Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Exploring Ground and Excited States via Single Reference Coupled-Cluster Theory and Algebraic Geometry

Svala Sverrisdóttir Department of Mathematics, The University of California, Berkeley, CA, USA    Fabian M. Faulstich Department of Mathematics, Rensselaer Polytechnic Institute, Troy, NY, USA faulsf@rpi.edu
Abstract

The exploration of the root structure of coupled cluster equations holds both foundational and practical significance for computational quantum chemistry. This study provides insight into the intricate root structures of these non-linear equations at both the CCD and CCSD level of theory. We utilize computational techniques from algebraic geometry, specifically the monodromy and parametric homotopy continuation methods, to calculate the full solution set. We compare the computed CC roots against various established theoretical upper bounds, shedding light on the accuracy and efficiency of these bounds. We hereby focus on the dissociation processes of four-electron systems such as (H2)2 in both D2h and D∞h configurations, H4 symmetrically distorted on a circle, and lithium hydride. We moreover investigate the ability of single-reference coupled cluster solutions to approximate excited state energies. We find that multiple CC roots describe energies of excited states with high accuracy. Remarkably, our investigations reveal that for systems like lithium hydride, CC not only provides high-accuracy approximations to several excited state energies but also to the states themselves.

{tocentry}[Uncaptioned image]

1 Introduction

Coupled cluster (CC) theory is a high-precision wave function method in computational quantum chemistry 1. Its origins date back to 1958 when Coester proposed to use an exponential parametrization of the wave function 2, i.e.

|Ψ=eT^|Φ0,ketΨsuperscript𝑒^𝑇ketsubscriptΦ0|{\Psi}\rangle=e^{\hat{T}}|{\Phi_{0}}\rangle,| roman_Ψ ⟩ = italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ , (1)

where T^^𝑇\hat{T}over^ start_ARG italic_T end_ARG is the new unknown, and |Φ0ketsubscriptΦ0|{\Phi_{0}}\rangle| roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ is a chosen reference Slater determinant (commonly the Hartree-Fock state). This parametrization was independently derived by Hubbard 3 and Hugenholtz 4 in 1957 as an alternative to summing many-body perturbation theory contributions order by order. A pivotal moment in the development of CC theory was Čížek’s seminal work in 1966 5. Čížek presented the first derivation of the CC working equations together with corresponding simulations. This work established CC theory within the theoretical framework that we recognize today, including second quantization applied to many-fermion systems, normal ordering, contractions, Wick’s theorem, normal-ordered Hamiltonians and Goldstone-style diagrammatic techniques. There exist several reviews of CC theory, summarizing different aspects of the method’s success over the past decades, see e.g. Kümmel 6, Čížek 7, Bartlett 8, Paldus 9, Arponen 10 and Bishop 11.

Employing the exponential ansatz in Eq. (1) yields the following reformulation of the stationary Schrödinger equation:

{Φ0|eT^H^eT^|Φ0=E,Φ|eT^H^eT^|Φ0=0|Φ|Φ0.\left\{\begin{aligned} \langle{\Phi_{0}}|e^{-\hat{T}}\hat{H}e^{\hat{T}}|{\Phi_% {0}}\rangle&=E,\\ \langle{\Phi}|e^{-\hat{T}}\hat{H}e^{\hat{T}}|{\Phi_{0}}\rangle&=0&\forall|{% \Phi}\rangle\perp|{\Phi_{0}}\rangle.\\ \end{aligned}\right.{ start_ROW start_CELL ⟨ roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | italic_e start_POSTSUPERSCRIPT - over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT over^ start_ARG italic_H end_ARG italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL = italic_E , end_CELL end_ROW start_ROW start_CELL ⟨ roman_Φ | italic_e start_POSTSUPERSCRIPT - over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT over^ start_ARG italic_H end_ARG italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL = 0 end_CELL start_CELL ∀ | roman_Φ ⟩ ⟂ | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ . end_CELL end_ROW (2)

Employing the standard Galerkin projection, the latter condition yields the projected CC equations (vide infra), i.e. a set of nonlinear algebraic equations. Typically, a single solution to this set of equations is computed using (quasi) Newton-type methods 12. However, as a set of nonlinear algebraic equations, the CC equations possess multiple roots, raising the following natural questions:

  • i)

    How many solutions to the CC equations exist?

  • ii)

    Which solution describes the ground state best?

  • iii)

    Do some of the solutions describe excited states?

Especially the last question is controversial, to say the least. On one side, it can be proven that as long a state is not strictly orthogonal to the chosen reference state, CC theory can – in its untruncated form – describe that state exactly. On the other side, excited states commonly have poor overlap with the Hartree-Fock ground state reference – hindering the convergence of (quasi) Newton-type methods. Besides such theoretical arguments, it has been numerically shown that higher-order CC solutions can approximate excited state energies well while the CC solutions provide poor approximations to the actual states 13. Conventionally, excited states are computed in a post-single-reference CC manner. This is done by either constructing and diagonalizing the similarity-transformed Hamiltonian in equation-of-motion CC 14, 15, 16, 17, 18, 19, 20 or by examining the poles of the linear-response function in linear-response CC 16, 21, 22, 23. In either case, the excited states provided by these methods are inherently biased toward the lowest eigenstate upon which they are built. As a direct consequence, the accuracy of such approaches is known to be quite poor for certain classes of eigenstates that are energetically higher than the ground state 24, 25, 26, 27, 28, 29.

Besides the fundamental nature of questions i)–iii), the existence of multiple roots to the CC equations can present challenges for the practical implementations of the method. For example, the roots may be difficult to adequately converge to 5, 30, 31, 32, and the convergence properties of the iterative solution methods can strongly depend on the employed initial guess 33, 34. See also Ref. 35 for a mathematical exposition of this problem and Ref. 36 for a potential remedy.

In this work, we elaborate on recent advances coming from algebraic geometry that provide new ideas on how to bound the number of roots to the CC equations. These new and significantly improved bounds can then be used to compute all roots to the CC equations, answering the above-posed questions i) – iii). In particular, we demonstrate that higher-order solutions of the CC equations can effectively approximate not only the energies of excited states but also the eigenstates themselves.

1.1 Previous works and perspective

We highlight that the inaugural investigation of the root structure of CC methods dates back to 1978 when Zivkovic and Monkhorst scrutinized the singularities and multifarious solutions in single-reference CC equations 37. Building upon this work, Paldus and colleagues subsequently conducted mathematical and numerical analyses during the early 1990s, focusing on the manifold of solutions of both single-reference and state-universal multi-reference CC equations, while also elucidating their singularities and analytical characteristics 38, 39. In 1998, Kowalski and Jankowski revived homotopy methods in connection with the single-reference CC theory and used them to solve a CCD system 40. Kowalski with Piecuch then extended the application of homotopy methods to the CCSD, CCSDT, and CCSDTQ 41 for a 4-electron system described in a minimal basis set. From these investigations, they derived the β𝛽\betaitalic_β-nested equations and proved the Fundamental Theorem of the β𝛽\betaitalic_β-NE Formalism 42, which enabled an explanation of curves connecting multiple solutions of the various CC polynomial systems, i.e. from CCSD to CCSDT, CCSDT to CCSDTQ, etc. This Kowalski-Piecuch homotopy was recently analyzed in-depth using topological degree theory 43, 44. Moreover, Kowalski with Piecuch used homotopy methods to obtain all solutions of the generalized Bloch equation 45 and explain symmetry breaking processes in the CC formulation 46, 47, 48, 49, 50. It is noteworthy that these fundamental results of CC theory have spurred the development of the method of moment CC approach 51, 52, 53, 54, which in turn has led to the highly efficient completely renormalized CC approach 55, 56, 57. This emphasizes the pivotal role of this research trajectory in driving methodological innovations within CC theory. A perspective on the recent use of homotopy methods in the context of quantum chemistry can be found in Ref. 58.

A more algebraic and mathematically driven approach to this problem has recently gained momentum. Initialized by some of the authors, a first algebraic perspective was introduced in Ref. 59. This was followed by an in-depth algebraic investigation leading to the introduction of the CC truncation varieties 60 – which are the subject of this work. This work moreover contained open mathematical questions and conjectures, one of which was proven shortly after 61. More generally, these works are part of a presently increasing interest in the mathematical understanding of CC theory, driven by the applied mathematics community. In this context, Schneider performed the first local analysis of CC theory in 2009 employing a functional analytic framework 62. Subsequently, this approach was refined 63, 64 and applied to different CC methods 65, 66, 67. A more versatile framework for analyzing general CC variants using topological degree theory was later proposed by Csirik and Laestadius 43, 44. The most recent numerical analysis results regarding single reference CC were established by Hassan, Maday, and Wang 68, 69.

2 Coupled-cluster theory

The starting point of CC theory is the exponential ansatz in Eq. (1), leading to the CC energy expression

E=Φ0|eT^H^eT^|Φ0,𝐸quantum-operator-productsubscriptΦ0superscript𝑒^𝑇^𝐻superscript𝑒^𝑇subscriptΦ0E=\langle{\Phi_{0}}|e^{-\hat{T}}\hat{H}e^{\hat{T}}|{\Phi_{0}}\rangle,italic_E = ⟨ roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | italic_e start_POSTSUPERSCRIPT - over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT over^ start_ARG italic_H end_ARG italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ , (3)

where the cluster operator T^^𝑇\hat{T}over^ start_ARG italic_T end_ARG is determined through the CC equations

0=Φ|eT^H^eT^|Φ0,|Φ|Φ0.formulae-sequence0quantum-operator-productΦsuperscript𝑒^𝑇^𝐻superscript𝑒^𝑇subscriptΦ0perpendicular-tofor-allketΦketsubscriptΦ00=\langle{\Phi}|e^{-\hat{T}}\hat{H}e^{\hat{T}}|{\Phi_{0}}\rangle,\qquad\forall% |{\Phi}\rangle\perp|{\Phi_{0}}\rangle.0 = ⟨ roman_Φ | italic_e start_POSTSUPERSCRIPT - over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT over^ start_ARG italic_H end_ARG italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ , ∀ | roman_Φ ⟩ ⟂ | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ . (4)

The cluster operator is a linear combination of elementary particle-hole excitation operators, i.e.

T^(𝐭)=μ>0tμX^μ,^𝑇𝐭subscript𝜇0subscript𝑡𝜇subscript^𝑋𝜇\hat{T}({\bf t})=\sum_{\mu>0}t_{\mu}\hat{X}_{\mu},over^ start_ARG italic_T end_ARG ( bold_t ) = ∑ start_POSTSUBSCRIPT italic_μ > 0 end_POSTSUBSCRIPT italic_t start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , (5)

where we highlight the dependence of T^^𝑇\hat{T}over^ start_ARG italic_T end_ARG on its expansion coefficients by writing T^(𝐭)^𝑇𝐭\hat{T}(\bf t)over^ start_ARG italic_T end_ARG ( bold_t ). Recall that elementary particle-hole excitation operators X^μsubscript^𝑋𝜇\hat{X}_{\mu}over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT are merely compositions of projections onto a subset of single-particle basis elements (the occupied orbitals) and wedge products with another subset of single-particle basis elements (the virtual orbital). It is therefore convenient to label the elementary particle-hole excitation operators using multi-indices μ>0𝜇0\mu>0italic_μ > 0 that clarify the projections and wedge products involved 12, 62. Therefore, acting on the reference state |Φ0ketsubscriptΦ0|{\Phi_{0}}\rangle| roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩, the elementary particle-hole excitation operators define the Hilbert space of functions that are L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-orthogonal to |Φ0ketsubscriptΦ0|{\Phi_{0}}\rangle| roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩. We henceforth set 𝒱:={|Φ0}assign𝒱superscriptketsubscriptΦ0perpendicular-to\mathcal{V}:=\{|{\Phi_{0}}\rangle\}^{\perp}caligraphic_V := { | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ } start_POSTSUPERSCRIPT ⟂ end_POSTSUPERSCRIPT and note that T^(𝐭)|Φ0𝒱^𝑇𝐭ketsubscriptΦ0𝒱\hat{T}({\bf t})|{\Phi_{0}}\rangle\in\mathcal{V}over^ start_ARG italic_T end_ARG ( bold_t ) | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ ∈ caligraphic_V for all CC amplitudes 𝐭𝐭\mathbf{t}bold_t. Hence, we can express Eq. (4) in the more common form as a set of projective equations, i.e.

0=Φμ|eT^(𝐭)H^eT^(𝐭)|Φ0,μ>0,formulae-sequence0quantum-operator-productsubscriptΦ𝜇superscript𝑒^𝑇𝐭^𝐻superscript𝑒^𝑇𝐭subscriptΦ0for-all𝜇00=\langle\Phi_{\mu}|e^{-\hat{T}({\bf t})}\hat{H}e^{\hat{T}({\bf t})}|\Phi_{0}% \rangle,\quad\forall\mu>0,0 = ⟨ roman_Φ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT | italic_e start_POSTSUPERSCRIPT - over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT over^ start_ARG italic_H end_ARG italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ , ∀ italic_μ > 0 , (6)

where |Φμ=X^μ|Φ0ketsubscriptΦ𝜇subscript^𝑋𝜇ketsubscriptΦ0|{\Phi_{\mu}}\rangle=\hat{X}_{\mu}|{\Phi_{0}}\rangle| roman_Φ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ⟩ = over^ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ are the excited Slater determinants spanning 𝒱𝒱\mathcal{V}caligraphic_V. In practice, the expansion in Eq. (5) is commonly limited to e.g. only containing one and two particle-hole excitations leading to CCSD. In this case, the projective equations (6) are also restricted to the corresponding set of excited Slater determinants, yielding a square system of polynomial equations that we seek to solve. Hence, a key object in CC theory is the set

𝒮={𝐭𝔽K|Φμ|eT^(𝐭)H^eT^(𝐭)|Φ0=0,μ>0},𝒮conditional-set𝐭superscript𝔽𝐾formulae-sequencequantum-operator-productsubscriptΦ𝜇superscript𝑒^𝑇𝐭^𝐻superscript𝑒^𝑇𝐭subscriptΦ00for-all𝜇0\mathcal{S}=\{{\bf t}\in\mathbb{F}^{K}~{}|~{}\langle\Phi_{\mu}|e^{-\hat{T}({% \bf t})}\hat{H}e^{\hat{T}({\bf t})}|\Phi_{0}\rangle=0,\quad\forall\mu>0\},caligraphic_S = { bold_t ∈ blackboard_F start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT | ⟨ roman_Φ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT | italic_e start_POSTSUPERSCRIPT - over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT over^ start_ARG italic_H end_ARG italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ = 0 , ∀ italic_μ > 0 } , (7)

where 𝔽𝔽\mathbb{F}blackboard_F is the considered number field (either \mathbb{R}blackboard_R or \mathbb{C}blackboard_C) and K𝐾Kitalic_K is the “system size” (determined by the number of correlated electrons, the size of one-particle basis functions as well as further selection rules). Note that 𝔽𝔽\mathbb{F}blackboard_F can be \mathbb{R}blackboard_R or \mathbb{C}blackboard_C depending on whether we are seeking real- or complex-valued CC amplitudes. We emphasize that although the CC polynomial coefficients are real, the roots of the polynomial system may not be. Mathematically, the theory describing complex-valued solutions is simpler and more complete than the theory describing real-valued solutions. Therefore, we shall consider 𝔽=𝔽\mathbb{F}=\mathbb{C}blackboard_F = blackboard_C for mathematical considerations, however, solving the truncated CC equations over 𝔽=𝔽\mathbb{F}=\mathbb{C}blackboard_F = blackboard_C may yield unwanted complex valued energies.

Mathematically, the set 𝒮𝒮\mathcal{S}caligraphic_S is an (algebraic) variety. Since the concept of varieties may be unconventional within the computational chemistry community, we provide an illustrative example. A variety is a mathematical object defined by the zeros of a set of polynomial equations. Consider the multivariate polynomial system

{p1(x,y,z)=x2+y2z2,p2(x,y,z)=x+y2+z1.\left\{\begin{aligned} p_{1}(x,y,z)&=x^{2}+y^{2}-z^{2},\\ p_{2}(x,y,z)&=\frac{x+y}{\sqrt{2}}+z-1.\end{aligned}\right.{ start_ROW start_CELL italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) end_CELL start_CELL = italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) end_CELL start_CELL = divide start_ARG italic_x + italic_y end_ARG start_ARG square-root start_ARG 2 end_ARG end_ARG + italic_z - 1 . end_CELL end_ROW (8)

The zeros of p1subscript𝑝1p_{1}italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT form a cone, whereas the zeros of p2subscript𝑝2p_{2}italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT form a plane; the set of simultaneous roots defines a variety that corresponds to a parabolic curve, see Fig. 1. As a curve, this variety is one-dimensional, embedded in the ambient space which is three-dimensional in this case. A central quality for this work is the degree of a variety which is defined to be the maximal number of intersection points of the variety with a general linear space of complementary dimension. By requiring the linear space to be general we ensure the intersection is always finite. For the variety in Fig. 1 we see that intersecting the yellow curve with a plane (a linear space of complimentary dimension) yields at most two intersection points, hence, its degree is two. We add that over the complex field, the condition of “maximal number of intersection points” is redundant for the definition of degree since the number of intersection points is the same for generic complex spaces of complementary dimension.

Refer to caption
Figure 1: Root structure corresponding to Eq. (8). The cone describes the roots of p1subscript𝑝1p_{1}italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and the plane describes the roots of p2subscript𝑝2p_{2}italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The simultaneous roots are described by the yellow curve that defines the solution variety of the multivariate polynomial system.

We conclude this section by noting that in the case of untruncated CC, amplitudes that solve the full CC equations describe eigenstates of the Hamiltonian. Therefore, the cardinality of 𝒮𝒮\mathcal{S}caligraphic_S is exactly the number of eigenstates of H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG that are intermediately normalized and the number of roots to the CC equations is bounded from above by the number of Slater determinants. However, when truncations are imposed, the cardinality of 𝒮𝒮\mathcal{S}caligraphic_S increases, i.e. truncated CC theory yields (some) unphysical solutions. Therefore, it becomes less clear what the different elements in 𝒮𝒮\mathcal{S}caligraphic_S describe. Understanding the variety 𝒮𝒮\mathcal{S}caligraphic_S and characterizing (some of) its elements are the subject of this manuscript.

3 Truncation varieties

In CC theory one encounters various algebraic equations that define different varieties. The truncation varieties 60 arise from the exponential parametrization, i.e.

exp:𝕍𝒱,𝐭eT^(𝐭)|Φ0,:expformulae-sequence𝕍𝒱maps-to𝐭superscript𝑒^𝑇𝐭ketsubscriptΦ0{\rm exp}:\mathbb{V}\to\mathcal{V},~{}\mathbf{t}\mapsto e^{\hat{T}(\mathbf{t})% }|\Phi_{0}\rangle,roman_exp : blackboard_V → caligraphic_V , bold_t ↦ italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT | roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟩ , (9)

where 𝕍𝕍\mathbb{V}blackboard_V is the CC amplitude space. In the untruncated CC method, the exponential map is bijective. However, when truncations are introduced, this property no longer holds. Truncations correspond to restrictions of the exponential map to a subspace, e.g. for CCSD the exponential map is restricted to 𝕍{1,2}𝕍subscript𝕍12𝕍\mathbb{V}_{\{1,2\}}\subset\mathbb{V}blackboard_V start_POSTSUBSCRIPT { 1 , 2 } end_POSTSUBSCRIPT ⊂ blackboard_V consisting of cluster amplitudes with only one and two particle-hole excitations. The corresponding truncation variety, V{1,2}subscript𝑉12V_{\{1,2\}}italic_V start_POSTSUBSCRIPT { 1 , 2 } end_POSTSUBSCRIPT is then defined as the smallest variety containing exp(𝕍{1,2})expsubscript𝕍12{\rm exp}(\mathbb{V}_{\{1,2\}})roman_exp ( blackboard_V start_POSTSUBSCRIPT { 1 , 2 } end_POSTSUBSCRIPT ). This construction of the truncation variety can be generalized to any CC truncation: Let σ𝜎\sigmaitalic_σ denote the CC truncation rank, i.e. σ[[N]]={1,,N}𝜎delimited-[]delimited-[]𝑁1𝑁\sigma\subset[\![N]\!]=\{1,\dots,N\}italic_σ ⊂ [ [ italic_N ] ] = { 1 , … , italic_N }, where N𝑁Nitalic_N denotes the number of electrons in the system. Then, the smallest variety containing exp(𝕍σ)expsubscript𝕍𝜎{\rm exp}(\mathbb{V}_{\sigma})roman_exp ( blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) is the σ𝜎\sigmaitalic_σ-truncation variety denoted Vσsubscript𝑉𝜎V_{\sigma}italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT.

Remarkably, the truncation varieties can be used to express the projective CC equations (6) as an eigenvalue problem, i.e.

[H^|Φ]σ=E[|Φ]σ,|ΦVσ,formulae-sequencesubscriptdelimited-[]^𝐻ketΦ𝜎𝐸subscriptdelimited-[]ketΦ𝜎ketΦsubscript𝑉𝜎[\hat{H}|\Phi\rangle]_{\sigma}=E[|\Phi\rangle]_{\sigma},\quad|\Phi\rangle\in V% _{\sigma},[ over^ start_ARG italic_H end_ARG | roman_Φ ⟩ ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = italic_E [ | roman_Φ ⟩ ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , | roman_Φ ⟩ ∈ italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT , (10)

where []σsubscriptdelimited-[]𝜎[\,\cdot\,]_{\sigma}[ ⋅ ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT is the projection onto coordinates with excitation level in σ{0}𝜎0\sigma\cup\{0\}italic_σ ∪ { 0 }, i.e.

[|Ψ]σ=Φμ|Ψfor|μ|σ{0}.formulae-sequencesubscriptdelimited-[]ketΨ𝜎inner-productsubscriptΦ𝜇Ψfor𝜇𝜎0[|\Psi\rangle]_{\sigma}=\langle\Phi_{\mu}|\Psi\rangle\quad{\rm for}~{}|\mu|\in% \sigma\cup\{0\}.[ | roman_Ψ ⟩ ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = ⟨ roman_Φ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT | roman_Ψ ⟩ roman_for | italic_μ | ∈ italic_σ ∪ { 0 } . (11)

This formulation is indeed equivalent to the projective CC equations in Eq. (6) for CCD, as well as for rank complete CC truncations, i.e. CCSD, CCSDT, CCSDTQ etc., see Theorem 5.11 in Ref. 60. The number of roots for these CC variants is bounded by the number of solutions to Eq. (10) for a generic Hamiltonian matrix H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG. We denote this number the CC degree of the corresponding truncation variety. The CC degree serves as a complexity measure for solving the CC equations and Bézout’s theorem provides the upper bound

CCdeg(σ)(dim(Vσ)+1)deg(Vσ),CCdeg𝜎dimensionsubscript𝑉𝜎1degreesubscript𝑉𝜎\mathrm{CCdeg(\sigma)}\leq(\dim(V_{\sigma})+1)\deg(V_{\sigma}),roman_CCdeg ( italic_σ ) ≤ ( roman_dim ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) + 1 ) roman_deg ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) , (12)

where dim(Vσ)dimensionsubscript𝑉𝜎\dim(V_{\sigma})roman_dim ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) is the dimension of the truncation variety and deg(Vσ)degreesubscript𝑉𝜎\deg(V_{\sigma})roman_deg ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) is its degree. The dimension of the truncation variety is well understood. The reason is that cluster analysis allows us to bijectively map cluster amplitudes to CI-coefficients using polynomial equations, see e.g. Refs. 12, 70, 32 or Proposition 2.4 in Ref. 60 for a mathematical proof. Then, since Vσsubscript𝑉𝜎V_{\sigma}italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT arises as a deformation of the linear space 𝕍σsubscript𝕍𝜎\mathbb{V}_{\sigma}blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT under a polynomial with polynomial inverse, the dimension of Vσsubscript𝑉𝜎V_{\sigma}italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT is equal to the dimension of 𝕍σsubscript𝕍𝜎\mathbb{V}_{\sigma}blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT. For example in the case of CCSD, we obtain

dim(Vσ)=dim(𝕍σ)=nvirtnocc+(nvirt2)(nocc2).dimensionsubscript𝑉𝜎dimensionsubscript𝕍𝜎subscript𝑛virtsubscript𝑛occbinomialsubscript𝑛virt2binomialsubscript𝑛occ2\dim(V_{\sigma})=\dim(\mathbb{V}_{\sigma})=n_{\rm virt}n_{\rm occ}+\binom{n_{% \rm virt}}{2}\binom{n_{\rm occ}}{2}.roman_dim ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) = roman_dim ( blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) = italic_n start_POSTSUBSCRIPT roman_virt end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_occ end_POSTSUBSCRIPT + ( FRACOP start_ARG italic_n start_POSTSUBSCRIPT roman_virt end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) ( FRACOP start_ARG italic_n start_POSTSUBSCRIPT roman_occ end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) . (13)

The degree of Vσsubscript𝑉𝜎{V}_{\sigma}italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT on the other hand is not as easily computable and requires a more profound understanding of Vσsubscript𝑉𝜎{V}_{\sigma}italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT. For special cases, however, explicit formulas for the degree have been established; for example, for CCS, the truncation variety is the Grassmanian (see Theorem 3.5 in Ref. 60) which is mathematically well understood. In other cases we rely either on symbolic methods, provided by e.g. Macaulay2 71, or on numerical methods, accessible in e.g. HomotopyContinuation.jl 72, to compute the degree.

Despite the complexity of calculating deg(Vσ)degreesubscript𝑉𝜎\deg(V_{\sigma})roman_deg ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) the bound given by Eq. (12) provides significant insight into bounding the CC degree, since it is less complex to calculate deg(Vσ)degreesubscript𝑉𝜎\deg(V_{\sigma})roman_deg ( italic_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) than to calculate the CC degree directly. Moreover, this upper bound is notably better than the previously known upper bounds obtained by applying Bézout’s theorem directly to the projective CC equations, see Section 6 in Ref. 60.

4 Computational methods

In this section, we outline the computational procedure employed to compute the CC degree and the CC roots for a given CC truncation rank σ[[N]]𝜎delimited-[]delimited-[]𝑁\sigma\subset[\![N]\!]italic_σ ⊂ [ [ italic_N ] ]. We begin by introducing the CC family, a parametric polynomial family for Eq. (10), i.e.

CC(σ)={[(G^EIK)eT^(𝐭)]σ|G^𝔽K×K},subscriptCC𝜎conditional-setsubscriptdelimited-[]^𝐺𝐸subscript𝐼𝐾superscript𝑒^𝑇𝐭𝜎^𝐺superscript𝔽𝐾𝐾\mathcal{F}_{\rm CC}(\sigma)=\{[(\hat{G}-E\cdot I_{K})e^{\hat{T}(\mathbf{t})}]% _{\sigma}~{}\big{|}~{}\hat{G}\in\mathbb{F}^{K\times K}\},caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT ( italic_σ ) = { [ ( over^ start_ARG italic_G end_ARG - italic_E ⋅ italic_I start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT | over^ start_ARG italic_G end_ARG ∈ blackboard_F start_POSTSUPERSCRIPT italic_K × italic_K end_POSTSUPERSCRIPT } , (14)

where the energy E𝐸E\in\mathbb{C}italic_E ∈ blackboard_C and cluster amplitudes 𝐭𝕍σ𝐭subscript𝕍𝜎\mathbf{t}\in\mathbb{V}_{\sigma}bold_t ∈ blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT are the variables, and the entries of the matrix G^^𝐺\hat{G}over^ start_ARG italic_G end_ARG are the parameters. We will ultimately set G^=H^^𝐺^𝐻\hat{G}=\hat{H}over^ start_ARG italic_G end_ARG = over^ start_ARG italic_H end_ARG to be the Hamiltonian of interest, but CC(σ)subscriptCC𝜎\mathcal{F}_{\rm CC}(\sigma)caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT ( italic_σ ) is a much broader family of CC-like systems with G^𝔽K×K^𝐺superscript𝔽𝐾𝐾\hat{G}\in\mathbb{F}^{K\times K}over^ start_ARG italic_G end_ARG ∈ blackboard_F start_POSTSUPERSCRIPT italic_K × italic_K end_POSTSUPERSCRIPT an arbitrary matrix. The parameter continuation theorem provides an upper bound to the number of roots of any polynomial system FG^CCsubscript𝐹^𝐺subscriptCCF_{\hat{G}}\in\mathcal{F}_{\rm CC}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG end_POSTSUBSCRIPT ∈ caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT. Moreover, this bound is tight and equality is obtained for all polynomial systems with parameters in a certain dense subspace of 𝔽K×Ksuperscript𝔽𝐾𝐾\mathbb{F}^{K\times K}blackboard_F start_POSTSUPERSCRIPT italic_K × italic_K end_POSTSUPERSCRIPT. We subsequently assume the validity of the parameter continuation theorem and refer the interested reader to Section 3 in Ref. 73. The upper bound provided by the parameter continuation theorem yields the CC degree mentioned in the previous section.

The numerical procedure used in this work consists of two steps: First, we solve a general system of equations from the CC family, CCsubscriptCC\mathcal{F}_{\rm CC}caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT, using monodromy methods. Second, we use parameter homotopy continuation to track the solutions from this solved but generic system of equations to solutions of the CC equations describing a specific chemical system of interest. Both steps are subsequently outlined in detail.

We emphasize that, at their core, both methods use a homotopy to track new solutions from known solutions 74, 75, 76, 77. More precisely, we define a homotopy for a (piecewise smooth) path γ:[0,1]𝔽K×K:𝛾01superscript𝔽𝐾𝐾\gamma:[0,1]\to\mathbb{F}^{K\times K}italic_γ : [ 0 , 1 ] → blackboard_F start_POSTSUPERSCRIPT italic_K × italic_K end_POSTSUPERSCRIPT in the parameter space via

H(E,𝐭,λ)=Fγ(λ)(E,𝐭),𝐻𝐸𝐭𝜆subscript𝐹𝛾𝜆𝐸𝐭H(E,\mathbf{t},\lambda)=F_{\gamma(\lambda)}(E,\mathbf{t}),italic_H ( italic_E , bold_t , italic_λ ) = italic_F start_POSTSUBSCRIPT italic_γ ( italic_λ ) end_POSTSUBSCRIPT ( italic_E , bold_t ) , (15)

where H(E,𝐭,λ)𝐻𝐸𝐭𝜆H(E,\mathbf{t},\lambda)italic_H ( italic_E , bold_t , italic_λ ) describes the continuous deformation of a start system G(E,𝐭)=H(E,𝐭,1)𝐺𝐸𝐭𝐻𝐸𝐭1G(E,\mathbf{t})=H(E,\mathbf{t},1)italic_G ( italic_E , bold_t ) = italic_H ( italic_E , bold_t , 1 ) to a target system F(E,𝐭)=H(E,𝐭,0)𝐹𝐸𝐭𝐻𝐸𝐭0F(E,\mathbf{t})=H(E,\mathbf{t},0)italic_F ( italic_E , bold_t ) = italic_H ( italic_E , bold_t , 0 ). The individual solution paths to the system H(E,𝐭,λ)𝐻𝐸𝐭𝜆H(E,\mathbf{t},\lambda)italic_H ( italic_E , bold_t , italic_λ ) are then tracked from G(E,𝐭)𝐺𝐸𝐭G(E,\mathbf{t})italic_G ( italic_E , bold_t ) to F(E,𝐭)𝐹𝐸𝐭F(E,\mathbf{t})italic_F ( italic_E , bold_t ) as λ0𝜆0\lambda\to 0italic_λ → 0. This is accomplished via an ordinary differential equation, known as the Davidenko differential equation 78, 79, i.e.

𝐱H(𝐱,λ)(ddλ𝐱(λ))+λH(𝐱,λ)=0,𝐱𝐻𝐱𝜆dd𝜆𝐱𝜆𝜆𝐻𝐱𝜆0\frac{\partial}{\partial\mathbf{x}}H(\mathbf{x},\lambda)\left(\frac{\mathrm{d}% }{\mathrm{d}\lambda}\mathbf{x}(\lambda)\right)+\frac{\partial}{\partial\lambda% }H(\mathbf{x},\lambda)=0,divide start_ARG ∂ end_ARG start_ARG ∂ bold_x end_ARG italic_H ( bold_x , italic_λ ) ( divide start_ARG roman_d end_ARG start_ARG roman_d italic_λ end_ARG bold_x ( italic_λ ) ) + divide start_ARG ∂ end_ARG start_ARG ∂ italic_λ end_ARG italic_H ( bold_x , italic_λ ) = 0 , (16)

initialized by a root 𝐱𝐱\mathbf{x}bold_x of the start system, i.e. G(𝐱)=0𝐺𝐱0G(\mathbf{x})=0italic_G ( bold_x ) = 0 where we set 𝐱=(E,𝐭)×𝕍σ𝐱𝐸𝐭subscript𝕍𝜎\mathbf{x}=(E,\mathbf{t})\in\mathbb{C}\times\mathbb{V}_{\sigma}bold_x = ( italic_E , bold_t ) ∈ blackboard_C × blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT.

4.1 Monodromy

The monodromy solver is a numerical technique specifically designed for solving generic polynomial systems within a parameterized family. We emphasize that the monodromy solver requires the system to be generic, meaning it is not suitable for solving a specific set of CC equations directly. However, it can be employed to establish a start system FG^CCsubscript𝐹^𝐺subscriptCCF_{\hat{G}}\in\mathcal{F}_{\rm CC}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG end_POSTSUBSCRIPT ∈ caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT within the CC family. This start system is then used by the parameter homotopy continuation method to solve a (physical) target system of interest. Note that the monodromy method is required only once for a particular system configuration, i.e. the number of electrons, the number of orbitals, and the CC truncation level σ[[N]]𝜎delimited-[]delimited-[]𝑁\sigma\in[\![N]\!]italic_σ ∈ [ [ italic_N ] ].

We start with a root 𝐱×𝕍σ𝐱subscript𝕍𝜎\mathbf{x}\in\mathbb{C}\times\mathbb{V}_{\sigma}bold_x ∈ blackboard_C × blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT of a generic system FG^CCsubscript𝐹^𝐺subscriptCCF_{\hat{G}}\in\mathcal{F}_{\rm CC}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG end_POSTSUBSCRIPT ∈ caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT in the CC family and a loop γ:[0,1]𝔽K×K:𝛾01superscript𝔽𝐾𝐾\gamma:[0,1]\to\mathbb{F}^{K\times K}italic_γ : [ 0 , 1 ] → blackboard_F start_POSTSUPERSCRIPT italic_K × italic_K end_POSTSUPERSCRIPT in the parameter space, based at the generic matrix G^^𝐺\hat{G}over^ start_ARG italic_G end_ARG, i.e. γ(0)=γ(1)=G^𝛾0𝛾1^𝐺\gamma(0)=\gamma(1)=\hat{G}italic_γ ( 0 ) = italic_γ ( 1 ) = over^ start_ARG italic_G end_ARG. Moreover, we let 𝐱(λ)𝐱𝜆\mathbf{x}(\lambda)bold_x ( italic_λ ) be the solution path of H(E,𝐭,λ)=0𝐻𝐸𝐭𝜆0H(E,\mathbf{t},\lambda)=0italic_H ( italic_E , bold_t , italic_λ ) = 0 along γ𝛾\gammaitalic_γ, starting at 𝐱(1)=𝐱𝐱1𝐱\mathbf{x}(1)=\mathbf{x}bold_x ( 1 ) = bold_x. Since γ𝛾\gammaitalic_γ is a loop, the start system in this homotopy is equal to the end system. In particular, the solution path ends again at a root of FG^subscript𝐹^𝐺F_{\hat{G}}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG end_POSTSUBSCRIPT. The monodromy solver exploits the fact that the root 𝐱(0)𝐱0\mathbf{x}(0)bold_x ( 0 ) need not necessarily be equal to the root 𝐱(1)𝐱1\mathbf{x}(1)bold_x ( 1 ). In fact, let (𝐱1,,𝐱M)subscript𝐱1subscript𝐱𝑀(\mathbf{x}_{1},\dots,\mathbf{x}_{M})( bold_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , bold_x start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ) be all roots to FG^subscript𝐹^𝐺F_{\hat{G}}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG end_POSTSUBSCRIPT, then the roots (𝐱1(0),,𝐱M(0))subscript𝐱10subscript𝐱𝑀0(\mathbf{x}_{1}(0),\dots,\mathbf{x}_{M}(0))( bold_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( 0 ) , … , bold_x start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ( 0 ) ) define a mere permutation the system’s roots.

We exemplify this procedure with the simple parametric polynomial family:

={x36x2+11zx6|z},conditional-setsuperscript𝑥36superscript𝑥211𝑧𝑥6𝑧\mathcal{F}=\{x^{3}-6\,x^{2}+11\,z\,x-6~{}\big{|}~{}z\in\mathbb{C}\},caligraphic_F = { italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 6 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 11 italic_z italic_x - 6 | italic_z ∈ blackboard_C } , (17)

where x𝑥x\in\mathbb{C}italic_x ∈ blackboard_C is the variable and z𝑧z\in\mathbb{C}italic_z ∈ blackboard_C is the parameter. For z=1𝑧1z=1italic_z = 1, the corresponding polynomial

x36x2+11x6=(x1)(x2)(x3)superscript𝑥36superscript𝑥211𝑥6𝑥1𝑥2𝑥3x^{3}-6\,x^{2}+11\,x-6=(x-1)(x-2)(x-3)italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 6 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 11 italic_x - 6 = ( italic_x - 1 ) ( italic_x - 2 ) ( italic_x - 3 ) (18)

has the roots x1=1,x2=2,x3=3formulae-sequencesubscript𝑥11formulae-sequencesubscript𝑥22subscript𝑥33x_{1}=1,x_{2}=2,x_{3}=3italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1 , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 2 , italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 3. The monodromy solver then tracks these roots along the path γ:[0,1],tei2πt:𝛾formulae-sequence01maps-to𝑡superscript𝑒𝑖2𝜋𝑡\gamma:[0,1]\to\mathbb{C},t\mapsto e^{i2\pi t}italic_γ : [ 0 , 1 ] → blackboard_C , italic_t ↦ italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_t end_POSTSUPERSCRIPT. Note that γ𝛾\gammaitalic_γ is a loop based at 1111. Indeed, this procedure permutes the roots, see Figure 2.

Refer to caption
Figure 2: The solution paths of the roots of x36x2+11zx6superscript𝑥36superscript𝑥211𝑧𝑥6x^{3}-6x^{2}+11zx-6italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 6 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 11 italic_z italic_x - 6 for z(λ)=γ(λ)𝑧𝜆𝛾𝜆z(\lambda)=\gamma(\lambda)italic_z ( italic_λ ) = italic_γ ( italic_λ ) along the path γ:[0,1],tei2πt:𝛾formulae-sequence01maps-to𝑡superscript𝑒𝑖2𝜋𝑡\gamma:[0,1]\to\mathbb{C},t\mapsto e^{i2\pi t}italic_γ : [ 0 , 1 ] → blackboard_C , italic_t ↦ italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_t end_POSTSUPERSCRIPT.

This approach can be effectively utilized to find all solutions to a generic system of equations. For instance, consider a scenario where the only known solution for the polynomial in Eq. (18) is x2=2subscript𝑥22x_{2}=2italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 2. By tracking this solution along the unit circle, we can determine another solution, x1=1subscript𝑥11x_{1}=1italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1. When we extend this tracking across sufficiently many loops, the method reveals all possible roots of the system.

This process is central to the monodromy solver. The solver begins with a single known solution and numerically traces known solutions through multiple loops within the parameter space. This tracking continues until no additional solutions are found over a predetermined number of loops, leading to the conclusion that all possible solutions to the system have been numerically identified.

The parameters in the CC family CCsubscriptCC\mathcal{F}_{\rm CC}caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT are linear, hence, we can find a generic start system along with a root by choosing a pair (E,𝐭)×𝕍σ𝐸𝐭subscript𝕍𝜎(E,\mathbf{t})\in\mathbb{C}\times\mathbb{V}_{\sigma}( italic_E , bold_t ) ∈ blackboard_C × blackboard_V start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT at random, and then compute one generic solution G^0subscript^𝐺0\hat{G}_{0}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT fulfilling

[(G^0EIK)eT^(𝐭)]σ=0.subscriptdelimited-[]subscript^𝐺0𝐸subscript𝐼𝐾superscript𝑒^𝑇𝐭𝜎0[(\hat{G}_{0}-E\cdot I_{K})e^{\hat{T}(\mathbf{t})}]_{\sigma}=0.[ ( over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_E ⋅ italic_I start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = 0 .

We then use the monodromy solver to compute all roots to FG^0subscript𝐹subscript^𝐺0F_{\hat{G}_{0}}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT.

4.2 Parametric homotopy continuation

Applying the parametric homotopy continuation method to find the roots of the CC equations corresponding to a system of interest is now straightforward. The goal is to solve

FH^(E,𝐭)=[(H^EIK)eT^(𝐭)]σ=0subscript𝐹^𝐻𝐸𝐭subscriptdelimited-[]^𝐻𝐸subscript𝐼𝐾superscript𝑒^𝑇𝐭𝜎0F_{\hat{H}}(E,{\mathbf{t}})=[(\hat{H}-E\cdot I_{K})e^{\hat{T}(\mathbf{t})}]_{% \sigma}=0italic_F start_POSTSUBSCRIPT over^ start_ARG italic_H end_ARG end_POSTSUBSCRIPT ( italic_E , bold_t ) = [ ( over^ start_ARG italic_H end_ARG - italic_E ⋅ italic_I start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ) italic_e start_POSTSUPERSCRIPT over^ start_ARG italic_T end_ARG ( bold_t ) end_POSTSUPERSCRIPT ] start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = 0 (19)

for a given Hamiltonian H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG, arising from a physical electronic system. By construction, both H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG and G^0subscript^𝐺0\hat{G}_{0}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are in the parameter space of CCsubscriptCC\mathcal{F}_{\rm CC}caligraphic_F start_POSTSUBSCRIPT roman_CC end_POSTSUBSCRIPT. In particular, we can define a path γ𝛾\gammaitalic_γ in the parameter space that continuously transforms G^0subscript^𝐺0\hat{G}_{0}over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT into H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG. Since we have computed all roots to the system FG^0subscript𝐹subscript^𝐺0F_{\hat{G}_{0}}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT, the Davidenko differential equation (16) allows the tracking of solutions of FG^0subscript𝐹subscript^𝐺0F_{\hat{G}_{0}}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_G end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT to FH^subscript𝐹^𝐻F_{\hat{H}}italic_F start_POSTSUBSCRIPT over^ start_ARG italic_H end_ARG end_POSTSUBSCRIPT using e.g. Newton’s method.

As λ0𝜆0\lambda\to 0italic_λ → 0, we may encounter different scenarios when getting close to λ=0𝜆0\lambda=0italic_λ = 0. Here, the endgame 80 should be employed to determine the target solutions and their “type”. We illustrate such different scenarios in Figure 3.

Refer to caption
Figure 3: Sketch of possible homotopy paths. The solid line shows a path with no finite limit as λ0𝜆0\lambda\to 0italic_λ → 0, the dashed lines have the same limit of a singular root, and the dotted-dashed line has a unique limit of a nonsingular root.

We conclude this section by highlighting the progress that has been made within homotopy continuation based software development exploiting parallel implementations, which can significantly extend their application areas in the coming years. In particular, we here want to emphasize PHCpack 81, Bertini, 82 HOM4PS, 83, 84 NAG4M2 74 and HomotopyContinuation.jl72

5 Numerical results

In this section we investigate the root structure and potential energy curves (PECs) for ground and excited states at the level of CCD and CCSD. We report PECs corresponding to different CC solutions obtained by the above-outlined computational procedure. The obtained energies and corresponding eigenstate approximations are then compared to exact eigenpairs of the respective Hamiltonian. Besides the computation and analysis of the CC states, we compare different bounds to the number of CC roots with the computed number of CC solutions along different bond stretching processes. We restrict our investigations to four electron systems, such as (H2)2 in D2h and D∞h configurations, H4 symmetrically disturbed on a circle 85, and LiH. The reported computations are performed in first quantization and the considered systems are all at half-filling, i.e. four electrons in eight spin orbitals. The one-body and two-body integrals for the considered systems are obtained from the Python-based Simulations of Chemistry Framework (PySCF86, 87, 88. For the algebraic simulations, we used the monodromy and parametric homotopy continuation methods in HomotopyContinuation.jl. 72

5.1 CC bounds for generic four electron systems at half-filling

We emphasize that the theoretical bounds hold for generic Hamiltonians, and therefore only depend on the number of electrons, the number of spin orbitals, and the CC truncation level.

Coupled cluster doubles:

The roughest bound is obtained by applying Bézout’s theorem directly to the CCD equations. Note that for four electron systems, the CCD equations are quadratic. Moreover, in first quantization, there are (42)2=36superscriptbinomial42236{4\choose 2}^{2}=36( binomial start_ARG 4 end_ARG start_ARG 2 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 36 doubly excited Slater determinants. Hence, the Bézout bound directly applied to the CCD equations yields 236superscript2362^{36}2 start_POSTSUPERSCRIPT 36 end_POSTSUPERSCRIPT as an upper bound for the number of roots. The second bound we compare with is the bound in Eq. (12). Again, in first quantization, there are 36 CCD amplitudes, hence, dim(V{2})=36dimsubscript𝑉236{\rm dim}(V_{\{2\}})=36roman_dim ( italic_V start_POSTSUBSCRIPT { 2 } end_POSTSUBSCRIPT ) = 36. Moreover we compute the degree of V{2}subscript𝑉2V_{\{2\}}italic_V start_POSTSUBSCRIPT { 2 } end_POSTSUBSCRIPT using Macaulay2 yielding deg(V{2})=2degsubscript𝑉22{\rm deg}(V_{\{2\}})=2roman_deg ( italic_V start_POSTSUBSCRIPT { 2 } end_POSTSUBSCRIPT ) = 2. Overall, this yields 74747474 as an upper bound. For a generic Hamiltonian, we can moreover compute the exact number of roots to the CCD equations using the monodromy method which yields CCdeg({2})4,8=73{}_{4,8}(\{2\})=73start_FLOATSUBSCRIPT 4 , 8 end_FLOATSUBSCRIPT ( { 2 } ) = 73. This quantifies the improvement provided by the bound in Eq. (12) over the direct application of Bézout’s theorem.

Coupled cluster singles and doubles:

When considering CCSD instead of CCD these numbers dramatically increase. Frist, we note that the Bézout bound directly applied to the CCSD equation yields the upper bound 3ns4ndsuperscript3subscript𝑛𝑠superscript4subscript𝑛𝑑3^{n_{s}}4^{n_{d}}3 start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT 4 start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, where for four electrons in eight spin orbitals ns=16subscript𝑛𝑠16n_{s}=16italic_n start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 16 and nd=36subscript𝑛𝑑36n_{d}=36italic_n start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 36. This leads to approximately 210292superscript10292\cdot 10^{29}2 ⋅ 10 start_POSTSUPERSCRIPT 29 end_POSTSUPERSCRIPT solutions. Second, we note that dim(V{1,2})=52dimsubscript𝑉1252{\rm dim}(V_{\{1,2\}})=52roman_dim ( italic_V start_POSTSUBSCRIPT { 1 , 2 } end_POSTSUBSCRIPT ) = 52 and numerical calculations yield deg(V{1,2})=442066degsubscript𝑉12442066{\rm deg}(V_{\{1,2\}})=442066roman_deg ( italic_V start_POSTSUBSCRIPT { 1 , 2 } end_POSTSUBSCRIPT ) = 442066, hence, the bound provided by Eq. (12) yields approximately 21072superscript1072\cdot 10^{7}2 ⋅ 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT solutions. Numerical calculations show that CCdeg({1,2})4,8=16 952 996{}_{4,8}(\{1,2\})=16\,952\,996start_FLOATSUBSCRIPT 4 , 8 end_FLOATSUBSCRIPT ( { 1 , 2 } ) = 16 952 996. This clearly illustrates the improvement provided by the bound in Eq. (12) over the direct application of Bézout’s theorem.

To connect these theoretical bounds to practical systems, we subsequently calculate the number of CC solutions that result in real-valued energies, as well as those that accurately predict excited state energies at different bond lengths for the considered systems. Additionally, we report the number of FCI solutions. This comparison allows us to assess the accuracy of the bounds established for generic Hamiltonians.

5.2 Dissociation of lithium hydride – CCD

We initiate our energy and state approximation investigations by looking at the dissociation process of lithium hydride. We consider bond distances ranging from 1.3751.3751.3751.375 to 5.955.955.955.95 bohr. Diagonalizing the Hamiltonian yields the full spectrum reported in Figure 16(a) (see supplementary material) together with all roots to the CCD equations. Due to the large number of FCI solutions, we minimize the distance between FCI solutions and CCD solutions showing a much clearer picture of potentially well-resolved PECs, see Figure 16(b) in the supplementary material. This graph allows us to extract the PECs of the different states that are well-approximated using CCD, see Figure 4(a). We find that 10 PECs are approximated up to 51025superscript1025\cdot 10^{-2}5 ⋅ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT hartree for the entire dissociation process, however, we also note that for various bond distances, more CCD energies approximate FCI energies well, see Figure 16(b) and Table 1. For the 10 PECs, we moreover compute the overlap between the CCD states and the approximated eigenstates, see Figure 4(b). Remarkably, three of the four lowest energy states are well approximated both in terms of energy and states. The highest energy state stands out as well since it is well-approximated near the equilibrium, both in terms of energy as well as the eigenstate. However, as we transition away from the equilibrium the level of approximation deteriorates. Note that poor overlap may be detected when the eigenspace is higher dimensional, that is if the corresponding eigenvalue has multiplicity two or higher. Therefore some of the higher energy states reporting bad overlap might have CCD states that are close to the actual eigenspace – or even within the eigenspace – since we are only reporting overlap with one representative from the eigenspace. A detailed investigation of this is left for future works.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: (a) PECs of LiH that can be accurately approximated by CCD energies. (b) Overlaps of the CCD states with the corresponding eigenstate.

Investigating the root structure, we observe that the total number of CCD roots fluctuates along the bond stretching process, see Table 1. We find that any bound for a generic Hamiltonian severely overestimates the true number of roots for this system.

Bond distances 1.375 1.6 1.9 2.2 2.5 3.1 4.0 4.9 5.8
##\## CCD real 33 40 33 40 34 28 30 33 28
##\## CCD approx. 22 25 20 23 18 19 17 18 14
Table 1: The number of CCD roots in LiH dissociation for selected bond distances. By "##\## CCD approx." we denote the energetically relevant CCD roots. For comparison, we recall that the Bézout bound led to 236superscript2362^{36}2 start_POSTSUPERSCRIPT 36 end_POSTSUPERSCRIPT solutions, the bound in Eq. (12) led to 74 solutions, and CCdeg4,8({2}) is 73.

5.3 Lithium hydride – CCSD

We moreover perform CCSD computations on lithium hydride at the bond distance 2.875 bohr – close to the molecule’s equilibrium. Computing all roots to CCSD equations yields 15 9541595415\,95415 954 roots, 2 17021702\,1702 170 of which yield real-values energies, and 1 28012801\,2801 280 are real-valued solutions. Visualizing all 15 9541595415\,95415 954 CCSD energies shows that the energies tend to cluster around the FCI solutions, see Figure 5(a). Being able to compare with the true Hamiltonian spectrum reveals that only 26 of these CCSD solutions yield energies that are close to an FCI energy up to 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT hartree, see Figure 5(b).

Refer to caption
(a)
Refer to caption
(b)
Figure 5: (a) FCI solutions (red lines) together with all CCSD energies (b) CCSD energies that approximate FCI energies up to 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT hartree.

For all 70 eigenstates, we compute the energetically closest CCSD state together with the overlap with the corresponding states, see Figure 6. We find that three of the 26 energetically relevant CCSD states have good overlaps with the targeted states. Therefore there are at least three CCSD states that are well approximated both in terms of energy and state.

Refer to caption
Figure 6: Overlap of the energetically closest CCSD state with the corresponding eigenstate. The blue dots show the 26 energetically relevant CCSD states.

5.4 Dissociating systems of hydrogen

We now investigate the dissociation of (H2)2 planar model systems in different geometries 38.

5.4.1 (H2)2 dissociation (D2h symmetry)

The bond stretching procedure for (H2)2 in D2h symmetry is sketched in Figure 7. We consider an intra-molecular bond distance of R=1.4𝑅1.4R=1.4italic_R = 1.4 bohr, which corresponds to the equilibrium geometry of H2; this is kept fixed during the dissociation process. The inter-molecular distance is varied from 1.5 to 4.0 bohr.

Refer to caption
Figure 7: Schematic depiction of the dissociation process of (H2)2 in D2h configuration. The parameter R=1.4𝑅1.4R=1.4italic_R = 1.4 bohr.

The full spectrum together with the CCD solutions can be found in Figure 13(a) in the Supplementary material, and the PECs together with the closest CCD solutions are reported in Figure 13(b). We extract the PECs that are well-approximated using CCD and report them in Figure 8(a). Here, we observe that eight CCD energies approximate FCI PECs well. For these eight CCD states, we compute the overlap with the corresponding eigenstate, see Figure 8(b). We see that the ground state is well approximated, both in terms of energy as well as the state. The following six states (i.e. states two to seven in Figure 8(b)) have intermediate overlaps with the targeted states; the last state has a poor overlap with the targeted state. This illustrates that CC amplitudes could yield good energy approximations while providing poor approximations to the actual eigenstates 13.

Refer to caption
(a)
Refer to caption
(b)
Figure 8: (a) PECs of (H2)2 in D2h symmetry that are accurately described by CCD energies. (b) Overlap of the CCD states with the corresponding eigenstate.

In Table 2 we report the number of CCD roots yielding real-valued energies and energetic relevant CCD roots for selected bond distances. We observe that the number of CCD roots fluctuates along the bond stretching procedure.

Bond distances 1.6 1.9 2.1 2.5 3.0 3.9
##\## CCD real 66 45 60 64 55 64
##\## CCD approx. 32 26 33 32 33 34
Table 2: The number of roots along the (H2)2 dissociation in D2h configuration for selected bond distances. By "##\## CCD approx." we denote the energetically relevant CCD roots. For comparison, we recall that the Bézout bound is 236superscript2362^{36}2 start_POSTSUPERSCRIPT 36 end_POSTSUPERSCRIPT, the bound in Eq. (12) yields 74, and CCdeg4,8({2}) is 73.

5.4.2 (H2)2 dissociation (D∞h symmetry)

Next, we investigate the (H2)2 dissociation in D∞h symmetry, see Figure 9 for a sketch of the dissociation process. The intra-molecular distance is again set to R=1.4𝑅1.4R=1.4italic_R = 1.4 bohr and is kept fixed during the dissociation process. The inter-molecular distance is varied from 1.5 to 4.0 bohr.

Refer to caption
Figure 9: Schematic depiction of the dissociation process of (H2)2 in D∞h configuration. The parameter R=1.4𝑅1.4R=1.4italic_R = 1.4 bohr.

The full spectrum and CCD solutions is presented in Figure 14(a) in the Supplementary material, with the PECs and their closest CCD solutions detailed in Figure 14(b). We have identified PECs that are accurately approximated by CCD, as shown in Figure 10(a), where four CCD energies closely match the FCI PECs. The overlap of these four CCD states with their respective eigenstates is analyzed in Figure 10(b). Consistent with observations from the (H2)2 in D2h configuration, the ground state demonstrates a high degree of accuracy, while the other three states exhibit intermediate to poor overlap with their targeted eigenstates.

Refer to caption
(a)
Refer to caption
(b)
Figure 10: (a) PECs of (H2)2 in D∞h symmetry that are accurately described by CCD energies. (b) Overlap of the CCD states with the corresponding eigenstate.

Investigating the root structure, we observe that – similar to (H2)2 in D2h symmetry – the total number of CCD roots fluctuates along the bond stretching procedure, see Table 3. Note that for (H2)2 in D∞h symmetry, the total number of CCD roots is lower than the number of CCD roots (H2)2 in D2h symmetry. In particular, any bound for a generic Hamiltonian severely overestimates the true number of roots for this system.

Bond distances 1.6 1.9 2.1 2.5 3.0 3.9
##\## CCD real 46 42 40 38 37 35
##\## CCD approx. 13 19 12 14 16 21
Table 3: The number of roots along the (H2)2 dissociation in D∞h configuration for selected bond distances. By "##\## CCD approx." we denote the energetically relevant CCD roots. For comparison, we recall that the Bézout bound led to 236superscript2362^{36}2 start_POSTSUPERSCRIPT 36 end_POSTSUPERSCRIPT solutions, the bound in Eq. (12) led to 74 solutions, and CCdeg4,8({2}) is 73.

5.5 H4 disturbed on a circle

We proceed by investigating a variant of the H4 model consisting of four hydrogen atoms symmetrically distributed on a circle of radius R=2𝑅2R=\sqrt{2}italic_R = square-root start_ARG 2 end_ARG bohr 89, see Fig. 11.

Refer to caption
Figure 11: Schematic depiction of the H4 model undergoing a symmetric disturbance on a circle modeled by the angle ΘΘ\Thetaroman_Θ.

The full spectrum together with the CCD solutions can be found in Figure 15(a) in the Supplementary material, and the PECs together with the closest CCD solutions are reported in Figure 15(b). We extract the PECs that are well-approximated using CCD which are reported in Figure 12(a). Note that it is well-known that for this system CCD does not perform well in the region close to 90, due to strong degeneracies 38. This results in a dramatic reduction of the total number of roots, we find three roots at 90superscript9090^{\circ}90 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT and six roots at 89superscript8989^{\circ}89 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. For completion, these energies are included in the supplementary material, however, since the focus of this work is on the root structure and potentially physically relevant roots accessible by single-reference CC theory, we focus on cases where single-reference CC theory provides reasonable approximations and therefore excluded these points in Figure 12(a). Outside of this challenging region, we find five solutions that approximate PECs well. For these five solutions, we compute the overlap with the corresponding eigenstate, see Figure 12(b). Consistent with the previous H4 systems, this model system shows the effect of CCD states accurately resolving excited energies but providing poor overlap to the exact eigenstates. The only state that shows a good approximation to its eigenstate is the CCD ground state.

Refer to caption
(a)
Refer to caption
(b)
Figure 12: (a) PECs of H4 symmetrically distributed on a cycle that can be accurately reconstructed by CCD roots. (b) Overlap of the CCD states with the corresponding eigenstate.

Similar to (H2)2 in D∞h symmetry, we observe that generic bounds severely overestimate the number of CCD roots, though this effect is further amplified for H4 symmetrically distributed on a circle, see Table 4.

Angles in degrees 45 55 65 75 85
##\## CCD real 35 36 35 36 28
##\## CCD approx. 18 18 19 21 22
Table 4: The number of roots of H4 symmetrically distributed on a circle for selected bond distances. By "##\## CCD approx." we denote the energetically relevant CCD roots. For comparison, we recall that the Bézout bound led to 236superscript2362^{36}2 start_POSTSUPERSCRIPT 36 end_POSTSUPERSCRIPT solutions, the bound in Eq. (12) led to 74 solutions, and CCdeg4,8({2}) is 73.

6 Conclusion

The introduction highlighted critical unanswered questions in the study of the coupled cluster (CC) equations’ algebraic structures, including the total number of solutions, the best approximation element, and the feasibility of approximating excited states with higher-order roots of single reference CC equations. The ability to compute the full solution spectrum to the CC equations enabled us to tackle these questions for concrete physical systems.

Most remarkably, our results demonstrate that for lithium hydride in multiple geometries, higher-order CC solutions not only provide high-accuracy approximations to the energies but also to the states themselves. For a broader range of systems including multiple (H2)2 planar model systems in different geometries, we found that high-order CC solutions accurately predict several full potential energy curves, though they possibly provide a poor approximation to the eigenstates. While this effect has been previously reported, it requires further investigation to fully understand it and ultimately leverage it for practical applications.

Additionally, we examined the total number of roots and the effectiveness of various bounds developed for generic Hamiltonians. Our simulations indicate that while the most recent bounds perform exceptionally well for generic systems, they significantly overestimate the number of solutions in physical systems. More precisely, for all considered systems, we computed between 30303030-60606060 real solutions of which only 10101010-30303030 approximate physical states, compared to CC degree of 72727272. This excess of solutions gets amplified for CCSD: Solving the CC equations of a generic Hamiltonian yields 16 952 9961695299616\,952\,99616 952 996 solutions, compared to 1280128012801280 solutions that yield real-valued energies of the CC equations describing LiH. This overestimation by a factor of over ten thousand indicates that to further improve the bounds of the CC equations, specific structures of the Hamiltonian have to be taken into account.

{acknowledgement}

The authors thank Piotr Piecuch for insightful discussions.

References

  • Bartlett and Musiał 2007 Bartlett, R. J.; Musiał, M. Coupled-cluster theory in quantum chemistry. Reviews of Modern Physics 2007, 79, 291.
  • Coester 1958 Coester, F. Bound States of a Many-Particle System. Nuclear Physics 1958, 7, 421–424.
  • Hubbard 1957 Hubbard, J. The description of collective motions in terms of many-body perturbation theory. Proceedings of the Royal Society A 1957, 240, 539–560.
  • Hugenholtz 1957 Hugenholtz, N. Perturbation approach to the Fermi gas model of heavy nuclei. Physica 1957, 23, 533–545.
  • Čížek 1966 Čížek, J. On the Correlation Problem in Atomic and Molecular Systems. calculation of wavefunction components in Ursell-type expansion using quantum-field theoretical methods. The Journal of Chemical Physics 1966, 45, 4256–4266.
  • Kümmel 1991 Kümmel, H. Origins of the coupled cluster method. Theoretica Chimica Acta 1991, 80, 81–89.
  • Čížek 1991 Čížek, J. Origins of coupled cluster technique for atoms and molecules. Theoretica Chimica Acta 1991, 80, 91–94.
  • Bartlett 2005 Bartlett, R. In Theory and Applications of Computational Chemistry (The First Forty Years), 1st ed.; Dykstra, C., Frenking, G., Kim, K., Scuseria, G., Eds.; 2005; pp 1191–1221.
  • Paldus 2005 Paldus, J. In Theory and Applications of Computational Chemistry (The First Forty Years), 1st ed.; Dykstra, C., Frenking, G., Kim, K., Scuseria, G., Eds.; 2005; pp 115–147.
  • Arponen 1991 Arponen, J. S. Independent-cluster methods as mappings of quantum theory into classical mechanics. Theoretica Chimica Acta 1991, 80, 149–179.
  • Bishop 1991 Bishop, R. An overview of coupled cluster theory and its applications in Physics. Theoretica Chimica Acta 1991, 80, 95–148.
  • Helgaker et al. 2014 Helgaker, T.; Jörgensen, P.; Olsen, J. Molecular Electronic-Structure Theory; John Wiley & Sons, 2014.
  • Piecuch and Kowalski 2000 Piecuch, P.; Kowalski, K. In Computational Chemistry: Reviews of Current Trends; Leszczynski, J., Ed.; Singapore: World Scientific, 2000; Vol. 5.
  • Rowe 1968 Rowe, D. Equations-of-motion method and the extended shell model. Reviews of Modern Physics 1968, 40, 153.
  • Emrich 1981 Emrich, K. An extension of the coupled cluster formalism to excited states (I). Nuclear Physics A 1981, 351, 379–396.
  • Sekino and Bartlett 1984 Sekino, H.; Bartlett, R. J. A linear response, coupled-cluster theory for excitation energy. International Journal of Quantum Chemistry 1984, 26, 255–265.
  • Geertsen et al. 1989 Geertsen, J.; Rittby, M.; Bartlett, R. J. The equation-of-motion coupled-cluster method: Excitation energies of Be and CO. Chemical Physics Letters 1989, 164, 57–62.
  • Stanton and Bartlett 1993 Stanton, J. F.; Bartlett, R. J. The equation of motion coupled-cluster method. A systematic biorthogonal approach to molecular excitation energies, transition probabilities, and excited state properties. The Journal of chemical physics 1993, 98, 7029–7039.
  • Comeau and Bartlett 1993 Comeau, D. C.; Bartlett, R. J. The equation-of-motion coupled-cluster method. Applications to open-and closed-shell reference states. Chemical Physics Letters 1993, 207, 414–423.
  • Watts and Bartlett 1994 Watts, J. D.; Bartlett, R. J. The inclusion of connected triple excitations in the equation-of-motion coupled-cluster method. The Journal of chemical physics 1994, 101, 3073–3078.
  • Monkhorst 1977 Monkhorst, H. J. Calculation of properties with the coupled-cluster method. International Journal of Quantum Chemistry 1977, 12, 421–432.
  • Dalgaard and Monkhorst 1983 Dalgaard, E.; Monkhorst, H. J. Some aspects of the time-dependent coupled-cluster approach to dynamic response functions. Physical Review A 1983, 28, 1217.
  • Koch and Jørgensen 1990 Koch, H.; Jørgensen, P. Coupled cluster response functions. The Journal of chemical physics 1990, 93, 3333.
  • Watson and Chan 2012 Watson, M. A.; Chan, G. K.-L. Excited states of butadiene to chemical accuracy: Reconciling theory and experiment. Journal of chemical theory and computation 2012, 8, 4013–4018.
  • Shu and Truhlar 2017 Shu, Y.; Truhlar, D. G. Doubly excited character or static correlation of the reference state in the controversial 21Ag state of trans-butadiene. Journal of the American Chemical Society 2017, 139, 13770–13778.
  • Barca et al. 2018 Barca, G. M.; Gilbert, A. T.; Gill, P. M. Excitation number: Characterizing multiply excited states. Journal of chemical theory and computation 2018, 14, 9–13.
  • Loos et al. 2019 Loos, P.-F.; Boggio-Pasqua, M.; Scemama, A.; Caffarel, M.; Jacquemin, D. Reference energies for double excitations. Journal of chemical theory and computation 2019, 15, 1939–1956.
  • Ravi et al. 2022 Ravi, M.; Perera, A.; Bartlett, R. J., et al. The intermediate state approach for doubly excited dark states in EOM-coupled-cluster theory. The Journal of Chemical Physics 2022, 156.
  • Do Casal et al. 2023 Do Casal, M. T.; Toldo, J. M.; Barbatti, M.; Plasser, F. Classification of doubly excited molecular electronic states. Chemical Science 2023, 14, 4012–4026.
  • Piecuch and Adamowicz 1994 Piecuch, P.; Adamowicz, L. Solving the single-reference coupled-cluster equations involving highly excited clusters in quasidegenerate situations. The Journal of chemical physics 1994, 100, 5857–5869.
  • Mihálka and Noga 2023 Mihálka, Z. É.; Noga, J. Exploring alternative approaches to improve the convergence pattern in solving the coupled-cluster equations. Molecular Physics 2023, 121, e2140084.
  • Shavitt and Bartlett 2009 Shavitt, I.; Bartlett, R. J. Many-body methods in chemistry and physics: MBPT and coupled-cluster theory; Cambridge university press, 2009.
  • Szakács and Surján 2008 Szakács, P.; Surján, P. R. Stability conditions for the coupled cluster equations. International Journal of Quantum Chemistry 2008, 108, 2043–2052.
  • Adams et al. 1981 Adams, B.; Jankowski, K.; Paldus, J. Symmetry-adapted coupled-pair approach to the many-electron correlation problem. II. Application to the Be atom. Physical Review A 1981, 24, 2316.
  • Faulstich 2024 Faulstich, F. M. Recent mathematical advances in coupled cluster theory. arXiv:2401.07383 2024,
  • Faulstich et al. 2024 Faulstich, F. M.; Khoo, Y.; Li, K. Augmented Lagrangian method for coupled-cluster. arXiv:2403.16381 2024,
  • Živković and Monkhorst 1978 Živković, T. P.; Monkhorst, H. J. Analytic connection between configuration–interaction and coupled-cluster solutions. Journal of Mathematical Physics 1978, 19, 1007–1022.
  • Paldus et al. 1993 Paldus, J.; Piecuch, P.; Pylypow, L.; Jeziorski, B. Application of Hilbert-space coupled-cluster theory to simple (H2)2 model systems: Planar models. Physical Review A 1993, 47, 2738.
  • Piecuch et al. 1990 Piecuch, P.; Zarrabian, S.; Paldus, J.; Čížek, J. Coupled-cluster approaches with an approximate account of triexcitations and the optimized-inner-projection technique. II. Coupled-cluster results for cyclic-polyene model systems. Physical Review B 1990, 42, 3351.
  • Kowalski and Jankowski 1998 Kowalski, K.; Jankowski, K. Towards complete solutions to systems of nonlinear equations of many-electron theories. Physical Review Letters 1998, 81, 1195.
  • Piecuch and Kowalski 2000 Piecuch, P.; Kowalski, K. In Computational Chemistry, Reviews of Current Trends; Leszczynski, J., Ed.; 2000; pp 1–104.
  • Kowalski and Piecuch 2000 Kowalski, K.; Piecuch, P. Complete set of solutions of multireference coupled-cluster equations: The state-universal formalism. Physical Review A 2000, 61, 052506.
  • Csirik, Mihály A. and Laestadius, Andre 2023 Csirik, Mihály A.,; Laestadius, Andre, Coupled-Cluster theory revisited - Part I: Discretization. ESAIM: Mathematical Modelling and Numerical Analysis 2023, 57, 645–670.
  • Csirik and Laestadius 2023 Csirik, M. A.; Laestadius, A. Coupled-Cluster theory revisited-Part II: Analysis of the single-reference Coupled-Cluster equations. ESAIM: Mathematical Modelling and Numerical Analysis 2023, 57, 545–583.
  • Kowalski and Piecuch 2000 Kowalski, K.; Piecuch, P. Complete set of solutions of the generalized Bloch equation. International Journal of Quantum Chemistry 2000, 80, 757–781.
  • Kowalski and Jankowski 1998 Kowalski, K.; Jankowski, K. Full solution to the coupled-cluster equations: the H4 model. Chemical Physics Letters 1998, 290, 180–188.
  • Jankowski and Kowalski 1999 Jankowski, K.; Kowalski, K. Physical and mathematical content of coupled–cluster equations: Correspondence between coupled–cluster and configuration–interaction solutions. The Journal of Chemical Physics 1999, 110, 3714–3729.
  • Jankowski and Kowalski 1999 Jankowski, K.; Kowalski, K. Physical and mathematical content of coupled-cluster equations. II. On the origin of irregular solutions and their elimination via symmetry adaptation. The Journal of Chemical Physics 1999, 110, 9345–9352.
  • Jankowski and Kowalski 1999 Jankowski, K.; Kowalski, K. Physical and mathematical content of coupled-cluster equations. III. Model studies of dissociation processes for various reference states. The Journal of Chemical Physics 1999, 111, 2940–2951.
  • Jankowski and Kowalski 1999 Jankowski, K.; Kowalski, K. Physical and mathematical content of coupled-cluster equations. IV. Impact of approximations to the cluster operator on the structure of solutions. The Journal of Chemical Physics 1999, 111, 2952–2959.
  • Kowalski and Piecuch 2000 Kowalski, K.; Piecuch, P. The method of moments of coupled-cluster equations and the renormalized CCSD[T], CCSD(T), CCSD(TQ), and CCSDT(Q) approaches. The Journal of Chemical Physics 2000, 113, 18–35.
  • Piecuch et al. 2002 Piecuch, P.; Kowalski, K.; Pimienta, I. S.; Mcguire, M. J. Recent advances in electronic structure theory: Method of moments of coupled-cluster equations and renormalized coupled-cluster approaches. International Reviews in Physical Chemistry 2002, 21, 527–655.
  • Piecuch et al. 2004 Piecuch, P.; Kowalski, K.; Pimienta, I.; Fan, P.-D.; Lodriguito, M.; McGuire, M.; Kucharski, S.; Kuś, T.; Musiał, M. Method of moments of coupled-cluster equations: a new formalism for designing accurate electronic structure methods for ground and excited states. Theoretical Chemistry Accounts 2004, 112, 349–393.
  • Shen and Piecuch 2012 Shen, J.; Piecuch, P. Biorthogonal moment expansions in coupled-cluster theory: Review of key concepts and merging the renormalized and active-space coupled-cluster methods. Chemical Physics 2012, 401, 180–202.
  • Kowalski and Piecuch 2000 Kowalski, K.; Piecuch, P. Renormalized CCSD(T) and CCSD(TQ) approaches: Dissociation of the N2 triple bond. The Journal of Chemical Physics 2000, 113, 5644–5652.
  • Piecuch et al. 2006 Piecuch, P.; Włoch, M.; Gour, J. R.; Kinal, A. Single-reference, size-extensive, non-iterative coupled-cluster approaches to bond breaking and biradicals. Chemical physics letters 2006, 418, 467–474.
  • Piecuch and Włoch 2005 Piecuch, P.; Włoch, M. Renormalized coupled-cluster methods exploiting left eigenstates of the similarity-transformed Hamiltonian. The Journal of chemical physics 2005, 123.
  • Faulstich and Laestadius 2023 Faulstich, F. M.; Laestadius, A. Homotopy continuation methods for coupled-cluster theory in quantum chemistry. Molecular Physics 2023, e2258599.
  • Faulstich and Oster 2024 Faulstich, F. M.; Oster, M. Coupled Cluster Theory: Toward an Algebraic Geometry Formulation. SIAM Journal on Applied Algebra and Geometry 2024, 8, 138–188.
  • Faulstich et al. 2023 Faulstich, F. M.; Sturmfels, B.; Sverrisdóttir, S. Algebraic varieties in quantum chemistry. arXiv:2308.05258 2023,
  • Borovik et al. 2023 Borovik, V.; Sturmfels, B.; Sverrisdóttir, S. Coupled cluster degree of the grassmannian. arXiv:2310.15474 2023,
  • Schneider 2009 Schneider, R. Analysis of the Projected Coupled Cluster Method in Electronic Structure Calculation. Numerische Mathematik 2009, 113, 433–471.
  • Rohwedder 2013 Rohwedder, T. The Continuous Coupled Cluster Formulation for the Electronic Schrödinger Equation. ESAIM: Mathematical Modelling and Numerical Analysis 2013, 47, 421–447.
  • Rohwedder and Schneider 2013 Rohwedder, T.; Schneider, R. Error Estimates for the Coupled Cluster Method. ESAIM: Mathematical Modelling and Numerical Analysis 2013, 47, 1553–1582.
  • Laestadius and Kvaal 2018 Laestadius, A.; Kvaal, S. Analysis of the extended coupled-cluster method in quantum chemistry. SIAM Journal on Numerical Analysis 2018, 56, 660–683.
  • Laestadius and Faulstich 2019 Laestadius, A.; Faulstich, F. M. The coupled-cluster formalism–a mathematical perspective. Molecular Physics 2019, 117, 2362–2373.
  • Faulstich et al. 2019 Faulstich, F. M.; Laestadius, A.; Legeza, Ö.; Schneider, R.; Kvaal, S. Analysis of the tailored coupled-cluster method in quantum chemistry. SIAM Journal on Numerical Analysis 2019, 57, 2579–2607.
  • Hassan et al. 2023 Hassan, M.; Maday, Y.; Wang, Y. Analysis of the single reference coupled cluster method for electronic structure calculations: the full-coupled cluster equations. Numerische Mathematik 2023, 155, 121–173.
  • Hassan et al. 2023 Hassan, M.; Maday, Y.; Wang, Y. Analysis of the Single Reference Coupled Cluster Method for Electronic Structure Calculations: The Discrete Coupled Cluster Equations. arXiv:2311.00637 2023,
  • Crawford and Schaefer III 2007 Crawford, T. D.; Schaefer III, H. F. An introduction to coupled cluster theory for computational chemists. Reviews in computational chemistry 2007, 14, 33–136.
  • 71 Grayson, D. R.; Stillman, M. E. Macaulay2, a software system for research in algebraic geometry. Available at http://www2.macaulay2.com.
  • Breiding and Timme 2018 Breiding, P.; Timme, S. HomotopyContinuation. jl: A package for homotopy continuation in Julia. Mathematical Software–ICMS 2018: 6th International Conference, South Bend, IN, USA, July 24-27, 2018, Proceedings 6. 2018; pp 458–465.
  • Breiding et al. 2024 Breiding, P.; Kohn, K.; Sturmfels, B. Metric algebraic geometry; Springer Nature, 2024.
  • Bates et al. 2023 Bates, D. J.; Breiding, P.; Chen, T.; Hauenstein, J. D.; Leykin, A.; Sottile, F. Numerical Nonlinear Algebra. arXiv:2302.08585 2023,
  • Garcia and Zangwill 1979 Garcia, C. B.; Zangwill, W. I. Finding all solutions to polynomial systems and other systems of equations. Mathematical Programming 1979, 16, 159–176.
  • Morgan 2009 Morgan, A. Solving polynomial systems using continuation for engineering and scientific problems; SIAM, 2009.
  • Sommese et al. 2005 Sommese, A. J.; Wampler, C. W., et al. The Numerical solution of systems of polynomials arising in engineering and science; World Scientific, 2005.
  • Davidenko 1953 Davidenko, D. On a new method of numerical solution of systems of nonlinear equations. Proceedings of the USSR Academy of Sciences. 1953; pp 601–602.
  • Davidenko 1953 Davidenko, D. On the approximate solution of systems of nonlinear equations. Ukrainian Mathematical Journal 1953, 5, 196–206.
  • Morgan et al. 1992 Morgan, A. P.; Sommese, A. J.; Wampler, C. W. Computing singular solutions to polynomial systems. Advances in Applied Mathematics 1992, 13, 305–327.
  • Verschelde 1999 Verschelde, J. Algorithm 795: PHCpack: A general-purpose solver for polynomial systems by homotopy continuation. ACM Transactions on Mathematical Software (TOMS) 1999, 25, 251–276.
  • Bates et al. 2006 Bates, D. J.; Hauenstein, J. D.; Sommese, A. J.; Wampler, C. W. Bertini: Software for numerical algebraic geometry. 2006.
  • Chen et al. 2014 Chen, T.; Lee, T.-L.; Li, T.-Y. Hom4PS-3: a parallel numerical solver for systems of polynomial equations based on polyhedral homotopy continuation methods. Mathematical Software–ICMS 2014: 4th International Congress, Seoul, South Korea, August 5-9, 2014. Proceedings 4. 2014; pp 183–190.
  • Lee et al. 2008 Lee, T.-L.; Li, T.-Y.; Tsai, C.-H. HOM4PS-2.0: a software package for solving polynomial systems by the polyhedral homotopy continuation method. Computing 2008, 83, 109–133.
  • Van Voorhis and Head-Gordon 2000 Van Voorhis, T.; Head-Gordon, M. Benchmark variational coupled cluster doubles results. J. Chem. Phys. 2000, 113, 8873–8879.
  • Sun et al. 2020 Sun, Q.; Zhang, X.; Banerjee, S.; Bao, P.; Barbry, M.; Blunt, N. S.; Bogdanov, N. A.; Booth, G. H.; Chen, J.; Cui, Z.-H., et al. Recent developments in the PySCF program package. The Journal of chemical physics 2020, 153.
  • Sun et al. 2018 Sun, Q.; Berkelbach, T. C.; Blunt, N. S.; Booth, G. H.; Guo, S.; Li, Z.; Liu, J.; McClain, J. D.; Sayfutyarova, E. R.; Sharma, S., et al. PySCF: the Python-based simulations of chemistry framework. Wiley Interdisciplinary Reviews: Computational Molecular Science 2018, 8, e1340.
  • Sun 2015 Sun, Q. Libcint: An efficient general integral library for g aussian basis functions. Journal of computational chemistry 2015, 36, 1664–1671.
  • Bulik et al. 2015 Bulik, I. W.; Henderson, T. M.; Scuseria, G. E. Can single-reference coupled cluster theory describe static correlation? Journal of chemical theory and computation 2015, 11, 3171–3179.
{suppinfo}

7 Dissociating systems of hydrogen

We begin by reporting the full energy spectrum of the Hamiltonian describing (H2)2 in D2h symmetry together with all CCD roots, see Figure 13(a). Given the large number of solutions, we minimize over the difference between spectral values of the Hamiltonian and the found CCD energies. This provides further insight into energies that are well approximated by CCD, see Figure 13(b).

Refer to caption
(a)
Refer to caption
(b)
Figure 13: (a) Full spectrum (b) Minimal difference.

Similarly, we proceed for (H2)2 in D∞h symmetry. We reporting the full energy spectrum of the Hamiltonian describing (H2)2 in D∞h symmetry together with all CCD roots, see Figure 14(a) Given the large number of solutions, we minimize over the difference between spectral values of the Hamiltonian and the found CCD energies. This provides further insight into energies that are well approximated by CCD, see Figure 14(b).

Refer to caption
(a)
Refer to caption
(b)
Figure 14: (a) Full spectrum (b) Minimal difference.

8 H4 disturbed on a circle

Similarly, we proceed for H4 disturbed on a circle. We report the full energy spectrum of the Hamiltonian describing H4 disturbed on a circle together with all CCD roots, see Figure 15(a). Given the large number of solutions, we minimize over the difference between spectral values of the Hamiltonian and the found CCD energies. This provides further insight into energies that are well approximated by CCD, see Figure 15(b).

Refer to caption
(a)
Refer to caption
(b)
Figure 15: (a) Full spectrum (b) Minimal difference.

9 Lithium hydride

We report the full energy spectrum of the Hamiltonian describing LiH together with all CCD roots, see Figure 16(a) Given the large number of solutions, we minimize over the difference between spectral values of the Hamiltonian and the found CCD energies. This provides further insight into energies that are well approximated by CCD, see Figure 14(b).

Refer to caption
(a)
Refer to caption
(b)
Figure 16: (a) Full spectrum (b) Minimal difference.