structure tor matrix C˜ (1) and the eigenvalue matrix (1) can be E E performed initially. This allows us to demonstrate how stored for later use. Therefore, it is sufficient to solve the large a structural perturbation may be in order for the subsystem eigenvalue problem in Eq. (75). approximate SCF to be sufficiently accurate. The results Selecting the occupied eigenvectors follows the standard of the calculations for the perturbed structures are sum- SCF procedure. Nocc eigenvectors with the lowest corre- marized in Fig. 6. Up to an elongation of around 0.3 A,˚ sponding eigenvalues are occupied with electrons. As the the error is within chemical accuracy of 1 kcal mol−1. Quantum System Partitioning at the Single-Particle Level 12

500 25 6

5 400 20 ] ] ] 1 1 4 l % l [ o

300

o 15 l e m m

E J

3 J / k

k [ l

[ e l 200 10 l e E e

E 2 E 5 100 1

0 0 0 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 Elongation [Å] Elongation [Å]

FIG. 6. Acetonitrile subsystem solvated in a water environment (middle). Electronic energy (given relative to that of the equilibrium structure) of the reference (solid) and approximate (dashed) SCF procedures for increasing C–H bond elongation (left). Error in electronic energy introduced by the approximate SCF scheme (right).

B. Reducing the number of integrals in the two-electron in a KS-DFT framework, the approximation will take a matrix slightly different form as only parts of the electron density do not need to be reevaluated,

(n) (n−1) The Fock matrix F = F[P ] is calculated from the NS (n−1) X (n−1) density matrix P . When evaluating the two-electron (n−1) ρ[P ] ≈ Pkl φkφl + δρ, (82) (n) (n−1) matrix V = V[P ] it is possible to introduce ap- kl proximations that make it possible to drastically reduce the number of required two-electron integrals. with

If subsystem and environment are well separated (either NS (0) X (0) with respect to molecular structure in real space or in δρ = ρ[P ] − Pkl φkφl. (83) Hilbert space indicated by some energy gap), it will be kl possible to reduce the number of two-electron integrals required. This may facilitate the evaluation of the two- (n) (n) In addition, the two-electron matrix V can be approxi- electron matrix V = V[P(n−1)] after the first SCF it- mated by restricting the iterative re-evaluation to certain eration step. matrix elements of V(n) only, most conveniently those in First, improvements in efficiency may be achieved by (n) the V11 block, leaving the coupling and the pure-environment blocks of ! the density matrix P unchanged, V(n) V(1) V(n) ≈ 11 12 , (84) ! (1) (1) (n) (0) V21 V22 (n) P11 P12 P ≈ (0) (0) . (79) P21 P22 which even further reduces the number of two-electron integrals to be evaluated. If P(n) is approximated as shown above, it is not nec- essary to re-evaluate certain terms of the two-electron matrix in each SCF iteration step. For example, the Coulomb matrix J(n) we may approximate as VI. PROJECTOR-BASED EMBEDDING THEORY

NS 1 (n−1) X (n−1) Jij[P ] ≈ Pkl hφiφj| |φkφli + δJij, (80) We are now in a position to discuss the embedding of r12 kl a subsystem described by a high-level theory within the with the correction δJ collecting all terms that are not low-level mean-field method. This allows us to directly ij relate our work to the recent work by Miller, Manby, purely subsystem-dependent and may be evaluated only 55 once in the beginning of the SCF procedure, and co-workers as well as to Huzinaga and Cantu and H´egelyet al.66. These embedding methods are based on NS 1 an augmentation of the Fock operator with a projection (0) X (0) δJij = Jij[P ] − Pkl hφiφj| |φkφli. (81) operators. Therefore, we refer to them as projector-based r12 kl embedding methods. An analogous expression can be written for the exchange For projector-based embedding methods, an initial calcu- (n) matrix K . For the exchange-correlation matrix Vxc lation on the whole system with a low-level DFT method Quantum System Partitioning at the Single-Particle Level 13 is usually carried out. Quantities from this initial calcu- principle, in the limit of µ tending to infinity, exact re- lation are denoted with a superscript 0(0)0. Then, the re- sults can be obtained. However, due to numerical issues, sulting eigenvectors in the eigenvector matrix C(0) from it is recommended to use values for µ of a few thousand this initial calculation are localized in such a way that Hartree. 0 they can be assigned to either the subsystem or the envi- The Huzinaga–Cantu matrix FHC commutes with the ronment. This produces C(0) and C(0), which introduce (0) (0)T S E matrix SCE CE S and therefore shares its eigenvectors a corresponding separation in the density matrix P(0), (0) (0) CE . being the direct sum of the two density matrices PS (0) (0) (0)T and PE , Here, we sketch how the additional term µSCE CE S 0 in the Miller–Manby Fock matrix FMM forces the com-  T P(0) = 2 C(0) C(0) (85) posite Fock matrix Fcomp to assume the eigenvectors S S,occ S,occ (0) CE . We present an intuitive (but by no means rigor- and ous) way to rationalize how these environment eigenvec- tors are recovered and how the energetic shift can be T (0) (0)  (0)  understood. PE = 2 CE,occ CE,occ , (86) First, we realize that the generalized eigenvalue problem which implicitly assigns a number of electrons to the sub- for Fcomp in the Roothaan–Hall equation in Eq. (16) can system and the environment. also be written as,

In the following, we seek the self-consistent solution of a T Fcomp = SCC S. (92) composite Fock matrix Fcomp, mixing the two mean-field methods, Splitting this representation of Fcomp into its subsystem and environment parts yields, Low (0) High Low (0) Fcomp = F [P ] + F [PS ] − F [PS ]. (87) T T Fcomp = S CS S C + CE E C S. (93) In projector-based embedding methods, it is required S E that the new subsystem eigenvectors in C are orthog- T S By adding the term µSC(0)C(0) S, the modified Fock onal to the environment eigenvectors in C(0) (which are E E E matrix F0 is obtained, kept constant), MM  T (0) 0 T T (0) (0) T F = S CS S C + CE E C + µC C S. CS SCE = 0. (88) MM S E E E The expression for the electronic energy of such a com- (94) posite calculation is given by, (0) Assuming that CE ≈ CE and that for all eigenvalues Low (0) High Low (0) | |  µ holds, we arrive at Ecomp = E [P ] + E [PS ] − E [PS ]. (89) ii

 T 0 T (0) (0) To ensure that orthogonality is preserved, the Fock ma- FMM ≈ S CS S CS + CE (E + µI)CE S. (95) trix has to be modified even further such that its eigen- (0) vectors include the eigenvectors in CE . This is done by Since µ is such a large number (usually a few orders of augmenting the Fock operator with projection operators. magnitude larger than the largest eigenvalues), the eigen- In that case, solving the eigenvalue equation for this new vectors C(0) dominate the matrix F0 . In this represen- 0 E MM Fock matrix F will lead to subsystem eigenvectors CS tation, it is easy to see why µ is also referred to as an that are necessarily orthogonal to those of the environ- energy shift applied to the environment eigenvectors. ment. Such modified Fock matrices F0 were proposed by Miller, Manby, and co-workers61,63, Interestingly, the projector-based embedding approach can be used in conjunction with our approximate SCF T F0 = F + µSC(0)C(0) S, (90) procedure. As noted in section V, in the approximate MM comp E E SCF procedure the same transformation matrix W may and also by Huzinaga and Cantu55,66, be applied throughout the whole procedure. Since the Fock matrix changes during the optimization, the block- (0) (0)T (0) (0)T diagonalization is not exact, introducing errors. How- F0 = F − SC C F − F C C S. HC comp E E comp comp E E ever, this is not the case for the modified Fock matri- (91) 0 0 ces FMM and FHC. Since the environment eigenvectors (0) Whereas the former is approximate, the latter is exact. CE are kept constant, exact block-diagonalization can The factor µ introduced in the Miller–Manby Fock ma- be achieved for the whole SCF procedure. This implies 0 trix FMM can be understood as an energy shift which is that once the transformation matrix W has been calcu- (0) applied to the molecular orbitals of the environment. In lated from the eigenvectors CS according to Eq. (60), Quantum System Partitioning at the Single-Particle Level 14 projector-based embedding methods can be applied, solv- While this seems somewhat reminiscent of the Fock ma- ing the eigenvalue problem for the subsystem only. trix expression in the projector-based embedding ap- proach introduced earlier, it is different. Here, F [P ] We may now compare the Miller–Manby embedding 11 11 denotes the F block of a Fock matrix where only the method with the approximate block-diagonalized variant 11 subsystem block of the density matrix was used to eval- suggested in this section. We chose again formaldehyde uate all contributions, and partitioned it by cutting through the C=O double bond. This leaves us with a subsystem containing a sin- (  P11 0  Fij 0 0 : i, j ≤ NS gle oxygen atom and the environment containing a car- Fij = . (97) bon and two hydrogen atoms. The eigenvectors were lo- 0 : else calized according to the cost function in Eq. (51). For the high- and low-level mean-field schemes we simply This Fock matrix can be used in the SCF procedure as chose Hartree–Fock and density functional theory with usual to obtain a self-consistent solution. In distinction the BP86 functional113,114, respectively. We varied the to the projector-based embedding approaches in the pre- parameter µ between 100 and 1000000 Hartree (values vious section, no initial calculation on the whole system of at least a few thousand Hartree are recommended in is required. Instead, both high- and low-level mean-field Refs. 61 and 63). The results are shown in Table I. For methods are converged to a self-consistent solution si- values of µ above 10000 Hartree, it is possible to obtain multaneously. The EMFT energy expression takes the the reference energy up to within five decimals with the composite form, approximate method. Low High Low Eel[P] = Eel [P] + Eel [P11] − Eel [P11]. (98) TABLE I. Electronic energies obtained with the Miller– Manby projector based embedding method and their approx- imate subsystem SCF counterparts for different values for B. EMFT with block-orthogonalized partitioning parameter µ (in Hartree) for a BP86-in-HF calculation of formaldehyde. After the introduction of EMFT, it became apparent that µ MM BD-MM some EMFT calculations exhibit an unphysical collapse 100 -113.7078797 -113.7080317 54 500 -113.7106458 -113.7106769 of the self-consistent solution. This has been attributed 1000 -113.7109922 -113.7110081 to a mismatch in the functional forms of the high- and 54 5000 -113.7112694 -113.7112732 low-level density functionals employed. 10000 -113.7113041 -113.7113063 The unphysical collapse manifests itself in a lowering 100000 -113.7113353 -113.7113362 1000000 -113.7113384 -113.7113391 of the electronic energy of several thousand Hartree. Also, electron population analysis reveals that the col- lapse is accompanied by extraordinarily high electron populations in the subsystem and environment blocks, Tr (P11S11) and Tr (P22S22), respectively. Since the VII. RELATION TO EMBEDDED MEAN-FIELD total number of electrons is constant, large negative THEORY populations in the coupling blocks Tr (P12S21) and Tr (P21S12) are generated. As a consequence, other ob- The embedded mean-field theory52 (EMFT) also aims at servables are also affected and can be wrong by several embedding a subsystem in an environment. EMFT is orders of magnitude (consider, for example, the dipole applied to describe the subsystem with a computation- moment). ally expensive, high-level density functional within an To prevent this collapse of the self-consistent EMFT solu- environment described with a cheaper, lower-level den- tion it was modified to operate on a different partitioning sity functional. Most importantly, in comparison to the of the system.54 Instead of partitioning the system based projector-based embedding schemes, it is not necessary on atomic orbitals, the scheme is applied in a block- to calculate a self-consistent solution with the low-level orthogonalized basis (BOEMFT). This means that all method in advance. matrix operators must be transformed accordingly. The transformation matrix W, which block-orthogonalizes the basis, is given by54 A. Embedded mean-field theory  I 0 W = −1 . (99) −S21S11 I The EMFT Fock matrix is a composite Fock matrix com- bining high- and low-level mean-field theories, Transforming a basis B into the basis B˜ with this trans-

EMFT Low High Low formation matrix W according to Eq. (23) leaves the F [P] = F [P] + F11 [P11] − F11 [P11]. (96) basis functions of the subsystem invariant. Therefore, Quantum System Partitioning at the Single-Particle Level 15 the subsystem subblock A11 is equal to the transformed block-diagonalization of the four-component Dirac-based subsystem block A˜ 11 of the transformed matrix A˜ . Fock operator from its eigenvectors,

Since the transformation is a block-orthogonalization, the X2C −1 U = −CSLC , (103) transformed overlap matrix S˜ is block-diagonal, LL   (where S and L refer to the small- and large-components ˜ S11 0 of the molecular 4-spinors, respectively). It was then S = ˜ , (100) 0 S22 realized that the free-particle Foldy–Wouthuysen trans- formation can be skipped and Eq. (103) can be applied with directly to the Dirac-based Fock operator128–130 (writ- ten in terms of Dyall’s modified Dirac equation131). This ˜ S22 = S22 − S21S11S12. (101) approach was later called exact two-component (X2C) decoupling.108,128–130,132–135 In analogy to the EMFT Fock matrix in Eq. (96), the transformed BOEMFT Fock matrix F˜ BOEMFT is con- As such, the X2C approach is identical to the block- 107 structed according to diagonalization approach used by Cederbaum et al. earlier, but in a different context (cf. Eq. (57)). Cau- ˜ BOEMFT ˜ Low ˜ High ˜ ˜ Low ˜ tion is advised when comparing different notations for F [P] = F [P] + F11 [P11] − F11 [P11]. (102) the block-diagonalization matrices of Eq. (60) with those encountered in the literature.107–110 At first glance, all of them seem to be slightly different, but they are, in fact, Instead of the subblock P11 of the density matrix, the all equivalent. transformed density matrix subblock P˜ 11 is used. This is an approximation, producing a Fock matrix evaluated X2C is usually viewed as a way to eliminate negative- for unphysical subsystem electron densities. energy states that are separated from the ’electronic’ The BOEMFT idea seems to resemble our approximate states by twice the electron rest energy per electron in SCF procedure, with a basis transformation partitioning the system. This energy gap is huge because it depends, subsystem and environment. However, it is fundamen- according to Einstein, on the squared speed of light times tally different. In BOEMFT the transformation is only the rest mass of an electron. The reduction of 4-spinors to applied to prevent the unphysical collapse of the self- 2-spinors essentially requires the elimination of all small- consistent field procedure, whereas in our case it is the component basis functions. The contribution of these decisive starting point. basis functions to ’electronic’, i.e., positive-energy molec- ular 4-spinors are usually atom centered and mostly con- served in chemical reactions.136 Accordingly, one is not surprised that, for physical reasons, X2C works so well at VIII. RELATION TO RELATIVISTIC DECOUPLING producing an effective electrons-only Hamiltonian for 2- THEORIES component calculations. However, in view of SSUB this separation of electronic and positronic states is nothing but a system-environment embedding, in which the effect The block-diagonalization of Dirac-based Fock operators of the positronic states is folded into the one-electron has been a desire for physical, formal, and numerical rea- two-component X2C Fock operator. Hence, no energy sons as one wants to decouple electronic bound states criteria need to be invoked to justify the separation of from the negative-energy (continuum) states in order to the negative-energy states, they are only required to iden- arrive at variational electrons-only Hamiltonians suitable tify the one-particle eigenstates whose spinor energies are for applications in molecular physics and chemistry.115 smaller than twice the rest energy of an electron. SSUB can be directly related to this block- In fact, our analysis of SSUB shifts the focus from the diagonalization of the one-electron Dirac operator116, physical picture to the actual mechanism in one-particle which is very efficiently formulated in basis-set represen- Hilbert space, which allows us to better understand the tation. In 2005, Jensen117 proposed a scheme to decouple decisive approximations that are in operation in prac- the negative-energy states (sometimes called ’positronic tical X2C implementations. SSUB applied to a four- states’) of the Dirac operator that avoids the involved component Dirac-based Fock operator produces the X2C algebra of Douglas–Kroll–Hess transformations.118–120 Fock operator, but requires all eigenvectors of the origi- This work was driven by the desire of rewriting the nal four-component operator. This implies that the so- Dirac Hamiltonian for electrons-only problems without lution must already be known for the construction of a introducing approximations120–127. Jensen started from unitarily transformed operator (note the relation to the a free-particle Foldy–Wouthuysen transformed Dirac formal projection operators built from the eigenstates of Hamiltonian, which in Douglas–Kroll–Hess theory is the Dirac Hamiltonian proposed by Mittleman137,138 and the mandatory first step120. However, his central Sucher139; in particular, see Ref. 140). In relativistic idea then was to construct the for the quantum chemistry this procedure is nevertheless advan- Quantum System Partitioning at the Single-Particle Level 16 tageous as subsequent electron-correlation methods ben- may as well exploit approaches developed to cope with efit largely from the elimination of the negative-energy such situations for X2C approaches. spinor states in the (preparatory) four-index transforma- tion, which also holds true for general embedding schemes The preparation of the X2C operator for large molecules based on SSUB-like ideas. has prompted the development of approximate decou- pling schemes at the level of the Hamiltonian itself.144–146 Approximate approaches are usually applied that restrict A more accurate approximation is obtained at the level and/or model the contributions to the potential. Var- of the unitary transformation itself, which decomposes it ious such approximations were proposed to obtain ap- into atomic blocks,109,110 proximate eigenvectors for the construction of the unitary M matrix in Eq. (60). The most popular one, which is also Q = QII , (104) applied in standard applications of sequential Douglas– I Kroll–Hess decoupling transformations118, is the com- plete neglect of all electron-electron interaction terms, where I denotes a subsystem (typically an atom) for which alleviates the problem of obtaining a transformed which one-particle eigenstates have been determined. form of these terms. In other words, the iterative con- which was called diagonal local approximation to the uni- struction of the unitary matrix discussed in the con- tary decoupling transformation (DLU) in Ref. 109 and text of the modified SCF scheme in section V is then local unitary transformation (LUT) in Ref. 110. avoided. The unitary transformation is then constructed from eigenvectors obtained by diagonalizing an approxi- mate Roothaan–Hall equation with the external electron- A. Local SSUB nuclei potential as the only potential. Naturally, at- tempts were made to improve on this restricted model by introducing mean-field potentials which affect the Here, we introduce the concept of an approximative lo- eigenvectors and, hence, the unitary block-diagonalizing 108,141–143 cal construction of the block-diagonalization matrix Q, matrix. As such, these approximate and popu- similar to the DLU and LUT schemes. Given the fact lar versions of the X2C approach fall well into the class that two names, DLU and LUT, are in use in relativistic of approximate solutions that we discussed in section V quantum chemistry for the same idea and that this idea above. In Fig. 7, we demonstrate how relying only on is important in the more general context of SSUB we the one-electron contributions of Eq. (5), i.e., neglecting may propose to call it L-SSUB for local approximation all other potential-energy contributions to the Fock op- to SSUB. erator, affects the accuracy of the decoupling in general embedding schemes. In comparison to the relativistic case, we see that this approximation does not work very well. In the relativistic case, the eigenvectors are mainly 1. Fragmenting the Basis dominated by the one-electron contributions (because of the huge energy gap separating positive- and negative- energy states). In mean-field theory, this is not the case. The local construction of the block-diagonalization ma- The eigenvectors of the eigenvector matrix H and the trix Q is based on the partitioning of the basis B into k Fock matrix F are fundamentally different. For this rea- fragment bases BFi , such that son, the approximation fails when applied to the general k case, where no large energy gap sustains it. [ B = BFi , (105) i=1

with IX. APPROXIMATIONS FOR HUGE SINGLE-PARTICLE SPACES F F B i ∩ B j = ∅. (106)

Fi Fi We have mentioned in the beginning that we will not Each fragment basis B consists of N basis functions. Fi Fi address the issue of how to construct large Fock matri- It can be partitioned into BS and BE , both of which ces that can then be subjected to SSUB. However, given are subsets of BS and BE , respectively, that one can construct very large Fock matrices (con- BFi = BFi ∩ B , (107) sider, for example, those in semi-empirical methods ap- S S Fi Fi plied to molecular systems with thousands of atoms), the BE = B ∩ BE . (108) issue of diagonalizing them will be pressing. Whereas one may apply subspace diagonalization to such large The fragment basis should be chosen such that it cap- Fock matrices (e.g., Lanczos or Davidson) and construct tures the dominant interactions between subsystem and an approximate unitary matrix of reduced dimension, we environment. Quantum System Partitioning at the Single-Particle Level 17

O1 C2 H3 H4 O CH2 O1 C2 H3 H4 O1 C2 H3 H4

FIG. 7. The transformed Fock matrix F˜, eigenvector matrix C˜ , density matrix P˜ , and transformation matrix W (from left to right) of formaldehyde, obtained by decoupling the oxygen atom from the other atoms with a transformation matrix W constructed from the eigenvectors of the orthogonalized one-electron matrix H˘ , which is an approximation inspired by standard decoupling methods considering only the nuclear potential for the block diagonalization of Dirac-based Fock operators. A transformation based on the one-electron matrix H only is advantageous as it does not require a self-consistent solution. However, this approximate scheme does not yield block-diagonal matrices. The color code is as in Fig. 3.

2. Constructing the local block-diagonalization matrix be done with a rather trivial permutation. Therefore, in order to block-diagonalize the Fock matrix F with the local block-diagonalization matrix QF , we first need to F A fragment Fock matrix F i is constructed from a con- transform the Fock matrix F into the fragment basis rep- verged mean-field calculation employing the fragment ba- resentation with the K. Afterwards, F sis B i only. For this mean-field calculation, a modi- the block-diagonal Fock matrix F˜ is recovered in the orig- fied Fock operator may be applied, considering a subset inal basis B by an inverse permutation, of nuclei only, not taking those nuclei into account on T which the basis functions are not located. This fragment F˜ = KTQF KFKT QF  K. (112) Fock matrix is then block-diagonalized with the block- F diagonalization matrix Q i as outlined in section IV. If This implies that the total block-diagonalization matrix F a fragment basis B i consists exclusively of subsystem Q in the basis B can be constructed as, or environment basis functions, the matrix QFi becomes the identity matrix. Q = KTQF K, (113) With all fragment block-diagonalization matrices QFi , the local block-diagonalization matrix QF can be con- which block-diagonalizes the Fock matrix directly. structed in analogy to Eq. (104),

k M 3. Quantifying approximate block-diagonalizations QF = QFi . (109) i=1 Since the local block-diagonalization is only approxi- F However, this local block-diagonalization matrix Q can- mate, we need to introduce some sort of measure that not be applied to the Fock matrix F directly. allows us to analyze the quality of the resulting block- The matrix QF , is implicitly set up in a basis in which diagonalization. A naive approach to this is to simply in- the order of basis functions is permuted. Let Oˆ be an spect the off-diagonal blocks of the approximately block- 0 operator that reveals the order of a set in terms of its diagonal Fock matrix F˜ . However, such a measure would subsets. While the basis order of the basis B, in which depend on the nature of the particular system since the the Fock matrix F is represented, is given by magnitude of the matrix elements in the Fock matrix de- pend on the potentials in the Fock operator. Here, we OBˆ = [BS ,BE ] , (110) choose a diagonostic that has a more universal scope. It is based on a comparison between the eigenvectors of the the fragmented basis BF has the following order, approximately block-diagonalized Fock matrix F˜ 0 and the exact block-diagonalized Fock matrix F˜. I.e., the mea- F h F F F F F F i OBˆ = B 1 ,B 1 ,B 2 ,B 2 , ··· ,B k ,B k . (111) sure di quantifies the similarity of the eigenvectorsc ˜i and S E S E S E 0 c˜i,

While it seems to be complicated to convert the Fock ma- T 0 trix F to its fragment basis representation, it can actually di = c˜i c˜i , (114) Quantum System Partitioning at the Single-Particle Level 18 which is 1 for identical one-particle states. To quantify single-particle basis to define subsystems in a very gen- the adequateness of a block-diagonalization for the whole eral way, which we call subsystem separation by uni- system, we introduce the measure D, which is simply the tary block-diagonalization (SSUB). The approach re- average of all measures di over all eigenvectors, quires one calculation on the full system and as such it shares this disadvantage with embedding theories pro- NB posed by Manby and Miller and also with the relativistic −1 X D = NB di. (115) exact two-component approaches. However, the disad- i=1 vantage is made up for by exploiting computational effi- ciency for a subsystem in computational protocols that The closer this measure D is to 1, the higher the accuracy build upon a single-determinant solution. For instance, of the approximate block-diagonalization. these subsequent steps of a computational protocol may involve a four-index transformation from the atomic or- bital to a molecular orbital basis before a correlation cal- B. Example I: Structural separation culation is considered. The latter is then significantly simplified for the subsystem (and usually would not have been possible for the total system, which is one reason Results of the local SSUB scheme for the formaldehyde for adopting an embedding approach). molecule in which the oxygen atom is separated from the other atoms are shown in Fig. 8. The fragment basis for which the block-diagonalization matrix was determined consists solely of the basis functions centered on the oxy- gen and the carbon atom. The general nature of SSUB allows us to view the block- orthogonalization correction of EMFT from a more gen- While this block-diagonalization is not exact, the decou- eral perspective. Moreover, projector-based embedding pling scheme appears to decouple the most significant in- can be simplified significantly without compromising the teractions and yields a Fock matrix F˜ that for the most inherent exactness of the method in the limit for large part is block-diagonal. However, the associated eigen- energy-shift parameter µ. It also relates the field of em- vector matrix C˜ and density matrix P˜ show that there bedding to very different ones such as electron-positron are interactions that could not be decoupled by this ap- separation in relativistic quantum chemistry and purely proach. This observation is also supported by a relatively electronic partitionings such as core-valence separation. low measure D = 0.9222. For certain eigenvectors, the As such, SSUB can be viewed as a framework for study- measure d is even below 0.7, which indicates that the ing quantum system partitioning at the single-particle separation is not perfect. level in a general way.

C. Example II: Core-Valence Separation We emphasize that a system separation by SSUB allows one to proceed with further calculations on the sepa- Results of the local SSUB scheme for a core-valence sep- rated parts of the system. In particular, it is possible aration applied to the formaldehyde molecule, in which to consider more accurate and, hence, computationally the carbon and oxygen 1s-like molecular orbitals are sep- more expensive ab initio calculations for a subsystem in arated from all other orbitals, are shown in Fig. 9. The the same spirit as in projector embedding and X2C the- basis was split into three fragment bases. The first frag- ories. Apart from this, one may exploit SSUB to ac- ment basis consisted of the basis functions centered on celerate standard protocols (e.g., for the solution of the the oxygen atom, the second one of the basis functions SCF general eigenvalue problem). In future work, we will centered on the carbon atom. The remaining hydrogen therefore investigate a change of parameters for which the basis functions constitute the third fragment basis. In single-particle equations are formulated. In particular, this example, the transformed Fock matrix F˜ shows a for accelerating calculations on a subsystem, a change of perturbed block-diagonal structure. However, the rela- the nuclear framework — as it occurs in structure op- tively high measure D = 0.9998 implies that the sepa- timizations, first-principles molecular dynamics, and in- ration was in fact successful. teractive reactivity studies — represents such a change of parameters. This will require a thorough study of the transferability of the system-separating matrix transfor- X. CONCLUSIONS mations in order to assess when one may exploit a system separation, which has been obtained for some configura- tion, for related configurations. Moreover, a generaliza- In this work, we described a block-diagonalization ap- tion to time-dependent quantum dynamics should also proach directed toward an arbitrary separation of a be promising. Quantum System Partitioning at the Single-Particle Level 19

O1 C2 H3 H4 O CH2 O1 C2 H3 H4 O1 C2 H3 H4

FIG. 8. The transformed Fock matrix F˜, eigenvector matrix C˜ , density matrix P˜ , and transformation matrix W (from left to right) of formaldehyde, obtained by decoupling the oxygen atom from the other atoms with a block-diagonalization matrix Q obtained from a local approximation. The fragment on which the decoupling is based upon consists of only the oxygen and carbon atom. While the transformed Fock matrix F˜ appears to be block-diagonal, the eigenvector matrix C˜ and density matrix P˜ have off-diagonal contributions. The color code is as in Fig. 3.

1s O1val C2val H3 H4 1s val 1s O1val C2val H3 H4 1s O1val C2val H3 H4

FIG. 9. The transformed Fock matrix F˜, eigenvector matrix C˜ , density matrix P˜ , and transformation matrix W (from left to right) of formaldehyde, obtained by decoupling the 1s-like molecular orbitals of the carbon and oxygen atoms with a local approximation to the block-diagonalization matrix. The color code is as in Fig. 3.

ACKNOWLEDGMENTS This implies that subsystem and environment eigenvec- tors are orthogonal, leading to the expression,

˘ T ˘ We gratefully acknowledge financial support by ETH 0 = CE CS (A3) Zurich (Grant Number ETH-46 16-2). Part of this work ˘ T ˘ ˘ T ˘ was presented at the ICQC 2018 in Menton, France, in = C12C11 + C22C21 (A4) ˘ −T ˘ T ˘ ˘ −1 June 2018. We thank Professor Trond Saue for remarks = C22 C12 + C21C11 (A5) on the historical development of Jensen’s original block- T  ˘ ˘ −1 ˘ ˘ −1 diagonalization idea, mean-field spin–orbit Hamiltonians, = C12C22 + C21C11 , (A6) and on two orbital localization schemes related to SSUB. from which Eq. (A1) immediately follows.

Appendix A: Note concerning the construction of U Appendix B: Representation of the block-diagonalization matrix Q Here, we show that the equality in Eq. (57),

T ˘ ˘ −1  ˘ ˘ −1 107–110 −C21C11 = C12C22 , (A1) In the literature, there exist multiple representa- tions of the block-diagonalization matrix Q. These rep- holds by considering the orthogonality of the eigenvector resentations are very similar to the expression in Eq. (60) matrix C˘ , and it is trivial to show that they achieve the exact same block-diagonalization. The different representations arise ˘ T ˘ I = C C. (A2) from the following different definitions: Quantum System Partitioning at the Single-Particle Level 20

Si 1. The definition of the block-diagonalization as either We start by defining the localization function fj for each eigenvectorc ˘j for each subsystem Si, F˜ = QFQT, (B1) Pi t=1 NSt or Si X 2 fj = c˘j,r. (C1) T r=1+Pi−1 N F˜ = Q¯ FQ¯ , (B2) t=1 St

with Q¯ = QT. As in Eq. (51), this is simply a measure of the contribu- tion of the subsystem basis functions to the correspond- 2. The definition of the block-diagonalization matrix ing molecular orbital. After evaluation of the localization

Q as either functions, the NSi eigenvectors with the highest localiza- Si tion function fj are assigned to the subsystem Si. Q = QRQBD, (B3) However, this choice of eigenvectors suffers from the or aforementioned problems concerning the assignment of eigenvectors to multiple subsystems. This ambiguity can Q = QBDQR. (B4) be resolved by the following iterative scheme:

This is possible since QR and QBD commute. 1. For each pair of subsystems Sr and Ss, the assign- 3. The definition of the block-diagonalization matrix ment of eigenvectors is probed for a duplicate as- QBD as either signment. If no collision is detected, the procedure is completed. Otherwise, we continue with the next  I −UT Q = , (B5) step. BD UI 2. One of the detected collisions is chosen at random. or Letc ˘i be an eigenvector that was assigned to both  I −U¯  subsystems Sr and Ss, but with a larger contribu- Q = , (B6) BD U¯ T I tion to Sr than to Ss,

Sr Ss with U¯ = UT. Since it is also possible to swap the fi > fi . (C2) sign of both U¯ and U, this leaves us with four differ- ent definitions for the block-diagonalization matrix In this case, the eigenvectorc ˘i is removed from the choice of eigenvectors for subsystem S . QBD. s 3. Now that the subsystem S is missing an eigenvec- All of these different definitions may be combined, which s tor, we need to assign a new eigenvector to this leads to the large number of possible representations of subsystem for it to be considered complete. This the block-diagonalization matrix Q. new eigenvectorc ˘j is chosen as the eigenvector with Ss the highest localization function fj which has not yet been chosen for subsystem Ss. Appendix C: Eigenvector assignment for multiple subsystems 4. The procedure is repeated from the first step to probe for new collisions since the new choice of eigenvectors is not guaranteed to be free of mul- Assigning eigenvectors to multiple subsystems is not triv- tiple assignments. ial when considering multiple subsystems. When assign- ing an eigenvector to a subsystem with a localization function such as the one in Eq. (51), we must consider Unless there is a case in which two localization functions that this eigenvector may also contribute to another sub- for an eigenvectorc ˘i are equal, system. Ranking these contributions may prove to be Sr Ss hard and can lead to the assignment of a single eigen- fi = fi , (C3) vector to multiple subsystems. It is also possible that an eigenvector may not be chosen for any subsystem and this scheme is guaranteed to produce a unique and cor- is subsequently excluded from the block-diagonalization rect assignment of the eigenvectors to multiple subsys- procedure. This must not occur, and each eigenvector tems. However, it should also be noted that in the case can be assigned to exactly one subsystem only. Here, we of an assignment of an eigenvector to multiple subsys- introduce a scheme for an optimal and unique assignment tems, caution is advised since it may be an indication of of a set of eigenvectors to multiple subsystems. a particularly challenging partitioning. Quantum System Partitioning at the Single-Particle Level 21

REFERENCES 39R. G. Gordon and Y. S. Kim, J. Chem. Phys. 56, 3122 (1972). 40Y. S. Kim and R. G. Gordon, J. Chem. Phys. 60, 1842 (1974). 41G. Senatore and K. Subbaswamy, Phys. Rev. B 34, 5754 (1986). 1 C. Cheng, J. Hussels, M. Niu, H. L. Bethlem, K. S. E. 42P. Cortona, Phys. Rev. B 44, 8454 (1991). Eikema, E. J. Salumbides, W. Ubachs, M. Beyer, N. J. 43T. A. Wesolowski and A. Warshel, J. Phys. Chem. 97, 8050 Holsch¨ , J. A. Agner, F. Merkt, L.-G. Tao, S.-M. Hu, and (1993). C. Jungen, Phys. Rev. Lett. 121, 013001 (2018). 44 and , Int. J. Quantum Chem. 2 P. Cortona A. V. Monteleone L. M. Wang and Z.-C. Yan, Phys. Rev. A 97, 060501 (2018). 52, 987 (1994). 3 M. Puchalski, A. Spyszkiewicz, J. Komasa, and 45J. Neugebauer, M. J. Louwerse, E. J. Baerends, and T. A. K. Pachucki, Phys. Rev. Lett. 121, 073001 (2018). , J. Chem. Phys. 122, 094115 (2005). 4 Wesolowski V. Barone, M. Biczysko, and C. Puzzarini, Acc. Chem. Res. 46M. Iannuzzi, B. Kirchner, and J. Hutter, Chem. Phys. Lett. 48, 1413 (2015). 421, 16 (2006). 5 C. Puzzarini and V. Barone, Acc. Chem. Res. 51, 548 (2018). 47 , , and , J. Comput. 6 C. R. Jacob J. Neugebauer L. Visscher D. R. Glowacki, C.-H. Liang, C. Morley, M. J. Pilling, and Chem. 29, 1011 (2008). S. H. Robertson, J. Phys. Chem. A 116, 9545 (2012). 48 , , , , and 7 S. Fux C. R. Jacob J. Neugebauer L. Visscher A. Warshel and M. Levitt, J. Mol. Biol. 103, 227 (1976). , J. Chem. Phys. 132, 164101 (2010). 8 M. Reiher U. C. Singh and P. A. Kollman, J. Comput. Chem. 7, 718 49P. Elliott, K. Burke, M. H. Cohen, and A. Wasserman, (1986). Phys. Rev. A 82, 024501 (2010). 9 M. J. Field, P. A. Bash, and M. Karplus, J. Comput. Chem. 50J. D. Goodpaster, N. Ananth, F. R. Manby, and T. F. 11, 700 (1990). , J. Chem. Phys. 133, 084103 (2010). 10 Miller H. Lin and D. G. Truhlar, Theor. Chem. Acc. 117, 185 (2006). 51 and , WIREs Comput Mol Sci 4, 11 C. R. Jacob J. Neugebauer H. M. Senn and W. Thiel, Curr. Opin. Chem. Biol. 11, 182 325 (2014). (2007). 52 , , , , and 12 M. E. Fornace J. Lee K. Miyamoto F. R. Manby H. M. Senn and W. Thiel, Top. Curr. Chem. 268, 173 (2007). , J. Chem. Theory Comput. 11, 568 (2015). 13 T. F. Miller H. M. Senn and W. Thiel, Angew. Chem. Int. Ed. 48, 1198 53T. A. Wesolowski, S. Shedge, and X. Zhou, Chem. Rev. 115, (2009). 5891 (2015). 14 J. M. Olsen, K. Aidas, and J. Kongsted, J. Chem. Theory 54F. Ding, F. R. Manby, and T. F. Miller, J. Chem. Theory Comput. 6, 3721 (2010). Comput. 13, 1605 (2017). 15 J. M. H. Olsen and J. Kongsted, Adv. Quantum Chem. 61, 55S. Huzinaga and A. A. Cantu, J. Chem. Phys. 55, 5543 107 (2011). (1971). 16 K. Sneskov, T. Schwabe, O. Christiansen, and J. Kongsted, 56N. Govind, Y. A. Wang, and E. A. Carter, J. Chem. Phys. Phys. Chem. Chem. Phys. 13, 18551 (2011). 110, 7677 (1999). 17 M. Svensson, S. Humbel, R. D. J. Froese, T. Matsubara, 57T. Kluner¨ , N. Govind, Y. A. Wang, and E. A. Carter, J. S. Sieber, and K. Morokuma, J. Phys. Chem. 100, 19357 Chem. Phys. 116, 42 (2002). (1996). 58 and , J. Chem. Phys. 125, 084102 18 P. Huang E. A. Carter S. Dapprich, I. Komairomi´ , K. S. Byun, K. Morokuma, and (2006). M. J. Frisch, J. Mol. Struct. THEOCHEM 461-462, 1 (1999). 59 , , and , Phys. Chem. 19 A. S. P. Gomes C. R. Jacob L. Visscher H. Nakai, Chem. Phys. Lett. 363, 73 (2002). Chem. Phys. 10, 5353 (2008). 20 L. V. Slipchenko and M. S. Gordon, J. Comput. Chem. 28, 60C. Huang, M. Pavone, and E. A. Carter, J. Chem. Phys. 276 (2006). 134, 154110 (2011). 21 M. S. Gordon, L. Slipchenko, H. Li, and J. H. Jensen, 3, 61F. R. Manby, M. Stella, J. D. Goodpaster, and T. F. 177 (2007). , J. Chem. Theory Comput. 8, 2564 (2012). 22 Miller T. Akama, M. Kobayashi, and H. Nakai, J. Comput. Chem. 62S. Hofener¨ and L. Visscher, J. Chem. Phys. 137, 204120 28, 2003 (2007). (2012). 23 M. Kobayashi and H. Nakai, Int. J. Quantum Chem. 109, 63J. D. Goodpaster, T. A. Barnes, F. R. Manby, and T. F. 2227 (2009). , J. Chem. Phys. 140, 18A507 (2014). 24 Miller III M. S. Gordon, D. G. Fedorov, S. R. Pruitt, and L. V. 64C. Daday, C. Konig¨ , J. Neugebauer, and C. Filippi, Slipchenko, Chem. Rev. 112, 632 (2012). ChemPhysChem 15, 3205 (2014). 25 H.-P. Breuer and F. Petruccione, The theory of open quan- 65T. Dresselhaus, J. Neugebauer, S. Knecht, S. Keller, tum systems, Oxford University Press, 2002. , and , J. Chem. Phys. 142, 044111 (2015). 26 Y. Ma M. Reiher A. Amann and U. Muller-Herold¨ , Offene Quantensysteme, 66B. Hegely´ , P. R. Nagy, G. G. Ferenczy, and M. Kallay´ , J. Springer-Verlag Berlin Heidelberg, 2011. Chem. Phys. 145, 064107 (2016). 27 S. R. White, Phys. Rev. Lett. 69, 2863 (1992). 67 , , and , Phys. Rev. B 91, 28 A. A. Kananenka E. Gull D. Zgid S. R. White, Phys. Rev. B 48, 10345 (1993). 121111 (2015). 29 E. Schmidt, Math. Ann. 63, 433 (1907). 68 , , and , J. Chem. Phys. 30 T. N. Lan A. A. Kananenka D. Zgid U. Schollwock¨ , Ann. Phys. 326, 96 (2011). 143, 241102 (2015). 31 G. Knizia and G. K.-L. Chan, Phys. Rev. Lett. 109, 186404 69T. N. Lan and D. Zgid, J. Phys. Chem. Lett. 8, 2200 (2017). (2012). 70 and , Phys. Rev. B 45, 6479 (1992). 32 A. Georges G. Kotliar G. Knizia and G. K.-L. Chan, J. Chem. Theory Comput. 9, 71A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg, 1428 (2013). Rev. Mod. Phys. 68, 13 (1996). 33 I. W. Bulik, G. E. Scuseria, and J. Dukelsky, Phys. Rev. B 72A. Georges, AIP Conf. Proc 715, 3 (2004). 89, 035140 (2014). 73 , , , , 34 G. Kotliar S. Y. Savrasov K. Haule V. S. Oudovenko I. W. Bulik, W. Chen, and G. E. Scuseria, J. Chem. Phys. O. Parcollet, and C. A. Marianetti, Rev. Mod. Phys. 78, 141, 054113 (2014). 865 (2006). 35 M. Welborn, T. Tsuchimochi, and T. Van Voorhis, J. Chem. 74E. Fromager, Mol. Phys. 113, 419 (2015). Phys. 145, 074102 (2016). 75 , , , and , 36 B. Senjean M. Tsuchiizu V. Robert E. Fromager N. Ricke, M. Welborn, H.-Z. Ye, and T. Van Voorhis, Mol. Mol. Phys. 115, 48 (2017). Phys. 115, 2242 (2017). 76 , , , and , 37 B. Senjean N. Nakatani M. Tsuchiizu E. Fromager P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964). Phys. Rev. B 97, 235105 (2018). 38W. Kohn and L. J. Sham, Phys. Rev. 140, A1133 (1965). Quantum System Partitioning at the Single-Particle Level 22

77B. O. Roos, P. R. Taylor, and P. E. M. Siegbahn, Chem. 111M. Zi lkowski, B. Jans´ık, P. Jørgensen, and J. Olsen, J. Phys. 48, 157 (1980). Chem. Phys. 131, 124112 (2009). 78B. O. Roos, Int. J. Quantum Chem. 18, 175 (1980). 112Z. Li, H. Li, B. Suo, and W. Liu, Acc. Chem. Res. 47, 2758 79K. Ruedenberg, M. W. Schmidt, M. M. Gilbert, and S. T. (2014). Elbert, Chem. Phys. 71, 41 (1982). 113J. P. Perdew, Phys. Rev. B 33, 8822 (1986). 80J. Olsen, B. O. Roos, P. Jørgensen, and H. J. A. Jensen, J. 114A. D. Becke, Phys. Rev. A 38, 3098 (1988). Chem. Phys. 89, 2185 (1988). 115M. Reiher and A. Wolf, Relativistic Quantum Chemistry: 81T. Fleig, J. Olsen, and L. Visscher, J. Chem. Phys. 119, The Fundamental Theory of Molecular Science, Wiley-VCH, 2963 (2003). Weinheim, 2nd edition, 2015. 82J. Ivanic, J. Chem. Phys. 119, 9364 (2003). 116J.-L. Heully, I. Lindgren, E. Lindroth, S. Lundquist, and 83S. M. Parker, T. Seideman, M. A. Ratner, and T. Shiozaki, A.-M. M˚artensen-Pendrill, J. Phys. B: At. Mol. Phys. 19, J. Chem. Phys. 139, 021108 (2013). 2799 (1986). 84S. M. Parker and T. Shiozaki, J. Chem. Theory Comput. 10, 117H. J. A. Jensen, Talk on conference on relativistic effects in 3738 (2014). heavy elements — REHE, April 2005, M¨ulheim,2005. 85 S. M. Parker and T. Shiozaki, J. Chem. Phys. 141, 211102 118B. A. Hess, Phys. Rev. A 33, 3742 (1986). (2014). 119A. Wolf, M. Reiher, and B. A. Hess, J. Chem. Phys. 117, 86 W. Thiel, Tetrahedron 44, 7393 (1988). 9215 (2002). 87 M. J. S. Dewar, Int. J. Quantum Chem. 44, 427 (1992). 120M. Reiher and A. Wolf, J. Chem. Phys. 121, 2037 (2004). 88 T. Clark, Semiempirical Molecular Orbital Theory: Facts, 121M. Barysz, A. J. Sadlej, and J. G. Snijders, Int. J. Quantum Myths and Legends, pp. 369–380, Springer, Dordrecht, 1993. Chem. 65, 225 (1997). 89 W. Thiel, Adv. Chem. Phys. 93, 703 (1996). 122M. Barysz and A. J. Sadlej, J. Mol. Struct. (THEOCHEM) 90 W. Thiel, Thermochemistry from Semiempirical Molecular Or- 573, 181 (2001). bital Theory, chapter 8, pp. 142–161, American Chemical Soci- 123M. Barysz and A. J. Sadlej, J. Chem. Phys. 116, 2696 (2002). ety: Washington, DC, 1998. 124M. Filatov and D. Cremer, J. Chem. Phys. 119, 11526 (2003). 91 T. Clark, J. Mol. Struct. THEOCHEM 530, 1 (2000). 125M. Filatov and K. G. Dyall, Theor. Chem. Acc. 117, 333 92 T. Bredow and K. Jug, Theor. Chem. Acc. 113, 1 (2005). (2007). 93 J. J. P. Stewart, Rev. Comput. Chem. 1, 45 (2007). 126M. Reiher and A. Wolf, J. Chem. Phys. 121, 10945 (2004). 94 E. G. Lewars, : Introduction to the 127M. Reiher, Theor. Chem. Acc. 116, 241 (2006). theory and applications of molecular and quantum mechanics, 128W. Kutzelnigg and W. Liu, J. Chem. Phys. 123, 241102 Springer Science & Business Media, 2010. (2005). 95 T. Clark and J. J. P. Stewart, MNDO-Like Semiempirical 129M. Iliaˇs and T. Saue, J. Chem. Phys. 126, 064102 (2007). Molecular Orbital Theory and Its Application to Large Systems, 130T. Saue, ChemPhysChem 12, 3077 (2011). chapter 8, pp. 259–286, John Wiley & Sons: New York, 2011. 131K. G. Dyall, J. Chem. Phys. 106, 9618 (1997). 96 W. Thiel, WIREs Comput Mol Sci 4, 145 (2014). 132W. Liu and D. Peng, J. Chem. Phys. 125, 044102 (2006). 97 T. Bredow and K. Jug, Semiempirical Molecular Orbital 133W. Kutzelnigg and W. Liu, Mol. Phys. 104, 2225 (2006). Methods, chapter 6, pp. 159–202, American Cancer Society, 134W. Liu and W. Kutzelnigg, J. Chem. Phys. 126, 114107 2017. (2007). 98 T. Husch, A. C. Vaucher, and M. Reiher, Int. J. Quantum 135D. Peng and M. Reiher, Theor. Chem. Acc. 131, 1081 (2012). Chem. 118, e25799 (2018). 136L. Visscher, Theor. Chem. Acc. 98, 68 (1997). 99 C. C. J. Roothaan, Rev. Mod. Phys. 23, 69 (1951). 137M. H. Mittleman, Phys. Rev. A 5, 2395 (1972). 100 G. G. Hall, Proc. R. Soc. Lond. A 205, 541 (1951). 138M. H. Mittleman, Phys. Rev. A 24, 1167 (1981). 101 Q. Sun, T. C. Berkelbach, N. S. Blunt, G. H. Booth, 139J. Sucher, Phys. Rev. A 22, 348 (1980). S. Guo, Z. Li, J. Liu, J. D. McClain, E. R. Sayfutyarova, 140J. Sucher, Phys. Scr. 36, 271 (1987). S. Sharma, S. Wouters, and G. K.-L. Chan, WIREs Comput 141B. A. Heß, C. M. Marian, U. Wahlgren, and O. Gropen, Mol Sci 8, e1340 (2018). Chem. Phys. Lett. 251, 365 (1996). 102 F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys. 7, 142B. Schimmelpfennig, L. Maron, U. Wahlgren, C. Teichteil, 3297 (2005). H. Fagerli, and O. Gropen, Chem. Phys. Lett. 286, 261 103 P.-O. Lowdin¨ , Adv. Quantum Chem. 5, 185 (1970). (1998). 104 W. Givens, SIAM J. Appl. Math. 6, 26 (1958). 143J. Autschbach, D. Peng, and M. Reiher, J. Chem. Theory 105 C. G. J. Jacobi, Crelle’s Journal 30, 51 (1846). Comput. 8, 4239 (2012). 106 M. Robbe´ and M. Sadkane, BIT Numer. Math. 45, 181 (2005). 144J. E. Peralta and G. E. Scuseria, J. Chem. Phys. 120, 5875 107 L. S. Cederbaum, J. Schirmer, and H.-D. Meyer, J. Phys. A (2004). 22, 2427 (1989). 145J. E. Peralta, J. Uddin, and G. E. Scuseria, J. Chem. Phys. 108 J. Sikkema, L. Visscher, T. Saue, and M. Iliaˇs, J. Chem. 122, 084108 (2005). Phys. 131, 124116 (2009). 146J. Thar and B. Kirchner, J. Chem. Phys. 130, 124103 (2009). 109D. Peng and M. Reiher, J. Chem. Phys. 136, 244108 (2012). 110J. Seino and H. Nakai, J. Chem. Phys. 136, 244102 (2012).