A .A 4 .. r... .fiflufi. r5 Irma; .r a; a z“ warn ‘ u}. i a . q 2 . Rh: 3.: z 1%.. v ... :11: 1 a.“ , ¥ :4: f\«’ '9 in I u. , :3. 1;! v.u. 1.x! I: Rheum“ 32.2."... . an...) :1 . call I .2 If. , 41‘ A ¥ ,.fi....a.suw...3._ .J. 3 . . . . .73 u. . . 5E3. ..-.\.; LIBRARIES MICHIGAN STATE UNIVERSITY EAST LANSING, MICH 48824-1048 This is to certify that the dissertation entitled NEW ALTERNATIVES FOR ELECTRONIC STRUCTURE THEORY: THE APPLICATION OF TWO-BODY CLUSTER EXPANSIONS IN HIGH ACCURACY AB INITIO CALCULATIONS presented by Peng-Dong Fan has been accepted towards fulfillment of the requirements for the PhD. degree in chemistry Qlél/ GOOQA Maio'r Professor‘s Signature May 2. 2005 Date MSU is an Afiinnative Action/Equal Opportunity Institution PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE (.33 2 3 2007 AUG 2 9 2007 m NEW ALTERNATIVES FOR ELECTRONIC STRUCTURE THEORY: THE APPLICATION OF TWO-BODY CLUSTER EXPAN SION S IN HIGH ACCURACY AB IN ITIO CALCULATIONS By Peng-Dong Fan A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2005 ABSTRACT NEW ALTERNATIVES FOR ELECTRONIC STRUCTURE THEORY: THE APPLICATION OF TWO-BODY CLUSTER EXPANSIONS IN HIGH ACCURACY AB INITIO CALCULATIONS By Peng-Dong Fan In this thesis, the applicability of exponential cluster expansions involving one- and two-electron operators in high accuracy ab initio calculations is discussed. First, the extended coupled-cluster method with singles and doubles (ECCSD) is tested in the most demanding studies of potential energy surfaces involving multiple bond breaking. The numerical results for a few cases of multiple bond breaking show that the single-reference ECCSD method is capable of providing a qualitatively correct description of entire potential energy surfaces, eliminating, in particular, the fail- ures and the unphysical behavior of all standard coupled-cluster methods in similar cases. It is also demonstrated that one can obtain entire potential energy surfaces with millihartree accuracies by combining the ECCSD theory with the noniterative a posteriori corrections obtained by using the generalized variant of the method of moments of coupled-cluster equations. This is the first time when the relatively sim- ple single-reference formalism, employing one— and two-body clusters only, provides a highly accurate description of the dynamic and significant nondynamic correlation effects characterizing multiply bonded systems. Second, an evidence is presented that one may be able to represent the exact or virtually exact ground- and excited-state wave functions of many-electron systems by exponential cluster expansions employing general two-body operators. Calculations for small many-electron systems indicate the existence of finite two-body parameters that produce the numerically exact wave functions. This finding may have a significant impact on future quantum calcula- tions for many-electron systems, since normally one needs triply excited, quadruply excited, and other higher—than—doubly excited Slater determinants, in addition to all singly and doubly excited determinants, to obtain the exact or virtually exact wave functions. ACKNOWLEDGMENT I feel very privileged to have worked with my supervisor, Dr. Piotr Piecuch. To him I owe a great debt of gratitude for his patience, inspiration, and the warmest support. Piotr has taught me a great deal about electronic structure theory and guided me to explore new alternatives in this field. I also owe a great debt of gratitude to my Guidance Committee, including Dr. Katharine C. Hunt, Dr. Paul F. Mantica, and Dr. James K. McCusker, for advice, support, and setting up and overseeing my graduate study program. I would also like to thank Dr. Karol Kowalski, a former postdoctoral research asso- ciate in our group, who shared the results of his work when I started working on this project. I would like to thank Ian, Mike, Ruth, DJ, and Maricris for their friendship and help in practicing my seminar presentations. Finally, I would like to acknowledge the Department of Chemistry for providing the teaching assistantship and the Dissertation Completion Fellowship, and the US De- partment of Energy for the research assistantship. iv Table of Contents List of Figures vi List of Tables vii 1 Introduction 1 2 Practical Ways of Improving Standard Coupled-Cluster Methods Employing Singly and Doubly Excited Clusters via Extended Coupled- Cluster Theory 5 2.1 Extended Coupled-Cluster Theory: General Formalism ........ 5 2.2 Extended Coupled-Cluster Methods with Singles and Doubles . . . . 9 2.2.1 The Piecuch-Bartlett ECCSD Approach ............ 9 2.2.2 The Arponen—Bishop ECCSD Approach ............ 12 2.3 Numerical Results for Multiple Bond Breaking ............. 17 2.3.1 The Piecuch-Bartlett ECCSD Approach ............ 17 2.3.2 The Arponen-Bishop ECCSD Approach ............ 34 2.4 Conclusion ................................. 41 3 Noniterative Corrections to Extended Coupled-Cluster Energies for High Accuracy Electronic Structure Calculations: Generalized Method of Moments of Coupled-Cluster Equations 43 3.1 The Standard Method of Moments of Coupled-Cluster Equations . . 44 3.2 The Generalized MMCC Formalism ................... 50 3.3 The ECCSD(T), ECCSD(TQ), QECCSD(T), and QECCSD(TQ) Meth- 3.4 ods and their Performance in Calculations for 'Il'iple Bond Breaking in 4 Exactness of Two-Body Cluster Expansions in Many-Body Quantum Theory 70 4.1 Theory ................................... 71 4.1.1 The exp(X) Conjecture ..................... 72 4.1.2 Formal Arguments in Favor of the exp(X) Conjecture (Ground States) ............................... 75 4.1.3 Extension of the exp(X) Conjecture to Excited States ..... 81 4.2 Numerical Results ............................. 82 4.3 Summary ................................. 89 Bibliography 90 2.1 2.2 2.3 2.4 2.5 3.1 3.2 List of Figures Potential energy curves for the MBS S4 system. ECCD represents the Piecuch-Bartlett variant of the ECCD approach (in this case, the ECCD and ECCSD approaches are equivalent). ............ Potential energy curves for the N2 molecule, as described by the STO- 3G basis set. The LECCSD, BECCSD, QECCSD, and ECCSD meth- ods are based on the Piecuch-Bartlett variant of the ECC theory. (a) The overlaps of the normalized CCSD, BECCSD, QECCSD, and ECCSD wave functions, Eq. (2.62), with the normalized full CI wave function for the STO-3G N2 molecule, as functions of the N—N separa- tion R (in bohr). (b) The difference between the CCSD and ECCSD cluster operators T and the difference between the ECCSD cluster op- erators T and E, as defined by the quantity d(Y, Z), Eq. (2.63), for the STO-3G N2 molecule, as functions of the N—N separation R (in bohr). Ground-state potential energy curves of the N2 molecule, as described by the STO-3G basis sets. The QECCSD and ECCSD methods are based on the Arponen-Bishop ECC theory. .............. Ground-state potential energy curves of the N2 molecule, as described by the DZ basis sets. QECCSD represents the quadratic version of the ECCSD theory of Arponen and Bishop .................. Ground-state potential energy curves of the N2 molecule as described by the STO-3G basis set. The QECCSD and ECCSD calculations were performed using the ECC theory of Arponen and Bishop ........ Ground-state potential energy curves of the N2 molecule as described by the DZ basis set. The QECCSD calculations were performed using the ECC theory of Arponen and Bishop. ................ vi 19 26 32 38 40 68 69 2.1 2.2 2.3 2.4 2.5 3.1 3.2 4.1 4.2 List of Tables Ground-state energies of the MBS S4 model system as functions of the parameter a. ............................... Ground-state energies of the N 2 molecule, as described by the STO-3G basis set ................................... Ground-state energies of the N2 molecule, as described by the STO-3G basis set. ................................. The ground-state energies of the N2 molecule obtained for several in- ternuclear separations R with the STO-3G basis set. ......... The ground-state energies of the N2 molecule obtained for several in- ternuclear separations R with the DZ basis set. ............ The ground-state energies of the N2 molecule obtained for several in- ternuclear separations R with the STO-3G basis set. All ECC-related calculations were performed with the Arponen-Bishop ECC theory. The ground-state energies of the N2 molecule obtained for several in- ternuclear separations R with the DZ basis set. ............ A comparison of the ground-state energies of the MBS H8 system ob- tained with the exp(X)-like wave functions, where X = X2 (a purely two-body operator) or X1 + X2 (a sum of one- and two-body opera- tors), with the exact, full CI, energies and energies obtained in various CI and CC calculations. Full CI energies are total energies in hartree and all other energies are errors relative to full CI (also in hartree). We also give the overlaps of the normalized ground-state wave functions obtained with the exp(X) ansatz and~the full CI approach. The over- laps of the reference states Km) and |o) with the full CI ground-state wave function |\IIO) are given for comparison ............... Same as Table 4.1 for the first excited state of the 1Ag symmetry. vii 25 87 88 Chapter 1 Introduction Great advances have been made in ab initio quantum chemistry. Highly accurate calculations for closed-shell and simple open-shell molecular systems involving a few atoms are nowadays routine. This, in particular, applies to coupled-cluster (CC) theory1—5, which has become the de facto standard for high accuracy calcu- lations for atomic and molecular systems‘3—13. The basic single-reference CC meth- ods, such as CCSD (CC approach with singles and doubles)”, and the noniterative CCSD + T(CCSD) = CCSD[T]l5 and CCSD(T)16 approaches that account for the effect of triexcited clusters by using arguments based on the many-body perturbation theory (MBPT), in either the spin-orbitall4"16 and spin-freer"-19 or spin-adapted'm—22 forms, are nowadays routinely used in accurate electronic structure calculations. The idea of adding the a posteriori corrections due to higher-than—doubly excited clusters to CCSD energies, on which the CCSD[T] and CCSD(T) approaches and their more recent CCSD(TQf) extension23 are based, is particularly attractive, since it leads to methods that offer an excellent compromise between high accuracy and relatively low computer cost, as has been demonstrated over and over in numerous atomic and molecular applications7—10'12’13. There are, however, Open problems in CC theory. First and foremost is the per- vasive failing of the standard single-reference CC methods, such as CCSD, CCSD[T], CCSD(T), and CCSD(TQf), at larger internuclear separations, when the spin-adapted restricted Hartree-Fock (RHF) configuration is used as a reference, which limits the applicability of the standard CC approaches to molecules near their equilibrium ge- ometries. Second is the large computer effort associated with the need for using higher-than—doubly excited clusters in calculations involving quasi-degenerate elec- tronic states and bond breaking when larger many-electron systems are examined. Undoubtedly, it would be very useful to extend the applicability of the standard single-reference CC methods to entire molecular potential energy surfaces (PESs) involving bond breaking, and quasi-degenerate electronic states in general, without invoking complicated and time-consuming steps associated with the more traditional multi-reference approaches, in which one has to choose active orbitals and multi- dimensional reference spaces on an ad hoc molecule—by—molecule basis. Ideally, one would like to develop a straightforward theory which could provide a virtually exact description of many-electron wave functions with the exponential, CC-like, expansions involving one- and two-body operators only, since the molecular electronic Hamilto- nian does not include higher-than—two-body interactions. There are several specific challenges in all those areas. First of all, the RHF-based CCSD method, on which the noniterative CCSD[T], CCSD(T), and CCSD(TQf) approaches are based, is inadequate for the description of bond breaking and quasi- degenerate states, since it neglects all higher—than—doubly excited clusters, including the important triply and quadruply excited T3 and T4 components. Second, the non- iterative triples and quadruples corrections defining the CCSD[T], CCSD(T), and CCSD(TQf) methods aggravate the situation even further, since the usual arguments originating from MBPT, on which these standard noniterative CC approximations are based, fail due to the divergent behavior of the MBPT series at larger inter- nuclear separations (or when the strong configurational quasi-degeneracy and large nondynamic correlation effects set in). In consequence, the ground-state PESs ob- tained with the CCSD[T], CCSD(T), CCSD(TQf), and other noniterative CC ap- proaches are completely pathological when the RHF configuration is used as a ref- erence (cf., e.g., Refs. 9, 11—13, 24—37 and references therein). The iterative ex- tensions of the CCSD[T], CCSD(T), and CCSD(TQf) methods, including, among many others, the CCSDT-112135“41 and CCSDTQ-l42 approaches, and the nonitera- tive CCSDT + Q(CCSDT) = CCSDT[Q]42 and CCSDT(Q;)23 methods, in which the a posteriori corrections due to T4 cluster components are added to the full CCSDT 43"“ energies, improve the description of PBS in the (CC singles, doubles, and triples) bond breaking region, but ultimately all of these approaches break down at larger internuclear distances (see, e.g., Refs. 28, 29, 34). One might try to resolve the fail- ures of the standard single-reference CC approaches in the bond breaking region and for quasi-degenerate electronic states in a brute-force manner by including the triply excited, quadruply excited, pentuply excited, etc. clusters fully and in a completely iterative fashion, but, unfortunately, the resulting CCSDTQ (CC singles, doubles, triples, and quadruples)45_48, CCSDTQP (CC singles, doubles, triples, quadruples, and pentuples)”, etc. approaches are far too expensive for routine molecular appli- cations. For example, the full CCSDTQ method requires iterative steps that scale as ngnfi (no (nu) is the number of occupied (unoccupied) orbitals in the molecular orbital basis). This N10 scaling with the system size (N) restricts the applicability of the CCSDTQ approach to very small systems, consisting of ~ 2 - 3 light atoms de- scribed by small basis sets. For comparison, CCSD(T) is an ngnfi (or JV“) procedure in the iterative CCSD steps and an 723123 (or M) procedure in the noniterative part related to the calculation of the triples (T) correction. In consequence, it is nowadays possible to perform the CCSD(T) calculations for systems with 10—20 light atoms or a few heavier (transition metal) atoms. This indicates that in searching for new methods that would help to overcome the failures of the standard CC approaches in the bond breaking region, one should focus on the idea of improving the results of the low-order CC calculations, such as CCSD, with the noniterative corrections of the CCSD(T) type, since only such methods have a chance to be applied to larger molecular systems in the not—too—distant future. In view of the above discussion, the question: Can one improve the quality of standard CC wave functions in the bond breaking region at the basic CCSD level of the single-reference CC theory? seems to be particularly important. In this thesis, we show that this can be accomplished by exploring the so-called extended coupled- cluster (ECC) theory. The basic ECCSD results, particularly when multiple bonds are stretched or broken, are qualitatively much better than the corresponding standard CCSD results. However, they are not yet fully quantitative. This prompts another question: Can one improve the quality of the ECCSD results by adding a simple a posteriori correction to the ECCSD energy which is obtained by using the singly and doubly excited cluster amplitudes obtained with the ECCSD approach? In this thesis, 'we show that the answer to this question is affirmative if we use the generalized method of moments of coupled-cluster equations (GMMCC). Eventually, of course, one would prefer to use only one- and two-body clusters to obtain an exact or virtually exact description of many-electron systems, since, as we have already mentioned, the Hamiltonians used in quantum chemistry do not contain higher-than—pairwise interactions. This prompts the third and the final question of this thesis research: Can one obtain the exact or virtually exact many-electron wave functions by using two-body exponential cluster expansions? Chapter 2 Practical Ways of Improving Standard Coupled-Cluster Methods Employing Singly and Doubly Excited Clusters via Extended Coupled-Cluster Theory 2.1 Extended Coupled-Cluster Theory: General Formalism The extended coupled-cluster (ECC) theory is based on the asymmetric, doubly con- nected energy functional50—60, Elm) =<Ifil<1>>, (2.1) where |) is the independent particle model reference configuration (e.g., the Hartree- Fock determinant) and H = e2t(e'THeT)e’2t = eZtHe’El = (Eff-[)6 = [eEt(HeT)C]C (2.2) is the doubly transformed Hamiltonian, obtained by transforming the similarity trans- ‘ formed Hamiltonian H used in the standard CC theory, H = e-THeT = (HeT)C, (2.3) where H is the Hamiltonian and C stands for the connected part of the corresponding operator expression, with the exponential operator e'Et. T is the usual cluster opera- tor, which is a particle-hole excitation operator generating the connected components of the many-electron ground-state wave function I‘I/o) = 871(1)). (24) 5 and BI is the auxiliary hole-particle deexcitation operator. In the exact theory, T is a sum of all many-body components Tn with n = 1,. . . , N, N T = Z Tn, (2 5) n=1 where N is the number of electrons and T... is defined as _ i1...ifl a1...an Tn" Z ta1°"anEi1---in , (2.6) i1<~-) and a1, . . . ,an designate the unoccupied spin-orbitals. The auxiliary operator 21 is defined as N 2* = Z 231,, (2.8) n=1 where :31, = . Z 0511.1".an13311'33fi'n. (2.9) 0 Zn The operators E31221?" = II Cancun (2.10) K, = 1 . a . . . a . . . and the coefficrents Ui 1. . n are the corresponding deexc1tatlon operators and am- 1 o 0 Zn plitudes, respectively. The operators T and 2* (or the corresponding amplitudes till 1211.3" and 02911.1... ia") 78 defining the wave function I‘llo) through Eq. (2.4) and the energy EéECC) through Eq. (2.1) can be determined in various ways. In the ECC theory of Piecuch and Bartlett“, which can be applied to both ground and excited states, the T and El operators are determined by considering the doubly transformed form of the electronic Schrodinger equation, i.e., line) = E0|), (2.11) and its left-hand analog, (<1)le Eo(|, (2.12) where (| is the left eigenstate of II corresponding to the right reference eigenstate |). We obtain Eq. (2.11) by inserting the formula for the CC wave function, Eq. (2.4), into the Schrbdinger equation, Hl‘I’o> = Eol‘I’ola (2-13) and by premultiplying the resulting equation by eve—T, while utilizing the fact that |\Ilo) = eT|) = eTe‘EIIQ) (23t is a deexcitation operator, so that (El)"|<1>) = 0 for k > 0). In general, the (| dual state entering Eq. (2.12) depends on the values of T and El defining R. In the ECC theory of Piecuch and Bartlett, we simply require that T and El are such that (| = (|. Thus, the final system of equations used to determine the two different cluster operators T and El consists of Eq. (2.11) and (<12le = <|Eo. (2.14) which is the left-hand counterpart of Eq. (2.11). It is worth mentioning that Eq. (2.14) can also be obtained by considering the bra counterpart of the connected cluster form of the electronic Schréidinger equation, chp) = E0|), (2.15) where H is defined by Eq. (2.3), i.e. (|(1+ ME = E0(|(1+ A), (2.16) where A is the well-known “lambda” deexcitation operator of the analytic gradient CC theorym'”, and by identifying the left-hand ground eigenstate of H, (|(l + A), with (|ezl. An alternative reasoning that leads to Eq. (2.14) is based on considering the adjoint form of the electronic Schrodinger equation, i.e. (‘i’olH = Eo<‘i’o|, (2.17) where the dual state (\Ilol, satisfying the condition (\Ilolqlo) = 1, is the CC bra ground statemm’“, (\iIOI = (<1>|(1 + A)e"T. (2.18) Clearly, in the exact, full C1 or full CC, limit, there exists a deexcitation operator U, such that (1 + A) = e”, so that one can always give the dual CC state (Ilol a completely bi-exponential form, (to) = (slave-T. (2.19) By inserting Eq. (2.19) into Eq. (2.17) and by multiplying the resulting equation on the right by eTe‘Sl, we obtain the desired Eq. (2.14). In the original work by Arponen and Bishopm—59, the operators T and 21 of the ECC formalism are determined by imposing the stationary conditions on the energy functional ESECC), Eq. (2.1), with respect to operators T and 21, 6E3ECC) 6T 6 EéECC) 62" =0, =0. (2%) The bi-variational character of the ECC theory of Arponen and Bishop is particularly useful in calculations of molecular properties other than energy, since one can apply 8 the Hellmann-Feynmann theorem in such calculations‘55‘73. The question addressed in this thesis is how the bi-variational ECC theory of Arponen and Bishop and the ECC theory of Piecuch and Bartlett, which uses Eqs. (2.11) and (2.14) rather than Eq. (2.20) to determine T and 2", work when molecular PESs along bond breaking coordinates are examined. 2.2 Extended Coupled-Cluster Methods with Singles and Dou- bles 2.2.1 The Piecuch-Bartlett ECCSD Approach The approximate ECC methods, such as the ECCSD approaches tested in this work and developed in Refs. 74—77, are obtained as follows: First, as in all standard CC approximations, we truncate the many-body expansions of T and El, Eqs. (2.5) and (2.8), respectively, at some excitation level mA < N, so that T is replaced by T“), mA T“) = 2T", (2.21) n=1 and 2" is replaced by mA 2”” = Z 21,. (2.22) n=l Next, we use either the Piecuch-Bartlett approach (Eqs. (2.11) and (2.14)) or the Arponen-Bishop approach (Eq. (2.20)) to obtain a system of equations for the un- ”in ’ Z. a a! o o o a known cluster amplitudes ta1 (1 and 0-1 . ", l ’ ' ’ n 21 , _ 2.1.0.2.” ”Zn n — 1,...,mA. Once tal"'an and 0?". ° 1“ n are determined, we use the approximate form of the energy functional 1 .02" (2.1), 35,") = (cplfiwlcp), (2.23) where HM) ___ 82( ) (8_T(A)H6T(A))e-E(A) to calculate the ground-state energy. (2.24) In the specific case of the ECCSD approach, T is approximated by the sum of one- and two-body components, T1 and T2, respectively, where and with and T g T(ECCSD) = T1 + T2, i a = 2228 i,a __ ab T2 — Z tajbEij ’ i<fia|<1>) = 0, (2.33) <¢gjblfiaccsml¢> = 0, i < j, a < b, (2.34) (olfimsmlcg‘) = 0, (2.35) (¢|RIECCSD)|§‘JI-’) = 0, 2' < j, a < b, (2.36) where 111230050) = 621”); (e—Tl—T2 HeT1+T2) 6—21—23 = e£l+2lH(CCSD>e-2l-El (2.37) is the doubly transformed Hamiltonian of the ECCSD method, with H(CCSD) = €_T1_T2H€TI+T2 = (HeT1+T2)C (238) representing the similarly transformed Hamiltonian of the CCSD approximation. Once Eqs. (2.33)-(2.36) are solved for operators T1, T2, 21 and 2;, the ground-state energy ESECCSD) is calculated as follows: ESECCSD) = ((1,, fiscasmlé) : ((plezl-I-Eg (e—Tl —T2 H8T1+T2)€_EI_E;|¢) = (q)|e’3l+zi(e‘Tl‘T2HeT1+T2)|), (2.39) 11 where we used the fact that e‘zllQ) = IQ). Since the full ECCSD formalism defined by Eqs. (2.33)—(2.39) is rather complex, in this thesis, in addition to the full ECCSD method, we consider the systematic sequence of the linear (LECCSD), bi-linear (BECCSD), and quadratic (QECCSD) approaches, which represent approximations to full ECCSD. The LECCSD, BECCSD, and QECCSD methods are obtained by replacing the doubly transformed Hamiltonian 1330080), Eq. (2.37), in Eqs. (2.33)—(2.36) and (2.39) by ITLECCSD) = (1 + 21 + EDI-{(CCSDKI — El — 2;), (2.40) in the LECCSD case, 13000009 = (1 + 2) + 293000041 _ 21 _ 2;) +%(21 + 2;)2H(CCSD) + %H(CCSD)(EI + 2;)2’ (2.41) in the BECCSD case, and = 1 _ 91000009 = [1 + 2; + 1:; + 5(2:) + 2.).)0] 140001» 1 x [1 — )3} — 2; + 5(2)} + 2})2] , (2.42) in the QECCSD case. 2.2.2 The Arponen-Bishop ECCSD Approach In the case of the ECCSD method of Arponen and Bishop50—59, we obtain the system of equations for the singly and doubly excited cluster amplitudes t2 and thjb defining T1 and T2 and the deexcitation amplitudes 09 and of}? defining El and 2* from the z z] I 2 stationary condition represented by Eq. (2.20). The resulting equations can be given in the following form: BEéECCSD) 80“ = (leezl‘nglL-I‘CCSDHQ) = 0, (2.43) i 12 aE(ECCSD) b t f _ _0__ = ((1)3. [621+22H(CCSD)|) = 0, i < j, a < b, (2.44) 3029‘]? ‘7 (ECCSD) 6E0 . : <¢|ezl+zg[g(ccsn)’ qu» 3th I I — = (QIeElHSZ (H(CCSD) E?)C|q)) : 0, (2.45) (ECCSD) 6E0 .. = (@l €21.02; [ H(CCSD) EQbHQ) at” ’ 1.? ab 1' _ = (0|e2l+52 (H‘CCSDlEng’)C|) = 0, i< j, a < b, (2.46) where EéECCSD) is the ECCSD energy functional, Eq. (2.39), and H|(1+ 21 + 292003210) = (012003000) + (0211110039009 + <0I21H‘CCSD’I0> : <¢|H(CCSD)|¢) + Z 0.76:1. <¢?|H(CCSD)I¢> i, a ab ab ‘ (CCSD) + Z a” (<1,le |Q). (2.47) i < j,a < b The most expensive steps of the LECCSD method, as defined by Eq. (2.47), scale as ngnfi (or N6 with the system size), but, unfortunately, the LECCSD method based 13 on the Arponen-Bishop theory does not improve the CCSD results at all, since once we impose the stationary conditions on the energy functional ESLECCSD) with respect to operators T1, T2, 2}, and 2;, we obtain the following equations: aE(Lsccsn) _ 334 = <0ilHI> = o. (2.48) z E(LECCSD) __ L3)?»— = <0ijblH‘CCSD’l0> = 0. .-< j, a < 0, (2-49) which are the usual CCSD equations for T1 and T2, and aEmsccsn) _ JEF— = <l(1 + 21 + £1) (H‘CCSD’ 13,9980) = 0. (2.50) a aE(LECCSD) _ 05%,, = <|(1 + >31 + 2;) (H‘CCSD’ E#)c|<1>> = 0. ab i < j, a < b, (2.51) which are the El equations that are completely decoupled from the CCSD system, Eqs. (2.48) and (2.49), and solved only after the T1 and T2 cluster are determined. In other words, the LECCSD approximation does not provide for the coupling between the T and El equations, which is necessary for improving the quality of T1 and T 2 clusters in the bond breaking region. The above analysis indicates that in order to obtain the T1 and T2 clusters, which are better than those provided by the LECCSD = CCSD approximation, we must introduce nonlinear terms in 21 and 23; into the LECCSD energy functional that couple the T and 2" equations. The lowest-order approximation of this type is ob- tained by truncating the ECCSD energy functional, Eq. (2.39), at terms quadratic in (2} + 5.3;). We call it the QECCSD approximation. The formula for the QECCSD energy functional used in the Arponen-Bishop version of the QECCSD theory is 1 _ 123000000) = (010 + (21 + >31) + 5(21 + 292111208210) 14 = <0IHI0> + <¢|ElH‘eeSD’l> 1 _ +<I12£+ 5(212111‘0052109 +<Izizlfilccsml0> + §<I(>:;)2HI> (2.52) or, somewhat more explicitly, E(QECCSD)_ (q, I H(ccsn)|¢) + Z 030%” 1:1(ccsn) I‘D) i, a + Z (aijb + Aijaa 0?)( i W|H(CCSD>|¢)= (2.54) j < k, b < c 0E0 E(QECCSD) ab = (Qajb|H(CCSD)Iq)> + 20k<¢§3 bC|H(CCSD)|(p> 60 Uij k, c + Z ag‘l’(<1>al’cd|H||[1+(EI+E§ )+— %(2{+2;) 2](H CCSD)EC‘)C |> ati = (¢I(H(CCSD)E?)C|) + 20?. ((D2|(H(CCSD)EZQ)CI(I)> 1311 + Z (a ek+Aka jag)(<1>gg|(HE3)C|¢) j|[1 + (21 + 2;) + E(>31 + 21.)?) (HEgj )C|q>) atab = <|(H‘CCSD’E,‘-‘}’)cld>) + Zag(<1>il(H‘CCSD)E3})C|q>) k,c + Z (agghAklogaquWduH (CCSD)Eab)C (<5) k < l, c < d 0031) ab + Z Ak/lm'A c/de‘7k0'zm(‘I’ic‘zi ml(H( )Ei )0 VP) k < l < m c) k —2.00 '5 5 —2.02 > . 9 -2.04 3:: ’ ACCD -2-°6 _' VECCD - -2.08 _ OFullCl _2.1O . 1 . n . n . 1 . 2.0 3.0 4.0 5.0 6.0 7.0 0c(bohr) Figure 2.1: Potential energy curves for the MBS S4 system. ECCD represents the Piecuch-Bartlett variant of the ECCD approach (in this case, the ECCD and ECCSD approaches are equivalent). 19 The MBS S4 models with larger values of a create a serious challenge for the standard single-reference CC methods (even the genuine MRCC approaches may have a difficulty in describing these models-’8'“). This is related to the fact that larger a values correspond to a dissociation of the 84 model into four hydrogen atoms. This process is very difficult to describe by the RHF-based single-reference methods, particularly in the region of the intermediate a values where the ground-state wave function of the S4 system undergoes a significant rearrangement of its structure. As shown in Table 2.1 and Figure 2.1, the CCD = CCSD approximation breaks down at larger values of a. For a S 2.0 bohr, the unsigned errors in the CCD results, relative to full CI, do not exceed 2.834 millihartree. However, for 3.5 bohr 5 a g 4.5 bohr, the unsigned errors in the CCD results increase to 13—14 millihartree and the CCD potential energy curve corresponding to the dissociation of the S4 model into four hydrogen atoms goes significantly below the exact, full CI, curve (see Figure 2.1). For a g 2.0 bohr, the ground-state wave function of the MBS S4 model is dominated by two configurations: the RHF determinant and the doubly excited determinant corresponding to the HOMO —+ LUMO biexcitation78-80. For larger values of a, essentially all electron configurations present in the full CI expansion become very important, creating a strongly quasi-degenerate situation, and the role of the T4 cluster component becomes significant in the intermediate 2.5 bohr S a S 6.0 bohr region. This leads to the failure of the CCD approximation observed in Table 2.1 and Figure 2.1. 20 Table 2.1.: Ground-state energies of the MBS S4 model system as functions of the parameter a.“ a Full CI CCD b ECCD c (CCD)d (ECCD)d 0.5 3.952114 —0.093 0.015 0.032 0.015 1.0 —0.668783 —0.424 0.110 0.236 0.110 1.5 -1.694327 —1.235 0.376 0.862 0.376 2.0 —1.975862 —2.834 0.871 2.188 0.871 2.5 —2.043797 —5.674 1.609 4.447 1.609 3.0 —2.044850 —9.568 2.373 6.895 2.373 3.5 —2.028186 —13.083 2.754 7.698 2.754 4.0 —2.011713 —14.385 2.496 6.218 2.496 4.5 —2.000438 -13.094 1.812 3.896 1.812 5.0 —1.994021 —10.292 1.107 2.054 1.107 6.0 -1.989199 -4.748 0.297 0.431 0.296 7.0 —1.988164 —1.743 0.062 0.076 0.061 “The full CI total energies are in hartree. The CCD and ECCD energies are in millihartree relative to the corresponding full CI energy values. The parameter a is in bohr. ”For the MBS 84 model, CCD = CCSD. cThe ECCD method of Piecuch and Bartlett. For the MBS S4 model, ECCD = ECCSD = BECCSD = QECCSD. d(X) (X = CCD, ECCD) is the expectation value of the Hamiltonian with the 6T2 |) wave function, where T2 is obtained with method X (see Eq. (2.58)). 21 The results in Table 2.1 and Figure 2.1 show that the ECCD theory provides substantial improvements in the poor description of the dissociation of the S4 model into four hydrogen atoms by the CCD method. The 13-14 millihartree unsigned errors in the CCD energies in the 3.5 bohr g a g 4.5 bohr region reduce to 2-3 millihartree, when the ECCD approach is employed. The considerable reduction of errors in the CCD results is observed at all values of (1, even in the a S 2.0 bohr region, where the maximum unsigned error in the ECCD results is 0.871 millihartree, as opposed to 2.834 millihartree obtained with the CCD approach (see Table 2.1). Unlike the CCD potential energy curve shown in Figure 2.1, which is located significantly below the full CI curve, the ECCD potential energy curve describing the dissociation of the S4 system into four H atoms is located slightly above the full CI curve. Thus, in spite of its formally nonvariational character, the ECCD approach based on the ECC theory of Piecuch and Bartlett60 provides a highly accurate and variational description of the breaking of all four H—H bonds in the S4 system. The substantial improvement in the description of the S4 model system offered by the ECCD approach suggests that the T2 clusters resulting from the ECCD calcula- tions are considerably better than the CCD T2 values. This can be seen by calculating the expectation values of the Hamiltonian, designated by (X), where X = CCD or ECCD, with the normalized CCD-like wave functions 5113’”) = NeTéX’|<1>), (X = CCD, ECCD), (2.58) where Téx) (X = CCD, ECCD) are the T2 cluster components obtained with the CCD and ECCD methods, respectively, and N (X) = (@Ie‘TigX))teT2(X)|)‘1/2 are the appropriate normalization factors. Clearly, the (CCD) and (ECCD) values provide the upper bounds to the exact, full CI, energies. However, as demonstrated in Table 2.1, the (CCD) energies remain poor in the 3.5 bohr S a g 4.5 bohr region, whereas 22 the expectation values of the Hamiltonian calculated with the ECCD wave function IWSECCD)» Eq. (2.58), are very close to the corresponding full CI energies. The MBS S4 model is so simple that we can clearly understand the reasons of the excellent performance of the ECCSD = ECCD theory at larger 0 values. The MBS S4 system has only four electrons and the T1 and T3 components vanish due to the high symmetry of the Hamiltonian. Thus, in order to obtain a highly accurate description of the ground electronic state of the MBS S4 system, we must use a method which is capable of providing an accurate description of the effects due to connected quadruply excited (T4) cluster components, whiCh are missing in CCSD. It turns out that at least some of the T4 effects are brought into the ECC formalism as products of the low-order many-body components of 23‘ and T. Indeed, as shown, for example, in Ref. 23, the leading, fifth-order, contribution to the energy due to T4 clusters can be estimated by adding the E5]Q term, defined as 1 E818 = Z<2I>, (2.59) to 1 E814 = 5<9IZ|H|). For the MBS S4 model, the E8.) energy component, Eq. (2.60), vanishes, since Ty] = 0. Thus, the entire fifth-order effect due to T, can be estimated in this case by the E5]Q contribution, Eq. (2.59), which appears in the ECCD energy as the %(|(E;)2(VNT22)C|) term, since T 2 and 232 are similar when 23 MBPT converges. This means that the ECCD energy for the MBS S4 model contains a great deal of information about the leading effects due to T4 and these are sufficient to provide the excellent results at all values of the parameter a observed in Figure 2.1 and Table 2.1. The MBS S4 model allows us to obtain useful insights into the performance of the ECC approximations in the calculations for quasi-degenerate electronic states, but we cannot use it to test all important aspects of the ECC theory. For example, the MBS S4 system is too simple to analyze the importance of the T1 and 23} components and it does not allow us to understand the significance of terms that distinguish the LECCSD, BECCSD, QECCSD, and full ECCSD approximations defined in Section 2.2.1. Moreover, multiple bond breaking in real molecules can be considerably more complicated than the dissociation of the H4 cluster represented by the S4 system into four hydrogen atoms. A good example of the very challenging situation, which is considerably more complex than the situation created by the S4 model, is provided by the triple bond breaking in N2, where the standard CCSD approach displays a colossal failure (see Table 2.2 and Figure 2.2). We tested the ECCSD, LECCSD, BECCSD, and QECCSD methods, based on Eqs. (2.33)—(2.42), using the minimum basis set STD-3G81 model of N2. In all correlated calculations for N2 discussed below, the lowest two core orbitals were kept frozen. 24 Table 2.2.: Ground—state energies of the N2 molecule, as described by the STD-3G basis set.“ Rb Full CI CCSD BECCSDc QECCSDc ECCSDc 1.0 -101.791600 0.319 0.298 0.298 0.298 1.5 — 106.720117 1.102 0.885 0.885 0.885 2.0 -107.623240 3.295 1.897 1.897 1.897 2.5 — 107.651880 9.220 3.442 3.442 3.428 3.0 - 107.546614 13.176 3.919 3.908 3.758 3.5 — 107.473442 —38.645 5.280 5.322 4.746 4.0 -107.447822 — 140.376 15.580 15.968 14.122 4.5 —107.441504 — 184.984 26.795 27.769 24.039 5.0 -107.439549 -200.958 34.134 35.732 30.390 5.5 —107.438665 —206.974 38.368 40.491 33.867 6.0 —107.438265 —209.538 40.730 43.227 35.746 7.0 — 107.438054 —21 1.915 42.754 45.595 37.306 8.0 —107.438029 -—213.431 43.405 46.320 37.799 “The full CI total energies are in hartree. The CC and ECC energies are in millihartree relative to the corresponding full CI energy values. The lowest two occupied orbitals were frozen in the correlated calculations. bThe N—N separation in bohr. The equilibrium value of R is 2.068 bohr. “The BECCSD, QECCSD, and full ECCSD methods are based on the Piecuch-Bartlett variant of the ECC theory. 25 -107035 ' l v u v I v —107.40 - "figi : c 3 c ) -107.45 1' UCCSD ‘ Ia" , E \ 9 . <1~ . 0:, —107.55 (1‘46“ LIJ “Gm-43:] ‘ 11 —107.60 - ' - L‘ —107.65 7 = = = J r -107. 70 2.0 3.0 4.0 50 60 7.0 8.0 RN_N(bohr) Figure 2.2: Potential energy curves for the N2 molecule, as described by the STD-3G basis set. The LECCSD, BECCSD, QECCSD, and ECCSD methods are based on the Piecuch-Bartlett variant of the ECC theory. 26 The results in Table 2.2 and Figure 2.2 clearly demonstrate that the complete ECCSD formalism of Piecuch and Bartlett“, in which all nonlinear terms in (21 + 23;) and (T 1 + T2) are included, and its bilinear and quadratic variants, BECCSD and QECCSD, respectively, defined by the truncated Hamiltonians 1718300313) and H‘QECCSD), Eqs. (2.41) and (2.42), respectively, provide remarkable improvements in the very poor description of the potential energy curve of N2 by the standard CCSD method. The huge negative errors in the CCSD results at larger N—-N separations, R 2 4.5 bohr, of about -—200 millihartree, reduce to much smaller positive errors when the ECC methods are employed (24—38 millihartree in the full ECCSD case, 27—43 millihartree in the BECCSD case, and 28—46 millihartree in the QECCSD case). We also observe a considerable reduction of errors for smaller values of R, including the equilibrium, R z 2.0 bohr, region (see Table 2.2). As shown in Table 2.2 and Figure 2.2, the BECCSD, QECCSD, and full ECCSD approaches eliminate the pathological behavior of the CCSD method at larger N—N distances. As in the MBS S4 case, the BECCSD, QECCSD, and full ECCSD approaches restore the variational description of the potential energy curve of N2 at all internuclear separations. Our results for N2 show that it is not necessary to insist on the bi-variational determination of the 2" and T operators, exploited in the Arponen-Bishop ECC formalism, in order to obtain great improvements in the description of multiple bond breaking by the ECC theory. The fact that the ECC theory of Piecuch and Bartlett is not rigorously bi—variational seems to be of little significance, since our BECCSD, QECCSD, and ECCSD results obtained with this theory are of the same quality as the strictly bi-variational QECCSD results for N2 reported in subsection 2.3.2. The presence of quadratic terms in 2* in QECCSD), H‘BECCSD), and ITQECCSD), Eqs. (2.37), (2.41), and (2.42), respectively, and the use of two independent cluster operators T 27 and E in the ECC formalism, which are optimized by solving a coupled system of equations, are more important for improving the results in the bond breaking region than the particular way of obtaining the t3, tbjb’ of, and 09b amplitudes that defines 3.7 the ECCSD formalism of Piecuch and Bartlett. The importance of the quadratic terms in (21 + 23;), such as %(E{ + 2;)2 H(CCSD) and %H(CCSD) (21 + 2;)2, in the ECCSD equations becomes apparent when we com- pare the BECCSD or QECCSD results with the results of the LECCSD calculations. These quadratic terms are ignored in the LECCSD method (see Eq. (2.40)) and, in consequence, the LECCSD potential energy curve for N2 has the same type of hump for the intermediate values of R as the CCSD curve (see Figure 2.2). It is interesting to observe, though, a substantial reduction of errors in the CCSD re- sults at larger N—N separations, when the LECCSD approach is employed. This corroborates our earlier statements that the use of two independent cluster oper- ators, T and 2", in the ECC theory is more important for improving the poor CCSD results at larger R values than the specific procedure used to determine T and 2". It is also worth noticing that we can safely neglect the higher-order nonlinear terms, such as $031 + 23;)? ECCSD) (231 + 2;), $031 + 2;)?!(0051’) (2’; + 2;)”, and 1 4 methods (cf. Eqs. (2.42) and (2.37), respectively) and absent in the BECCSD ap- (21 + 2;)2 H(CCSD) (2‘; + 2;)2, which are present in the QECCSD and full ECCSD proach (see Eq. (2.41)). The BECCSD results are of the same quality as the QECCSD and full ECCSD results (see Figure 2.2 and Table 2.2). The BECCSD theory repre- sents the lowest-order ECC approach among various ECC methods examined in this work capable of providing the qualitatively correct description of triple bond breaking in N2. 28 Table 2.3.: Ground-state energies of the N2 molecule, as described by the STD-3G basis set.3 Rb Full CI (CCSD)c (BECCSD)c (QECCSD)c (ECCSD)c 1.0 —101.791600 0.298 0.298 0.298 0.298 1.5 — 106.720117 0.890 0.888 0.888 0.888 2.0 — 107.623240 2.004 1.946 1.946 1.946 2.5 —107.651880 4.316 3.775 3.775 3.775 3.0 — 107.546614 5.288 4.160 4.160 4.161 3.5 — 107.473442 16.755 3.378 3.387 3.388 4.0 -107.447822 80.696 6.922 7.206 7.145 4.5 — 107.441504 95.003 10.603 11.506 11.277 5.0 — 107.439549 91.561 12.391 13.877 13.327 5.5 —107.438665 86.652 13.056 14.931 14.224 6.0 — 107.438265 83.037 13.276 15.356 14.553 7.0 —107.438054 79.607 13.393 15.571 14.724 8.0 -107.438029 78.563 13.451 15.578 14.758 aThe full CI total energies are in hartree. The remaining energies are in millihartree relative to the corresponding full CI energy values. The lowest two occupied orbitals were frozen in the correlated calculations. bThe N—N separation in bohr. The equilibrium value of R is 2.068 bohr. c(X) (X = CCSD, BECCSD, QECCSD, ECCSD) is the expectation value of the Hamilto- nian with the eT1+T2|) wave function, where T1 and T2 are obtained with method X (cf. Eq. (2.62)). 29 Finally, before describing the calculations employing the Arponen-Bishop ECC theory, let us discuss the quality of the T1 and T2 clusters resulting from various types of the ECCSD calculations for N2. The remarkable improvements in the description of triple bond breaking in N2, offered by the BECCSD, QECCSD, and full ECCSD methods, imply that the T1 and T2 clusters resulting from the bilinear, quadratic, and full ECCSD calculations are much more accurate than the T1 and T 2 operators obtained with the standard CCSD approach. As in the case of the MBS S4 model, we examined the quality of the T1 and T2 clusters obtained in the CCSD and various ECC calculations for N2 by computing the expectation values of the Hamiltonian, designated by (X), where X = CCSD, BECCSD, QECCSD, and ECCSD, with the normalized CCSD-like wave functions 103’”) = N“) efo’+Téx’|), (X = CCSD, BECCSD, QECCSD, ECCSD), (2.62) where 7",“) and 7.)") (X = CCSD, BECCSD, QECCSD, ECCSD) are the T1 and T2 cluster components obtained with the CCSD, BECCSD, QECCSD, and complete ECCSD methods, respectively, and N (X l = (<1>|e(T{X))t+(TéX))teTlxb'Tén|)"1/2 are the corresponding normalization factors (see Table 2.3). As demonstrated in Table 2.3, the expectation values of the Hamiltonian obtained with the CCSD wave function are extremely poor at larger N—N separations, whereas the expectation values of the Hamiltonian calculated with the BECCSD, QECCSD, and full ECCSD wave SBECCSD)), I‘I’SQECCSD)), and I‘IIEECCSD)), respectively, are very close to functions, |\II the corresponding full CI energies at all values of R. The fact that we observe a fairly substantial (2—3-fold) reduction of unsigned errors, when the standard CCSD energy EéCCSD) = (®|H(CCSD)| D . QWUUm—OHN D 0 Wm LomoomehflN 5802—3? 4‘ . #0 \m r QWUUMQNX 0 g 00.0 W a ...A omooux n. N . md IZx 3.0 bohr, the differences between the CCSD and BECCSD/QECCSD/ECCSD oper- ators T increase so much (cf. Figure 2.3 (b)) that the behavior of the CCSD method and the behavior of the BECCSD/QECCSD/ECCSD approaches for stretched nuclear geometries are totally different. As we have already discussed, the standard CCSD method completely fails at larger R values, whereas the BECCSD/QECCSD/ECCSD 33 x1 approaches provide very good results. Figure 2.3 (b) illustrates another important feature of the ECCSD theory, namely, the similarity of the T and 23 operators in the equilibrium region and the significant difference between the T and )3 operators ob— tained in the ECCSD calculations in the region of larger N -N distances. As mentioned earlier, the lowest-order MBPT estimates of the operators 2 and T are identical (cf. Refs. 60—62, 83, 84). In consequence, in the equilibrium region, where the MBPT series rapidly converges, we. have 2 z T (see Figure 2.3 (b)). The situation changes, when the convergence of the MBPT series is slow or when the MBPT series diverges, as is the case when the N —N bond is stretched or broken. For larger N—N separations, the operators 2 and T become completely different. Figure 2.3 (b) provides us with a direct evidence that this is indeed what happens at larger R values. 2.3.2 The Arponen-Bishop ECCSD Approach So far, we have tested the ECCSD methods based on the ECC formalism of Piecuch and Bartlett. We have demonstrated considerable improvements offered by the ECCSD approximations when multiple bonds are broken. The question is if similar improve- ments can be obtained when the alternative formulation of the ECC theory, proposed by Arponen and Bishop, is exploited. This question is addressed in this subsection. The usefulness of the bi—variational ECCSD theory of Arponen and Bishop based on Eqs. (2.39) and (2.43)—(2.46) in improving the results for multiple bond breaking becomes apparent when we examine the results for the STD-3G model of N2 shown in Table 2.4 and Figure 2.4. As one can see, the ECCSD method of Arponen and Bishop employing the ground-state RHF determinant as a reference provides remarkable im- provements in the very poor description of the potential energy curve of N2 by the standard CCSD method. The results are as good as those obtained with the ECCSD 34 approach of Piecuch and Bartlett. Indeed, the huge negative errors in the CCSD re- sults at larger N—N separations, which exceed —200 millihartree in the R > 4.5 bohr region, reduce to much smaller positive errors (on the order of 31—38 millihartree when R > 4.5 bohr) in the ECCSD/Arponen-Bishop case. As shown in Figure 2.4, the ECCSD approach of Arponen and Bishop completely eliminates the pathologi- cal behavior of the standard CCSD method at larger N—N distances, restoring the variational description of the potential energy curve of N2 at all internuclear separa- tions. As in the case of the Piecuch-Bartlett theory, the ECCSD approach of Arponen and Bishop is capable of capturing the most essential nondynamic correlation effects (which the small STD-3G basis set used in these calculations already describes) in spite of the use of the single, spin— and symmetry-adapted, RHF determinant as a reference. The comparison of the CCSD and ECCSD potential energy curves for N2 shown in Table 2.4 and Figure 2.4 confirms once again that the T1 and T2 cluster components obtained in the ECCSD calculations are of much higher quality in the bond breaking region than the T1 and T2 clusters resulting from the standard CCSD calculations. As shown in Table 2.4 and Figure 2.4, the QECCSD/Arponen-Bishop results for the STD-3G model of the N2 molecule, obtained by truncating the ECCSD energy functional at terms quadratic in (2} + 2;) (see Eqs. (2.52)-(2.57)), are much better than the corresponding CCSD results and almost as good as the results of the full ECCSD calculations. The huge negative errors in the CCSD results in the R > 4.5 bohr region, which exceed -200 millihartree, reduce to much smaller positive errors on the order of 35—46 millihartree, when the QECCSD method is employed. As in the case of the full ECCSD approach, the QECCSD approximation employing the Arponen-Bishop ECC formalism eliminates the pathological behavior of the CCSD 35 method at larger N —N distances, restoring the variational description of the potential energy curve of N2 at all internuclear separations. The variational and qualitatively correct behavior of the QECCSD method based on the Arponen-Bishop ECC theory is independent of the basis set. Indeed, as shown in Table 2.5 and Figure 2.5, the QECCSD potential energy curve for the double zeta (DZ)85 model of N2 is much better than the corresponding CCSD curve. The QECCSD potential energy curve for the N2 molecule obtained with the DZ basis set, shown in Figure 2.5, is located above the full CI curve. The large negative errors in the CCSD results in the R 2 2Re region of (—70) — (—121) millihartree (RC = 2.068 bohr is the equilibrium value of R) are replaced by the considerably smaller positive errors of 40—50 millihartree when the QECCSD method is employed. As in the case of the STD-3G basis set, the QECCSD approach based on the Arponen-Bishop ECC theory eliminates the well pronounced hump on the CCSD potential energy curve, when the DZ basis set is employed. Thus, the QECCSD approach exploiting the Arponen-BishOp ECC theory pro- vides a practical method of capturing the large nondynamic correlation effects in N2, in spite of the single-reference nature of the ECC formalism, in spite of the use of the RHF determinant in the QECCSD calculations, and, what is probably most re- markable, in spite of the two-body character of the QECCSD (or ECCSD) theory, which uses one- and two-body cluster T1, T2, 21, and 22 only. As a matter of fact, the QECCSD results at larger internuclear separations of N2 are much better than those obtained with the standard CC methods with singly, doubly, triply, and even quadruply excited clusters (cf. the QECCSD results in Table 2.5 and Figure 2.5 with the corresponding CCSD(T), CCSDT, CCSD(TQf), and CCSDT(Qf) results). 36 Table 2.4.: The ground-state energies of the N2 molecule obtained for several inter- nuclear separations R with the STO-3G basis set.“ Rb Full CI CCSD QECCSDc ECCSDc 1.5 -106.720117 1.102 0.886 0.885 2.0 —107.623240 3.295 1.898 1.897 2.5 —107.651880 9.220 3.443 3.427 3.0 — 107.546614 13.176 3.909 3.757 3.5 - 107.473442 —38.645 5.294 4.746 4.0 —107.447822 — 140.376 15.815 14.148 4.5 —107.441504 - 184.984 27.792 24.198 5.0 —107.439549 —200.857 35.335 30.590 5.5 —107.438665 —206.974 39.983 34.158 6.0 —107.438265 -209.538 42.609 36.082 7.0 — 107.438054 -211.915 44.839 37.671 8.0 —107.438029 —213.431 45.508 38.161 a"The full CI energies are in hartree. The CCSD, QECCSD, and ECCSD energies are in millihartree relative to the corresponding full CI energy values. The lowest two occupied orbitals were kept frozen. bThe N—N separation in bohr. The equilibrium value of R is 2.068 bohr. cThe QECCSD and full ECCSD methods are based on the Arponen-Bishop ECC theory. 37 -107.35 - . —107.40 - -107.45 - f 3 . . 9 “C -107.50 P 2 {2 .FullCl v \ 00050 g; \ Aoecoso 5 —107.55 \ veccso - c | Lu \ 0 -107.60 \ 4 \\ \ —107.65 Q‘Qom—o--- —107.7o**1‘4-'-‘*-‘ 2.0 3.0 4.0 5.0 6.0 7.0 8.0 RN_N(bohr) Figure 2.4: Ground-state potential energy curves of the N2 molecule, as described by the STD-3G basis sets. The QECCSD and ECCSD methods are based on the Arponen-Bishop ECC theory. 38 Table 2.5.: The ground-state energies of the N2 molecule obtained for several inter- nuclear separations R with the DZ basis set.“ Method 0.7512. 12.b 1.2512. 1.512. 1.7512. 212. 2.2512. CCSD 3.132 8.289 19.061 33.545 17.714 —69.917 —120.836 CCSDT“ 0.580 2.107 6.064 10.158 —22.468 —109.767 -155.656 CCSD(T)“ 0.742 2.156 4.971 4.880 —51.869 —246.405 —387.448 CCSD(TQf)“ 0.226 0.323 0.221 —2.279 —14.243 92.981 334.985 CCSDT(Q;)“ 0.047 -0.010 —0.715 —4.584 3.612 177.641 426.175 QECCSD“ 2.506 6.236 13.609 23.485 31.060 40.085 49.741 “All energies are in millihartree relative to the corresponding full CI energy values, which are —108.549027, —109.105115, -109.054626, —108.950728, —108.889906, —108.868239, and —108.862125 hartree at R = 0.75R., Rm 1.25R., 1.5R., 1.75R., 2R., and 2.25R., respec- tively. The lowest two occupied and the highest two unoccupied orbitals were frozen in correlated calculations. bThe equilibrium value of R, R. = 2.068 bohr. cFrom Ref. 31. dFrom Ref. 34. eThe quadratic approximation to the full ECCSD theory of Arponen and BishOp. 39 -108.5 . . Full Cl -108.6 - -—- ccso ,’ - aleccsor I ' Dccsom I _ . OCCSDG'Q.) I . 1 08'7 AQECCSD , -108.8 2 —108.9 - Energy (hartree) —109.0 - —109.1 ~ -109.2 . _109.3 --.1....1....1...L11...1L..L14 1.5 2.0 2.5 3.0 3.5 4.0 4.5 RM, (bohr) Figure 2.5: Ground-state potential energy curves of the N2 molecule, as described by the DZ basis sets. QECCSD represents the quadratic version of the ECCSD theory of Arponen and Bishop. 40 2.4 Conclusion We can summarize this chapter by stating that the ECCSD approach of Piecuch and Bartlett, and its approximate BECCSD and QECCSD variants—”“76, and the ECCSD method of Arponen and Bishop and its QECCSD {variant77 represent interesting new alternatives for accurate electronic structure calculations of quasi-degenerate elec- tronic states and bond breaking. The BECCSD, QECCSD, and full ECCSD methods remove the pervasive failing of the standard CCSD approach at larger internuclear separations and provide very good values of the T1 and T2 cluster amplitudes in the bond breaking region, in spite of using the RHF configuration as a reference and in spite of the two-body character of all ECCSD approximations. The BECCSD, QECCSD, and full ECCSD approaches improve the quality of the T1 and T2 cluster components so much that we can start thinking about using the BECCSD, QECCSD, or full ECCSD theories to design new single-reference ab initio methods for quantita- tive, high-accuracy calculations for bond breaking. For example, by having an access to very good T1 and T2 cluster amplitudes, resulting from the BECCSD, QECCSD, or ECCSD calculations, which are much better than the T1 and T2 clusters resulting from the standard CCSD calculations, we should be able to propose simple nonitera- tive corrections to the BECCSD, QECCSD, or ECCSD energies, which may provide further improvements in the ECC results in the bond breaking region (see Chapter 3). Based on a comparison of the ECCSD and QECCSD results for N2 obtained with the Piecuch-Bartlett and Arponen-Bishop theories, we do not expect the differences between the ECC methods of Arponen and Bishop‘w—59 and Piecuch and Bartlett60 to be large in the context of bond breaking. It seems to us that the rigorously bi- variational character of the ECC formalism of Arponen and Bishop is of the secondary 41 importance in the calculations of PESs involving bond breaking (cf. Section 2.3). Our experiences to date indicate that the most important factor that contributes to significant improvements in the quality of the T1 and T 2 cluster components in the bond breaking region is the flexibility of the ECC theories, which rely on two independent sets of cluster amplitudes that are optimized by solving coupled systems of equations. The standard CC theory uses only one set of cluster amplitudes and this is not sufficient to obtain a correct description of multiple bond breaking by the standard CCSD method. On the other hand, costs of the ECCSD calculations employing the ECC theory of Arponen and Bishop are somewhat smaller than those charactering the ECCSD approach of Piecuch and Bartlett. This, in particular, applies to the QECCSD approx- imation, which is simpler when the Arponen-Bishop ECC theory is employed. The QECCSD method based on Arponen’s and Bishop’s formulation of the ECC theory is an .N’6 procedure. The relatively low cost of the QECCSD approximation, the single reference (“black-box”) character of the QECCSD calculations, and the qualitatively correct description of multiple bond breaking by the QECCSD approach make the QECCSD method an attractive theory for the design of noniterative corrections to CC energies. These corrections are discussed next. 42 Chapter 3 Noniterative Corrections to Extended Coupled-Cluster Energies for High Accuracy Electronic Structure Calculations: Generalized Method of Moments of Coupled-Cluster Equations In Chapter 2, we showed that we can provide substantial improvements in the quality of the calculated potential energy curves and electronic quasi-degeneracies if we switch from the standard CCSD theory to its extended ECCSD counterpart. The question arises if we can improve the ECCSD (or QECCSD) results even further and obtain the quantitative description of bond breaking by adding the a posteriori corrections to ECCSD energies that would be reminiscent of the popular triples and quadruples corrections of the CCSD(T) and CCSD(TQf) methods and that would eliminate fail- ures of these methods at larger internuclear separations. In this chapter, we show that such corrections to ECCSD energies can be developed if we use the generalized version of the method of moments of CC equations (the MMCC theory) of Piecuch and Kowalskill-13’3k32’74’75. Before describing the generalized MMCC approach, we discuss the standard MMCC formalism. 43 3.1 The Standard Method of Moments of Coupled-Cluster Equations The main idea of the standard MMCC theory‘1‘13’30—9’257‘L75 is that of the noniterative energy correction 53") E E0 — E3“, (3.1) which, when added to the energy ESA) obtained in some standard approximate CC calculation A, such as CCSD, recovers the exact, full CI, energy E0. The objective of the approximate MMCC methods is to approximate corrections 6),"), such that the resulting MMCC energies, defined as EgMMCC) = 133") + 63"), (3.2) are close to the corresponding full CI energies E0. The ground-state MMCC for- 12’13’3‘H38 and genuine multi-reference CC malism can be extended to excited states theories32’89'90. In this thesis, we focus on the ground-state problem and merging the single-reference MMCC formalism with the non-standard CC theories, such as the ECC method of Arponen and Bishop50—59 (see Refs. 75 and 77). In the standard formulation of the ground-state MMCC theory, we use the non- iterative corrections 66’4) to improve the results of the standard CC calculations. By the standard CC calculation, we mean any single-reference CC calculation in which the many-body expansion for the cluster operator T, defining the CC ground state |\Ilo), Eq. (2.4), is truncated at some excitation level mA < N (recall that N is the number of electrons in a system). The general form of the truncated cluster operator T”), defining the standard CC approximation A characterized by the excitation level mA, is given by Eq. (2.20). An example of the standard CC approximation is the 44 CCSD method. In this case, m A = 2 and the cluster operator T is approximated by T z T(CCSD) = Tl + T2, (3.3) where T1 and T2 are defined by Eqs. (2.25) and (2.26), respectively. Other examples of the standard CC approximations are the full CCSDT, CCSDTQ, and CCSDTQP approaches mentioned in the Introduction, in which mA 2 3, 4, and 5, respectively. The standard CC system of equations for the cluster amplitudes tall 1112.3" defining the Tn components of T (A) has the following general form: (@2111 °"ia"|H(A)|) = 0, i1 < < in, a1 < < an, (3.4) o o o n where n = 1,...,mA, H“) = e"T(A)HeTW = (HeTW)C (3.5) is the similarity-transformed Hamiltonian of the CC theory, and Id);l ' ‘ '19") = E2911 ° ’ 'ia"|) are the n-tuply excited determinants. In particular, the n n standard CCSD equations for the singly and doubly excited cluster amplitudes t3 and 4sz defining operators T1 and T2, respectively, are (‘I’flH‘CCSD’l‘M = 0, (3-6) (¢g]b|H|<1>) = 0, 2‘ < j, a < b, (3.7) where H‘CCSD) is the similarity-transformed Hamiltonian of the CCSD approach de- fined by Eq. (2.37). Once the system of equations, Eq. (3.4), is solved for T”) (or Eqs. (3.6) and (3.7) are solved for T1 and T 2), the CC energy corresponding to the standard approximation A is calculated as E5” = (<1>|H(A)|). (3.8) 45 The fundamental formula of the ground-state MMCC formalism introduced in Refs. 11 and 30 (cf., also, Refs. 12, 30—32, 74, 75), which expresses the energy differ- ence 6“), Eq. (3.1), in terms of the generalized moments of the single-reference CC equations of method A, has the following form: 6W=E— E”): ZN: Z (-eonC..(mA)M.(mA)I>/<\II0IT""I«I>>. (3.9) n=m4+lk=m4+l Iiere, Cn—k(mA) = (6TH) )n—k (3-10) is the (n — k)-body component of the wave operator 8TH) defining the CC method A, |\Ilo) is the full CI ground state, and M.(mA)I> = Z Max:113: (m mg“ ,, an. (3.11) i] < ' ° ' < ik 0.1 < ' ' ' < ak The coefficients Mal iffk(m,() entering Eq. (3.11) represent the general moments of the CC equations of method A, Maizzzé‘Jm/a) = ((pgllxjiikIHMnQ). (3.12) By comparing Eqs. (3.12) and (3. 4), we can state that moments .Mallz .i" k(mA) entering 63") represent the projections of the CC equations of method A on excited determinants Willi?) with k > m ,4. Equation (3.9) states that by calculating quantities Cn_k(m,4), Eq. (3.10), and moments M311'_'.'.ic’fk (mA), Eq. (3.12), with k > mA, we can determine the nonit- erative energy correction 66A) to the CC energy BSA) that we recovers the full CI energy E0. The determination of moments Mill'ffjt’z‘JmA) for the low-order CC methods, such as CCSD, is relatively straightforward (cf. Eqs. (3.15)-(3.18) below). The Cn_k(mA) terms entering Eq. (3.9) are easy to calculate too. The zero-body 46 term, Co(m,4), equals 1; the one-body term, Cl(m,4), equals T1; the two-body term, 1 02(mA), equals T2 +%T12 if mA 2 2; the three-body term, C3(mA), equals TIT; + 6T13 if 172); = 2 and T3 + T 1T2 + -1-T ,3 if mA 2 3, etc. Thus, the above formula for 63"), 6 Eq. (3.9), is an excellent starting point for developing noniterative CC approaches, in which the ground-state energies are calculated by adding corrections 68A) to the stan- dard CC energy EéA). For example, we can develop a hierarchy of approximations, in which the energy corrections based on Eq. (3.9) are added to the CCSD energies. In this case, the formula for the correction 66cc”), which must be added to the CCSD energy ESCCSD) to recover the full CI energy E0, is N min(n,6) 68‘3“”) 2 E0 — E3003”) = Z Z (wolc.-.(2)M.(2)I>/<%IeT1+T2|>, (3.13) n=3 k=3 where Mk(2)I> = Z M23112... (2) I331 2'. a). (3.14) i1 < ' ' ' < ik 0.1 < - ° - < a), with moments M211 11125,: (2), k = 3 — 6, defined as projections of the CCSD equations on triply, quadruply, pentuply, and hextuply excited determinants, 1 1 1 1 MWQ): (cpgjk|[HN(T2+TlT2+ —T22+ +—T2T2+—T1T§+—T?T2)]c|<1>), (3-15) Mabc 2 2 1 2 6 1 1 M23“(2)=(<1>ab,g;1|(H~(-;- Tg+ 41T2+——T:+1T3T2)]C|q>), (3.16) 3] 2 6 4 kl Mijbcd’g‘e): g<<1> 3,”,§§‘,$,IlH~(T§ +T1T3>JCI> (3.17) ijklmn __ abcdef Mam, (2) — 24<¢,,k,mnI(H~13)cl>. (3.18) The exact MMCC corrections 63"), Eq. (3.9), or 630030), Eq. (3.13), have the form of the complete many-body expansions involving all n-tuply excited configura- tions with n = mA + 1, . . . , N, where N is the number of electrons in a system (cf. 47 the summations over 17. in Eqs. (3.9) and (3.13)). Thus, in order to develop prac- tical methods based on the MMCC theory, we must first truncate the many-body or 6(ccsn) expansions for corrections (SW at some, preferably low, excitation level m3 satisfying mA < 7113 < N. This leads to the MMCC(mA, m3) schemes, in which we calculate the energy as followsl1‘13’3°_32'74’75: EéMMCC)(mAa m3) : E(()A) + 60(mAa m3), (3'19) where E3” is the energy obtained with the CC method A and 60>/ > (3.20) n=mA+l k: mA-l-l is the relevant MMCC correction. Examples of the MMCC(mA, m3) schemes are the basic MMCC(2,3) and MMCC(2,4) approximations, in which energies are calculated as followsl“1330—3523435. E‘MMCC’e 3): E‘CCSD’ + (‘I’olM3(2)|‘I’>/(‘I’o IeTI+T2I> (3.21) E), (3.23) where T1 and T2 are the singly and doubly excited clusters obtained in the CCSD calculations, T91 is an approximation of the connected triply excited clusters T3, defined by Eq. (2.61), and Zsl<1>) = Rés’VNTIId» (3.24) is the disconnected triples correction, which is responsible for the difference between the [T] and (T) triples corrections of the standard CCSD[T] and CCSD(T) ap- proaches. The higher-order CR-CCSD(T Q) methodsl1‘13’30’31’33735174'75 are obtained in a similar manner, by inserting the MBPT(2)[SDTQ]-like expressions for |\Ilo) into the MMCC(2,4) formula, Eq. (3.22). For example, the wave function |\Ilo) defin- ing variant “b” of the CR-CCSD(TQ) approach (the CR-CCSD(TQ),b method) is defined as follows: IwECSD‘TQ”) = Wm) + 223231». (3.25) where |W€CSD(T)) is given by Eq. (3.23). As shown in Refs. 11—13, 30—35, 37, 74, 75, and 94—96, the CR—CCSD(T) and CR-CCSD(TQ) approaches eliminate or considerably reduce the failures of the stan- dard CCSD(T) and CCSD(TQf) methods at larger internuclear separations and for 49 diradicals without making the calculations substantially more difficult or expensive. In particular, the CR—CCSD(T) and CR-CCSD(TQ) methods provide a very good description of single bond breaking. Unfortunately, performance of these methods for multiply bonded systems is often much less impressive, partly due to the very poor quality of the singly and doubly excited cluster amplitudes resulting from the CCSD calculations, on which the CR—CCSD(T) and CR-CCSD(TQ) methods are based, in calculations involving multiple bond breaking. The purpose of this work is to ex- amine an alternative approach, in which the MMCC corrections 63A) or 65,0030) are constructed using the cluster components obtained in the ECC or ECCSD calcula- tions. As shown in Chapter 2, the ECCSD method provides much better values of T1 and T2 clusters than the CCSD approach when multiple bonds are stretched or broken, so that we may be able to improve the CR-CCSD(T) and CR-CCSD(TQ) results in cases where the standard CCSD values of T1 and T2 are of very poor quality. The use of the ECC or ECCSD values of cluster amplitudes in MMCC calculations requires the generalization of the MMCC theory to non-standard CC methods, which is described in the next section. 3.2 The Generalized MMCC Formalism Interestingly, Eqs. (3.9) and (3.13) can be generalized to a situation where the trun- cated cluster operator T“) is not determined by solving the standard CC equa- tions, Eq. (3.4). Here is how this works: When the cluster operator T“) is ob- tained in a non-standard way, we can no longer assume that the generalized moments Mall'jfffik (mA) with k = 1,. . . , mA vanish; they only vanish in the standard CC case (cf. Eqs. (3.4) and (3.12)). It has been shown in Refs. 13 and 30 that when mo- 50 ments M211°.°.'.z(’1‘k(m,4) with k = 1, . . . m A do not vanish, we have to use the following expression for the exact, full Cl energy E0 instead of Eq. (3.9): N E. = ZZ>/<\IIoIeT“’I> n=0 k=0 N n = Moon.)+ZZ>/<%IeT“’I>. (3.26) n=l k=1 Here, Mo(mA) designates the zero-body moment, which is calculated in exactly the same way as the CC energy Eé"), i.e. (cf. Eq. (3.8)) Mum.) = <IFI""I>. (3.27) Although the formulas for Mo(m,4) and E84) are identical, there is a fundamental difference between Mo(mA) and Eff). The energy E5,“ is determined using the cluster amplitudes originating from the standard CC equations, Eq. (3.4), and E84) is the energy expression used in the CC theory. The zero—body moment Mo(mA) can be computed with any cluster operator T“), obtained, for example, by performing some nonstandard calculations, such as the ECC calculations, and Mo(mA) does not have to represent the energy expression of the nonstandard CC theory used to generate T(A) (e.g., Mo(m,4) does not represent the energy expression of the ECC formalism). As in the case of the standard MMCC formalism defined by Eq. (3.9), the Cn-k(m,4) and Mk(m,4)|) quantities entering Eq. (3.26) are defined by Eqs. (3.10) and (3.11), respectively, although we must consider now all quantities Mk(mA)|) with k 2 1, not just those with k > mA. In other words, since we no longer assume that the generalized moments M311 'o'.'.igk(m,4) with k = 1,. . . , mA vanish, since Eq. (3.4) is no longer satisfied, we must consider all generalized moments Mill '. ’. 11.3,: (m A), which a given cluster operator T“) produces. In particular, if we want to use the T1 and T2 clusters to construct the full CI energy E0, which are no longer determined 51 by solving the standard CCSD equations, we must use the following formula for the exact energy E0: N min(n,6) E0 = 140(2) +2 2 (10(0,_,.(2) Mk(2)|)/(\Ilo|eT1+T2|), (3.23) =1 k=l where Mk(2)|) is defined by Eq. (3.14). As one can see, Eq. (3.28) is very similar to Eq. (3.13). In particular, we do not have to consider momentsMé‘lzng) with k > 6, since for Hamiltonians containing up to two-body interactions the generalized moments Mall‘jji’fka) with k > 6 vanish, independent of the source of T1 and T2 clusters. There is, however, a difference between Eqs. (3.13) and (3.28): in Eq. (3.28) we consider the singly and doubly excited moments, 1143(2) = ((bgll—Iwcsmkb) and MZJAQ) = (¢?Jb|H(CCSD)|), respectively, which are no longer zeroed, along with moments M31, 11115,,(2) with k = 3 — 6 considered in the standard CCSD case; Eq. (3.13) uses moments M31115), (2) with k = 3 — 6 only, since Mg(2) = Mijb(2) = 0 in the standard CCSD case. The formulas for the M32) and M329) moments, entering Eq. (3.28), in terms of the T1 and T2 clusters, are identical to the left-hand sides of the standard CCSD equations. Thus, we obtain (cf. Eqs. (3.6) and (3.7)), - 1 1 M242) = (¢?|[H~(1 + T, + T2 + 5T1? + T171, + 6T3)]C|), (3.29) M11 (2) = («31.11.1|[H~(1+ T1+ T2 +1T‘1 + T1T2 +1T11 +12"1 + 171173 + —1-T4)]C|). ab 2] 2 1 6 1 2 1 2 1 24 1 (3.30) Equations (3.26) and (3.28) define the generalized version of the ground-state MMCC theory, designated as GMMCC. Clearly, the GMMCC theory reduces to the standard MMCC formalism if cluster components Tfl defining operator T“) are determined by solving the standard CC equations, Eq. (3.4). In this case, the M311 '.'.'.ic’{k(m,4) moments with k = 1, . . . , m A vanish and the summations 25:1 22:, entering Eq. (3.26) reduce to ELM +1 22:,“ +1, giving Eq. (3.9). The obvious ad- 52 vantage of Eqs. (3.26) and (3.28) is that they are much more general than Eqs. (3.9) and (3.13), enabling us to use the non-standard values of T“) clusters. Otherwise, we use Eqs. (3.26) and (3.28) in exactly the same way as Eqs. (3.9) and (3.13). Thus, once the cluster components defining the truncated cluster operator T”) are determined, we calculate the relevant moments Mo(m,4) and M311'_'.'.i§k (mA) and use these moments to determine the ground-state energy E0. In analogy to the standard MMCC theory, a few issues have to be addressed be- fore using the GMMCC formalism in practical calculations. First of all, the exact GMMCC expressions, Eqs. (3.26) or (3.28), represent the complete many-body ex- pansions involving all n-tuply excited configurations with n = 1,. ..,N, where N is the number of electrons in a system (see the summations over n in Eqs. (3.26) and (3.28)). Thus, in order to develop the computationally tractable GMMCC methods, we must first truncate the many-body energy expansions, Eqs. (3.26) or (3.28), at some, preferably low, excitation level m3, where mA < 1713 < N. This leads to the GMMCC(mA, m3) schemes, which are the non-standard analogs of the MMCC(mA,mB) approximations discussed in Section 3.1. By limiting ourselves to the wave functions |\Ilo) that do not contain higher—than—mB-tuply excited compo- nents relative to reference |) and by restricting the summation over n in Eq. (3.26) accordingly, we obtain the following energy expression for the GMMCC(mA,m3) methods: mg n EéGMMCC1(mA, m3) = Mum.) + 2 2(2010.-.(m..) Mk(mA)|‘I’)/(‘I’0|€Tw (<2). n=1 k=l (3.31) where Mo(mA) is defined by Eq. (3.27). The exact GMMCC formalism, equivalent to calculating the full CI energies, is obtained when |\Ilo) in Eq. (3.31) is the full CI ground-state wave function and m3 = N. In this case, it is irrelevant what is the 53 value of mA and where is the cluster operator T“) taken from. When I‘Ilo) is exact and m3 = N, Eq. (3.31) produces the exact energy, independent of the excitation level m A and the source of cluster amplitudes defining T”). In this thesis, we focus on the GMMCC(mA,m3) schemes with mA = 2, which can be used to correct the results of the non-standard CCSD-like (e.g., ECCSD or QECCSD) calculations. In analogy to the standard MMCC theory, two GMMCC(mA, m3) approximations are expected to be particularly useful: the GMMCC(2,3) method and the GMMCC(2,4) approach. According to Eq. (3.31), the GMMCC(2,3) and GMMCC(2,4) energies are calculated as follows: EéGMMCC’a. 3) = M0(2) + (on{M1(2) + (142(2) + T1M1(2)] + [M3(2) +T.M2(2) + (T: + §T3)M1(2)1}I>/<%Ie11+11I> . (3.32) ESGMMCC’Q. 4) = Mo(2) + <%I{M1(2) + (142(2) + T1M1(2)1 +[M3(2) + T1M2(2) + (T2 + %T3)M1(2)] +[M.,(2) + T1M3(2) + (T2 + éTE)M2(2) 1 +(T1T2 + 6 T13)M1 (2)1}|)/(1I’oleT1+T2 I‘P) . (333) As one can see, in the case of the GMMCC(2,3) approximation, we must calculate moments M3(2), MZ)(2), and M3£(2), using Eqs. (3.29), (3.30), and (3.15), re- spectively, in addition to moment Mo(2), Eq. (3.27). In the case of the GMMCC(2,4) approach, we must determine moments M3132) and M2520), Eqs. (3.15) and (3.16), respectively, along with moments Mo(2), M2(2) and M320), Eqs. (3.27), (3.29) and (3.30), respectively. Clearly, when the singly and doubly excited moments M39) and MZJAQ) vanish, the GMMCC(2,3) and GMMCC(2,4) energy expressions, Eqs. (3.32) and (3.33), respectively, reduce to the MMCC(2,3) and MMCC(2,4) for- 54 mulas given by Eqs. (3.21) and (3.22). This can only happen when the T1 and T 2 clusters are obtained by solving the standard CCSD equations, Eqs. (3.6) and (3.7). The second issue that needs to be addressed before the GMMCC(2,3) and GMMCC(2,4) methods and other GMMCC(mA,m3) approximations are used in practice is the issue of the wave function I‘llo) that enters Eqs. (3.31)-(3.33), which in the exact theory is a full CI ground state. Clearly, in order to make the GMMCC(2,3), GMMCC(2,4), and other GMMCC (m ,4, m3) schemes usable in practical applications, we must suggest some approximate forms of |\Ilo) that can be easily generated with one of the inexpensive ab initio approaches. Finally, the third issue is the source of the T1 and T2 cluster amplitudes which are needed to construct the GMMCC(2,3) and GMMCC(2,4) energy expressions, Eqs. (3.32) and (3.33), respectively. As mentioned earlier, in calculating the GMMCC(2,3) and GMMCC(2,4) energies we would like to use the T1 and T2 cluster components which are more accurate in cases involving multiple bond breaking than those obtained in the standard CCSD calculations. As shown in Chapter 2, the T1 and T2 cluster components resulting from various types of ECCSD calculations are much better than their standard CCSD counterparts, when multiply bonded systems are examined. Thus, it is worth examining the pos- sibility of combining the GMMCC(2,3) and GMMCC(2,4) schemes with the ECCSD methods. The resulting ECCSD(T), ECCSD(TQ), QECCSD(T) and QECCSD(TQ) approaches are discussed next. 55 3.3 The ECCSD(T), ECCSD(TQ), QECCSD(T), and QECCSD(TQ) Methods and their Performance in Cal- culations for Triple Bond Breaking in N2 The ECCSD results for triple bond breaking in N2 are so much better than their standard CCSD analogs that it is very important to analyze the effect of replacing the CCSD values of the T1 and T2 cluster components in the MMCC calculations by the ECCSD values of these clusters. Since the ECCSD and QECCSD methods are no longer the standard CC theories, so that we can no longer assume that the singly and doubly excited moments, MZ(2) and M2,.)(2), Eqs. (3.29) and (3.30), respectively, vanish, we must use the GMMCC formalism discussed in Section 3.2 rather than the standard MMCC formalism discussed in Section 3.1 in such considerations. In the following, we test an idea of using the QECCSD and ECCSD values of T1 and T2, resulting from the application of the Arponen-Bishop variant of the ECC theory, in the GMMCC calculations. As explained in Section 3.2, in practice we are interested in a truncated form of the GMMCC theory that leads to relatively low costs of calculating the final energy. All of our tests to date, including the calculations for N2 discussed below, indicate that the lowest-order GMMCC scheme, employing the ECCSD or QECCSD values of T1 and T2, which provides substantial improvements in the results for multiple bond breaking, is the GMMCC(2,4) approach defined by Eq. (3.33). In this approach, we only consider the generalized moments Millijzigkfl) with k = 1 — 4, i.e. mo- ments corresponding to the projections of H(CCSD)|<1>) on singly, doubly, triply, and quadruply excited determinants. The lower-order GMMCC(2,3) approach, defined by Eq. (3.32), employing the ECCSD or QECCSD values of T1 and T2 and ignoring the quadruply excited moments szbgla), provides improvements too, but the ne- 56 glect of the quadruply excited moments “3331(2) in the GMMCC calculations has a negative impact on the results for multiply bonded systems, such as N2. As in all approximate MMCC calculations, we must decide what to do with the wave function |\Ilo) that enters the GMMCC(2,3) and GMMCC(2,4) energy formulas, Eqs. (3.32) and (3.33). Since we are mainly interested in the “black-box” GMMCC approaches of the CCSD(T) or CCSD(TQ) type, in this study of the performance of the GMMCC theory employing the ECCSD and QECCSD values of T1 and T2, we use the same types of the wave functions I‘llo) in Eqs. (3.32) and (3.33) as the wave func- tions (1130511111) and (113051111‘1111’) used in the CR—CCSD(T) and CR-CCSD(TQ),b calculations (cf. Eqs. (3.23) and (3.25)). Depending on the choice of lilo) and the pre- cise source of T1 and T2 clusters for the GMMCC(2,3) and GMMCC(2,4) calculations, we introduce the following four approximations: o ECCSD(T): The ECCSD(T) approach is defined as the GMMCC(2,3) method, in which I‘IIO) is defined by Eq. (3.23) and in which T1 and T2 clusters originate from the full ECCSD calculations of the Arponen-Bishop type. 0 QECCSD(T): The QECCSD(T) approach is defined as the GMMCC(2,3) method, in which [\110) is defined by Eq. (3.23) and in which T1 and T2 clusters originate from the QECCSD calculations of the Arponen-Bishop type. 0 ECCSD(TQ): The ECCSD(TQ) approach is defined as the GMMCC(2,4) method, in which |\Ilo) is defined by Eq. (3.25) and in which T1 and T2 clusters originate from the full ECCSD calculations of the Arponen-Bishop type. 0 QECCSD(TQ): The QECCSD(TQ) approach is defined as the GMMCC(2,4) method, in which I‘IIO) is defined by Eq. (3.25) and in which T1 and T2 clusters originate from the QECCSD calculations of the Arponen-Bishop type. 57 Based on the above definitions, it is easy to verify that the ECCSD(T)/QECCSD(T) energies can be calculated as follows: E31Q1ECCSD‘T11 = 1140(2) + N“) mm, (3.34) where N111 = <IT.1 M1(2)I> + <4>|TJ (142(2) + T1M1(2)]|> +<|(T3[2’ + 232121432) + T. 1142(2) + (T: + 3T3)M1(2)]I) (3.35) and 011" = 1+<|TIT1I>+<|TJ (T. +17?) |>+<I(T§2]+Za)1(T1T2+éTf)I), (3.36) with T3] and 23 defined by Eqs. (2.61) and (3.24), respectively, and T1 and T2 obtained in the ECCSD/QECCSD calculations. The ECCSD(TQ)/QECCSD(TQ) energies are calculated as 23319110051111”) = 1140(2) + N/D, (3.37) where N111” = N1“ + é<I(T2‘ )1 [M.(2) + T1M3(2) + (T2 + 1T3)M2(2) +(T1T2 + éT?)M1(2)]I> (3.33) and 011‘” = D11" + %<|(T2* H17“: + H.171 + #312), (3.39) with N (T) and Dm defined by Eqs. (3.35) and (3.36), respectively, and T1 and T2 obtained in the ECCSD/QECCSD calculations. The ECCSD(T)/QECCSD(T) methods reduce to the CR-CCSD(T) approach of Refs. 11 and 30 (cf. Section 3.1) if the T1 and T2 clusters originating from the 58 ECCSD/QECCSD calculations are replaced in Eqs. (3.34)—(3.36) by their stan- dard CCSD values. Similarly, the ECCSD(TQ)/QECCSD(TQ) methods reduce to the CR—CCSD(T),b approach of Ref. 31 (cf., also, Section 3.1) if the T1 and T2 clusters originating from the ECCSD/QECCSD calculations are replaced in Eqs. (3.37)—(3.39) by their CCSD analogs. These straightforward relationships between the ECCSD(T)/QECCSD(T) and ECCSD(TQ)/QECCSD(TQ) methods on the one hand and the CR—CCSD(T) and CR-CCSD(T),b approaches on the other hand im- mediately imply that once the T1 and T2 clusters are determined by solving the ECCSD/QECCSD equations, the costs of calculating the ECCSD(T)/QECCSD(T) and ECCSD(TQ)/QECCSD(TQ) energies are essentially the same as the costs of the cor- responding CR-CCSD(T) and CR-CCSD(T),b calculations or their standard CCSD(T) and CCSD(TQf) counterparts. In particular, the most expensive steps of the ECCSD/QECCSD-based ECCSD(T)/QECCSD(T) calculations (if we ignore the costs of the ECCSD/QECCSD calculations) scale as 11311:. The most expensive steps of the ECCSD(TQ)/QECCSD(TQ) calculations (again, ignoring the costs of the ECCSD/QECCSD calculations) scale as either nfinf’, or 113. The ECCSD(T) and ECCSD(TQ) methods are much less practical, since the underlying ECCSD calcula- tions that provide T1 and T2 clusters have steps that scale as N10 with the system size. However, the QECCSD(T) and QECCSD(TQ) methods are very promising in this regard, since both the underlying QECCSD calculations that provide T1 and T2 clusters and the calculations of the final QECCSD(T) and QECCSD(TQ) energies, Eqs. (3.34) and (3.37), respectively, have steps that scale, at worst, as ngnf’, or n2 (M or .N'6 with the system size). We must keep in mind, however, that the QECCSD(T) and QECCSD(TQ) methods are approximations to the more complete ECCSD(T) 59 and ECCSD(TQ) approaches. The questions, therefore, are: (i) Do the QECCSD(T) and QECCSD(TQ) methods provide the results of the full ECCSD(T) and ECCSD(TQ) quality? (ii) Are the QECCSD(T) and QECCSD(TQ) methods sufficiently accurate to elimi- nate the problems observed in the standard and completely renormalized CCSD(T) and CCSD(TQ) calculations for triple bond breaking in N2? The answers to both questions can be provided if we examine the results of the benchmark ECCSD(T), ECCSD(TQ), QECCSD(T), and QECCSD(TQ) calculations for N2 shown in Tables 3.1, 3.2 and Figures 3.1, 3.2, which we performed with the computer codes developed in this thesis work". As shown in Table 3.1 and Figure 3.1, the ECCSD(TQ) approach employing the T1 and T2 clusters obtained in the full ECCSD (Arponen-Bishop) calculations is capable of providing spectacular improvements in the description of triple bond breaking in the N2 molecule, as described by the STO-3G basis set, reducing the large unsigned errors in the CCSD and CR—CCSD(TQ),b results in the R 2 5.0 bohr region, on the order of 201—213 and 39—54 millihartree, respectively, and the 31—38 millihartree errors in the ECCSD results to less than 4 millihartree. Remarkably enough, the ECCSD(TQ) results for the N2 molecule, as described by the STO-3G basis set, in the entire R _<_ 8.0 bohr (2: 4H,) region do not exceed ~ 4 millihartree, being much smaller in the R m Re region. As shown in Table 3.1 and Figure 3.1, the ECCSD(TQ) potential energy curve is located only slightly above the full CI curve and there is no unphysical bump on it. Interestingly enough, the zero-body moment Mo(2), Eq. (3.27), calculated with the ECCSD values of T1 and T2, which is corrected in the ECCSD(TQ) energy formula, Eq. (3.37), by adding terms expressed via moments 60 M311°.'.'.igk(2) with k = 1 — 4, is a poor approximation to the exact, full CI, energy in the region of large N —N separations (see Table 3.1). This demonstrates the remarkable ability of the MMCC (or GMMCC) formalism to restore high accuracies in the bond breaking region even when the CC energy that we are trying to correct (in this case, the Mo(2)/ECCSD energy) is itself very poor. The ECCSD(TQ) results for the STO-3G model of N2 shown in Table 3.1 and Figure 3.1 are very encouraging, but, as mentioned earlier, the ECCSD(TQ) ap- proach is not too practical due to the expensive N10 steps of the underlying ECCSD calculations. It is, therefore, important to examine if the much more manageable QECCSD(TQ) approximation, which relies on the T1 and T2 clusters obtained in the QECCSD calculations and which is the N“ -.N‘7 procedure, provides the results of the ECCSD(TQ) quality. The results in Table 3.1 indicate that the QECCSD(TQ) and ECCSD(TQ) energies are virtually indistinguishable when the internuclear separation R does not exceed 3.5 bohr (z 1.75Re). Only when the N—N separations exceed 3.5 bohr, the differences between the QECCSD(TQ) and ECCSD(TQ) results become larger. Although the errors in the QECCSD(TQ) results in the R > 3.5 bohr are greater than the corresponding errors obtained with the ECCSD(TQ) method, the QECCSD(TQ) method provides an excellent description of the potential energy curve of the STO-3G N2 molecule, reducing the 35—46 millihartree errors in the QECCSD results and much larger errors in the CCSD and CR—CCSD(TQ),b results in the R > 4.5 bohr region to 5—6 millihartree (see Table 3.1). As in the ECCSD(TQ) case, the QECCSD(TQ) potential energy curve of the STO-3G N2 molecule is located very close to and above the exact, full CI curve (see Figure 3.1). Thus, the QECCSD(TQ) approach provides a highly accurate description of the large nondynamic correlation effects characterizing the N2 molecule at larger N—N separations. 61 The question is if the above observations obtained for the very small STO-3G basis apply to larger basis sets. In order to answer this question, we performed the QECCSD(TQ) calculations for the DZ model of N2 (see Table 3.2 and Figure 3.2). As shown in Table 3.2, the errors in the QECCSD(TQ) results are somewhat greater than in the case of the STO-3G basis set, but the overall patterns are the same. Thus, the QECCSD(TQ) method provides considerable improvements in the CCSD and QECCSD results, reducing the large negative, (—70) — (—121) millihartree, errors in the CCSD results in the R = 2R9 —2.25Re region and the relatively large positive, 40— 50 millihartree, errors in the QECCSD results in the same region to 16—20 millihartree. For smaller values of R, the improvements offered by the QECCSD(TQ) approach are even greater. For example, the QECCSD(TQ) method reduces the 6.236 and 8.289 millihartree errors in the QECCSD and CCSD energies at R = Re and the 23.485 and 33.545 millihartree errors in the QECCSD and CCSD energies at R = 1.5Re to 1.002 and 6.011 millihartree, respectively. As shown in Table 3.2 and Figure 3.2, the QECCSD(TQ) potential energy curve for the DZ N2 molecule is located above the full CI curve and the QECCSD(TQ) approach completely eliminates the unphysical humps on the CR-CCSD(T) and CR—CCSD(TQ) curves, obtained with the CCSD values of T1 and T2. This shows that the use of the QECCSD rather than CCSD values of the T1 and T2 clusters improves the results of the MMCC calculations. All of this is very encouraging, since the QECCSD(TQ) method is a relatively inexpensive single-reference approach employing the spin-adapted RHF reference. As shown in Table 3.2 and Figure 3.2, all standard CC methods using the RHF determinant as a reference, including the very expensive CCSDT and CCSDT(Q;) approaches, which require iterative steps that scale as JV8 with the system size, completely break down at larger N—N separations. The QECCSD(TQ) method provides a much smoother 62 and more accurate description of triple bond breaking in N 2, with the relatively small errors which monotonically increase with R, while eliminating all of the pathologies observed in the standard CC calculations. Interestingly enough, even the simplest QECCSD(T) method works reasonably well, when the DZ model of N2 is examined (see Table 3.2), although one has to keep in mind that the errors in the QECCSD(T) energies in the R S 1.5Rf region are 2—3 times larger than in the QECCSD(TQ) case (see Table 3.2). Moreover, the QECCSD(T) energies do not vary with R as smoothly as the QECCSD(TQ) energies (cf. the nonmonotonic changes in errors in the QECCSD(T) results shown in Table 3.2). The same behavior is observed when the STO-3G basis set is employed. The calculations for the STO-3G basis set show that the QECCSD(T) and ECCSD(T) methods are incapable of providing significant improvements in the corresponding QECCSD and ECCSD results when the internuclear separation R becomes large (see 213(2) in the QECCSD(T) and ECCSD(T) energy expressions. It is interesting to observe, Table 3.1). This is related to the absence of the quadruply excited moments M though, that the QECCSD(T) results for the DZ model of N2 are much better than the results of the CR-CCSD(T) calculations (not to mention the CCSD(T) results), in which the quadruply excited moments M23219) are neglected too. This shows once again that the T1 and T2 clusters resulting from QECCSD (and ECCSD) calculations are of much higher quality than the T1 and T2 clusters obtained with the standard CCSD approach, improving the results of the MMCC calculations. 63 3.4 Summary We have demonstrated that the noniterative ECCSD(TQ) and QECCSD(TQ) meth- ods, obtained by merging the ECC formalism of Arponen and Bishop with the gen- eralized version of the MMCC theory that enables one to use the non-standard clus- ter components to design the noniterative corrections to CC energies, provide an accurate and variational description of potential energy surfaces involving multiple bond breaking with the ease of a single-reference “black-box” calculation. In partic- ular, the GMMCC-based ECCSD(TQ) and QECCSD(TQ) approximations employ- ing T1 and T2 clusters obtained in ECCSD and QECCSD calculations do not suffer from the non-variational collapse or unphysical behavior observed in the standard CCSD, CCSD(T), CCSD(TQf), CCSDT, and CCSDT(Qf) calculations. The use of the ECCSD and QECCSD values of the singly and doubly excited clusters has a very positive impact on improving the results of the MMCC calculations in the bond breaking region. In particular, the ECCSD(TQ) and QECCSD(TQ) methods em- ploying the ECCSD and QECCSD values of T1 and T2 clusters improve the results of the CR-CCSD(T) and CR-CCSD(TQ) calculations for triple bond breaking in N2, which also use the MMCC theory but rely on the T1 and T2 clusters obtained with the standard CCSD approach. The results obtained in this thesis show that the new GMMCC theory is a flex- ible formalism, which enables one to design the relatively inexpensive noniterative CC methods employing the non-standard values of cluster components. The gener- alized version of the MMCC theory provides us with precise information about the many-body structure of the exact energy and the corrections that must be added to standard or non-standard CC energies to recover full CI results. This is particularly valuable in situations involving multiple bond stretching or breaking, where the stan- 64 dard arguments based on MBPT fail due to the divergence of the MBPT series at larger internuclear separations. The fact that one can use the standard as well as non-standard cluster amplitudes in the GMMCC calculations is a very useful feature, which gives us an opportunity to improve the results by using non-traditional sources of cluster amplitudes that are more suitable for the applications of interest. The results obtained in this thesis, in which we used the ECCSD and QECCSD methods to generate the T1 and T2 clusters for the GMMCC calculations, are a clear demon- stration of how useful the idea of using the non-standard values of cluster amplitudes might be in the most challenging cases involving multiple bond breaking where all standard single-reference CC methods (including high level methods such as CCSDT or CCSDT(Qf)) fail. We tested a few ECCSD- and QECCSD-based GMMCC approximations, in- cluding ECCSD(T), QECCSD(T), ECCSD(TQ), and QECCSD(TQ). Although the best results are obtained with the ECCSD(TQ) method, we cannot recommend this method at this time for practical calculations due to the large costs of the under- lying ECCSD calculations. The QECCSD(TQ) approach employing the QECCSD values of T1 and T2 clusters is more promising in this regard, offering an accurate description of large nondynamic and substantial dynamic correlation effects with the computational steps that scale as N6 — .IV7 with the system size. 65 .3“: am .3 82%: 3: e2 as 23 2.... .6 822 800m 2: :22: 22.23.: :82 @6822: 2E... .323 am 3 25:2. 2.: 22 as 22 C. co 32.? omoomc 2: :22. 8:22.. :32: 3.300220 25 .32: em .3 22:2. 2.1 28 as 23 .5 :o 832 800m 2: 22.2: 2.22% :32 3.30022: 2:: .82: em 3 82:3 2.2 23 as 23 .s s 8:? omoomo 2: 222. 822% 2:82 3.300220 23... .NE 98 mu mo $3.? Dmoom 2: mafia 83328 45.8 .5 6068538 3.8:: 85de0 on» .8 3:89: 35:68: 23.: .NE was Cu mo 8:?» QmOOMO 2: mafia @3an18 5.68 .cm £068.:me 3.8:: SEA—moo one .8 3:82: mgaéumu 235 .38 $3 2 m :0 2:2, 225528 2a. .38 2 8:53... 2.2 2a... 6:833:33 Bad—ounce 2 88: 22: 2238 8388 25 :83 2a. 59:83.2 .32 o.» 22 .3 .3. .3 5.... .2. .3. .3 a..." .2 .3. .2 u m a 22:2: 883.37 23 «8332: $83.87 583321 53323:: £33.87 £23.37 23.3237 £83.37 £38.87 53.5.37 583.87 23 833 22:2, 923 6 =3 3:82.328 2: 3 2:22 2:23:52 2 23 32:38 3... m5.” azé wfiodm wwafim mag: wawaH Sfiwm womdv 3””.va Hmvdfiml 9w wved :56 $9.3 $3.3 Swag :5.me :bNm amwév wmfidml mSAHNI cs vac.” wowd Svdm 3%.; vowém: mad: mmoém 8ch «8.va wmmdoml c6 :3.” 3.9m Stag mmmdm ezfima 3353 wwfivm mmadm «3.ch Eméaml m6 awn.” and #5.: www.mw @9162 95.”: 23.8 www.mm Hmmdml umwdoml 9m 3”.” m3: wand“ 3de 25.2: vva: wafivm NEE.“ hmméml vwaéwfil m6 onad 3m.” wSA wad www.mc 25.8 mid; magma omwdl SEdEI o6 a3.” «mud Seél mwadl 58.: wmoéfi 35.: gm...” wwfiw 39%| 0m main mm?” mama. Sod onwél ummél RE.” moafi mwvd 3%.? 9m aha; c3; movd 3:6 $.de mmhdl haw.” mvvd SYN ommd md «Ed «36 gm; 8w; ~36 momd Swé mam; god mam.” ON mafia mafio vcvd vovd 35.: 3nd mmwd cwwd wand «34 m4 66 259580: 23.5803 .588: 295885 388E632 38885:: 88: 8025 23580.5 :80 3: 58:5 00m— aoaflmémaoedw 23 4:3 Entered :83 32.5338 veg—mudom =< 3.6: mag OTOHm 23 firs @ 3038320. Ham—osfimufi _Egom no: 353% 2:838 «Z 2: mo 8&8qu 83?:52w 25. ”Ad @359 Table 3.2.: The ground-state energies of the N 2 molecule obtained for several inter- nuclear separations R with the DZ basis set.“ Method 0.7512. 12..b 1.2512. 1.512. 1.7511. 2R. 2.2512. CCSD 3.132 8.289 19.061 33.545 17.714 -69.917 —120.836 CCSDTc 0.580 2.107 6.064 10.158 —22.468 —109.767 —155.656 CCSD(T)“ 0.742 2.156 4.971 4.880 —51.869 —246.405 —387.448 CCSD(TQ.)c 0.226 0.323 0.221 —2.279 —14.243 92.981 334.985 CCSDT(Q;)d 0.047 —0.010 —0.715 —4.584 3.612 177.641 426.175 CR—CCSD(T)C 1.078 3.452 9.230 17.509 -2.347 —86.184 -133.313 CR-CCSD(TQ),ac 0.448 1.106 2.474 5.341 1.498 -40.784 -69.259 CR—CCSD(TQ),bc 0.451 1.302 3.617 8.011 13.517 25.069 14.796 QECCSD 2.506 6.236 13.609 23.485 31.060 40.085 49.741 Mo(2)/QECCSD° 1.908 4.164 7.935 13.623 33.109 70.120 106.399 QECCSD(T)‘ 1.000 2.941 7.121 11.661 8.454 10.330 17.977 QECCSD(TQ)S 0.412 1.002 2.443 6.011 11.393 ' 16.103 19.958 “All energies are in millihartree relative to the corresponding full CI energy values, which are —108.549027, -109.105115, -109.054626, —108.950728, —108.889906, —108.868239, and —108.862125 hartree at R = 0.75Re, Re, 1.25Re, 1.5Re, 1.75Re, 2B.” and 2.25Re, respec- tively. The lowast two occupied and the highest two unoccupied orbitals were frozen in correlated calculations. bThe equilibrium value of R, 1?.: = 2.068 bohr. cFrom Ref. 31. dFrom Ref. 34. eThe zero-body moment or the CCSD-like energy expression, Eq. (3.27), calculated using the QECCSD values of T1 and T2 (obtained with the Arponen-Bishop ECC theory). fThe GMMCC(2,3) result obtained using the QECCSD values of T1 and T2 and |\Ilo) defined by Eq. (3.23) (obtained with the Arponen-Bishop ECC theory). “The GMMCC(2,4) result obtained using the QECCSD values of T1 and T2 and |\Ilo) defined by Eq. (3.25) (obtained with the Arponen-Bishop ECC theory). 67 -107.35 r I ' I ' I V v f r 7 -107.40 - :9 . fl” 4; '_‘—:;:m‘M-‘ -107.45 - f A m E *~)|$__*__ t —107.50 - - 2 Q .FulICl V 4 \ 00080 < 5‘: \ .'*'CR-CCSD(TQ).b 5 -107.55 - '\ Aoeccso - c " VECCSD u.| . uosccsocro) b QECCSD(TQ) —107.60 r \ . \ §>\ * l ‘- eke» -107.65 9---<>--- -107.7o-‘A--‘-‘-‘A 2.0 3.0 4.0 5.0 6.0 7.0 8.0 RN_N (bohr) Figure 3.1: Ground-state potential energy curves of the N2 molecule as described by the STD-3G basis set. The QECCSD and ECCSD calculations were performed using the ECC theory of Arponen and BishOp. 68 7"er 'tvr vvvv 'i" rii" fit" 7 U I Y I T I -108.5 q Full Cl -108.6 r ——— ccso I - alt-CCSDT I ’ ECCSD(T) / _ . CCSD(TQ.) I . 1 08'7 AQECCSD ,' Voeccscxm) ¢ ‘ -108.8 L -108.9 - Energy (hartree) —109.0 - -109.1 - -109.2 - _109.3 ..LLL.;L.1..-.1....I....1...;1.. 1.5 2.0 2.5 3.0 3.5 4.0 4.5 RN_N(bohr) Figure 3.2: Ground-state potential energy curves of the N2 molecule as described by the DZ basis set. The QECCSD calculations were performed using the ECC theory of Arponen and BishOp. 69 Chapter 4 Exactness of Two-Body Cluster Expansions in Many-Body Quantum Theory As demonstrated in Chapters 2 and 3, we can considerably improve the description of chemical bond breaking and quasi-degenerate electronic states by performing ECCSD calculations and by adding new types of noniterative energy corrections, based on the GMMCC formalism, to the ECCSD energies. The ECCSD approaches and the ECCSD-based GMMCC approaches are capable of eliminating the failures of the standard CCSD, CCSD(T), and CCSD(TQf) methods, in spite of the fact that we only use one— and two—body cluster operators in the ECCSD and ECCSD-based GMMCC calculations. Our positive experiences with the ECCSD-based methods imply that there is a lot of unexplored flexibility in the CC theories using only one- and two-body clusters. The natural question arises if one can improve the quality of many-electron wave functions based on the cluster expansions involving one— and two—body (or only two-body) operators to a degree where the results become exact or virtually exact. After all, the many-electron Hamiltonians used in quantum chemistry and atomic and molecular physics do not contain higher—than-two—body terms. Thus, one may wonder if there is a way to obtain the exact or virtually exact description of many- electron systems with the CC—like theories that does not use higher—than—two—body operators to construct the corresponding many-particle wave functions. An issue of the exactness of the exponential cluster expansions employing two-body operators is 70 discussed in this chapter. 4.1 Theory It is well known that one can always obtain the exact solution of the electronic Schrodinger equation within a given basis set by performing the full CI calculations. Unfortunately, the dimension of the full CI eigenvalue problem can easily run into astronomical figures, even for small many-electron systems. This can be easily veri- fied by applying the well-known Weyl’s formula97 for the number f (n, N, S) of spin- adapted electron configurations that enter the full CI expansion of an eigenstate of the electronic Hamiltonian H, i.e., 2S+1 n+1 n+1 f(n’N’S)=m(N/2—S)(N/2+S+1)’ (4'1) where n represent the number of correlated orbitals, N is the number of correlated electrons, S is the total spin of an eigenstate under consideration, and (’1?) = m!/[k!(m — k)!] is the binomial coefficient. For example, a full CI calculation of a sin- glet electronic state for a system consisting of 10 electrons (e.g., the HF molecule or the Ne atom) and described by only 20 orbitals requires using f (20, 10, O) = 52, 581, 816 configurations. A modest increase of the number of orbitals from 20 to 30 results in a steep increase in the number of configurations defining the full CI problem to 4.04 x 109. Those numbers should be compared with the much smaller numbers of 22,155 and 108,345 two-electron integrals defining the electronic Hamiltonians for the n = 20 and 30 cases, respectively. The point-group symmetry and other symmetries of the Hamiltonian can, on occasion, reduce the dimension of the full CI problem, but savings resulting from the use of symmetry are minimal compared to the rapidly in- creasing numbers of configurations defining the full CI wave functions with n and N. 71 Indeed, the realistic calculation for a 10 electron system would require using ~ 100 orbitals. According to Eq. (4.1), the n = 100, N = 10, and S = 0 case leads to an astronomical number of 9.94 x 1014 configurations in the corresponding full CI expan- sion. This should be compared to a much smaller number of 12,753,775 two-electron integrals defining the electronic Hamiltonian for the n = 100 case. Clearly, the num- bers of full CI configurations would be significantly larger for N > 10 and for larger n values due to the factorial scaling of the dimension of the full CI problem with the system size. This should be contrasted with the relatively slow, n4-like, increase of the number of two-electron integrals defining the Hamiltonian H with n. Thus, there seems to be a conflict between the huge dimensionality of the full CI problem, which prevents us from performing exact ab initio calculations for larger systems, and the fact that the Hamiltonians for many-electron systems involve only one- and two-body integrals, whose numbers are much smaller than the numbers of full CI coefficients defining the exact wave functions. In spite of the tremendous progress in the area of full CI calculations and computer technology (cf., e.g., Ref. 98 and references therein), the exact ab initio calculations employing the full CI method remain limited to a few electron systems described by small (usually, n ~ 20 — 30) basis sets. 4.1.1 The exp(X) Conjecture It has recently been suggested that it may be possible to represent the exact or virtually exact ground-state wave function of an arbitrary many-fermion pairwise in- teracting system by an exponential cluster expansion involving a general two-body 99-105. If these statements were true, completely new ways of performing Operator ab initio quantum calculations for many-fermion (e.g., many-electron) systems might be suggested, which could provide enormous reductions in computational require- 72 ments for accurate quantum calculations for pairwise-interacting many-fermion sys- tems, eliminating the astronomical costs of generating the exact many-particle wave functions by solving the full CI eigenvalue problem. Specifically, it has been proposed that the exact ground-state wave function I‘IIO) of a given many-fermion system de- scribed by the Hamiltonian, 1 H = zgcch + ivggopcqcscr, (4.2) containing up to two-body terms, obtained in a finite spin-orbital basis set, has the following simple form”: No) 5 I‘I’0(X)) = exlq’o), (43) where X is a general two-body operator and |o) is a normalized reference state, which in principle is an arbitrary wave function that has a nonzero overlap with |\Ilo), but in practice should provide us with a reasonable approximation of I‘IIO). In Eq. (4.2) and equations presented below, we use, whenever possible, the Einstein summation convention over repeated upper and lower indices. As in the earlier parts of this thesis, the op and op operators represent the usual creation and annihilation operators, respectively, associated with the one particle basis {p}. The 25% = (plilq) and 2253 = (pqlfilrs) represent the usual one- and two-particle integrals defining the Hamiltonian. In the language of second quantization, 1 X=X2=2 £3019 cqcscr, (4.4) where 23% are some coefficients. According to Nooijen99 (cf., also, Refs. 104, 105), the number of independent coefficients 2:53 should be identical to the number of two- particle integrals v53 entering the Hamiltonian H, Eq. (4.2). One could redefine the operator X by considering the one- and two—body components in Eq. (4.4)99’101'102 73 and write 1 X = X1 + X2 = xgopql + Exacpcqcscr, (4.5) but this is not really necessary, since for a fixed number of particles (N), one can always rewrite the Hamiltonian H, Eq. (4.2), in terms of two-body terms only. A straightforward manipulation shows that l H 2 Ehggcpcqcscr, (4.6) where fig}; 2 v53 + (2;,63 + 6523)/(N — 1), (4.7) with 6;], representing the usual Kronecker delta. On the other hand, it may be bene- ficial to use Eq. (4.5) rather than Eq. (4.4) in actual calculations, since the presence of one-body term X1 = xgcpccl in the operator X may accelerate the convergence of the resulting wave functions and energies. The above representation of the exact ground-state wave function, Eq. (4.3), is reminiscent of the exponential ansatz of the single-reference CC theory, Eq. (2.4). There is, however, a fundamental difference between Eqs. (4.3) and (2.4). The cluster operator T entering Eq. (2.4) is defined in terms of the particle-hole excitation operators Exist-2", Eq. (2.7), where i1, . . . , in (a1, . . . , an) are the spin-orbitals that are occupied (unoccupied) in the reference configuration |o) (which is then a single Slater determinant), and it contains all many-body terms Tn with n = 1,. . . , N in the exact case. The operator X, Eq. (4.4), or its analog defined by Eq. (4.5), entering Eq. (4.3), has at most two-body terms, but of the general type (excitations, deexcitations, and all other combinations of indices p, q, r, 3). Moreover, the reference configuration |o) can be a multi-determinantal state. One could, of course, truncate the many- body expansion for the cluster operator T in Eq. (2.4) at a two-body component 74 T2, Eq. (2.27), but then the resulting wave functions eT2|o) and eT1+T2|o), which are used in the CCD and CCSD methods, respectively, are only approximate wave functions. Thus, as we can see, there are significant differences between Eqs. (4.3) and (2.4). However, because of the formal similarity of the CCD wave function, eT2|o), and Eq. (4.3), Nooijen and Lotrich106 and Van Voorhis and Head-Gordon104 call the wave function ansatz defined by Eqs. (4.3) and (4.4) generalized CCD (GCCD) (the wave function ansatz using Eqs. (4.3) and (4.5) is then called generalized CCSD or GCCSD). A similar terminology has been used in Refs. 100—103. 4.1.2 Formal Arguments in Favor of the exp(X) Conjecture (Ground States) There are several facts that speak in favor of the correctness of Eq. (4.3). Nooijen based his reasoning99 on the fact that the number of two-body coefficients $53 is iden- tical to the number of components of the Nakatsuji two-particle density equation107, which is, in turn, equivalent to the time—independent Schriidinger equation for Hamil- tonians containing up to two-body terms. The problem that was left unsolved by Nooijen is the solubility of a rather complicated exponential variant of the Nakatsuji density equation, which forms an essential part of Nooijen’s analysis (cf. Eq. (11) in Ref. 99). Van Voorhis and Head-Gordon based their reasoning104 on the fact (ex- ploited in Quantum Monte Carlo techniques) that one can always obtain the exact wave function by considering the expression I‘I’o) = £13330 eztléo), (48) where the two-body operator Z, is defined as follows: Z, = —(H - Eo)t, (4.9) 75 with E0 representing the exact energy. They used Eq. (4.8) to write the two-body operator X defining the exact wave function |\Ilo) via Eq. (4.3) as t—mo Similar arguments and equations have been presented by N akatsujimr‘“103 , who also considered the conditions for the wave functions parameterized by two-body oper- ators to be exactloo. The problem with using Eq. (4.8) in this fashion is that the operator 2,, Eq. (4.9), provides the exact wave function only in the t —+ 00 limit, whereas the Operator X, Eq. (4.4) or (4.5), entering Eq. (4.3), is a finite operator. This immediately implies that the operator X defining the exact wave function |\Ilo) through Eq. (4.3) cannot be constrained to be of the Hamiltonian form, Eq. (4.9) (cf. below for additional remarks). Eq. (4.8) does not Open up the possibility of the existence of a finite two-body operator X, which is not necessarily defined through infinite coefficients 3;; (which Eq. (4.10) produces). There also seem to exist some contradictions between statements made in Refs. 100 and 102, 103. The exactness of the GCCD or GCCSD wave functions was questioned in Ref. 100 and supported in Refs. 102, 103 (see Refs. 108—112 for further debate; see the discussion below for additional comments). We have recently provided an evidence105 that the exact ground state of a many- fermion system, described by the Hamiltonian containing one- and two-body terms, may indeed be represented by the exponential cluster expansion employing a general two-body operator, Eq. (4.3), by connecting the problem with the Horn-Weinstein formula for the exact energym, E0 = tl_i+m ((pole—tHHI‘IHD/(cPMC-‘tflI‘po) = 11m Em; =31“) E(Xt), (4.11) t—mo 76 where 130.2 E(X.) = (eolexi‘ HextIO>/<olexi‘extl<1>o> (4.12) and 1 X1: —§tH, (4.13) and by determining the Operator X entering Eq. (4.3) through a direct minimization of the expectation value expression E00?) = <ole*’He*Io>/<ole*‘e*Io> (4.14) over general two-body Operators X = -~ggcpcqcscr (4.15) This analysis (taken from Ref. 105) is described below. Let us consider the family M of all two-body operators X , Eq. (4.15), that are defined by finite coefficients 5;}; and that have a general structure of the Hamiltonian H, Eq. (4.6). This means that M consists of all two-body operators that are, for example, Hermitian, since H is Hermitian; that satisfy relations, such as 5:53 = 53;), since has] = hag, etc. Obviously, the number of independent parameters 553 is identical to the number of coefficients hgg or 1253 defining the Hamiltonian. It should be noticed that all operators Xi, Eq. (4.13), and Z,, Eq. (4.9), belong to M, although M is a much larger operator family, which contains infinitely many Operators that are not multiples of H. This remark is important for the considerations discussed in this section, since one can always Obtain the exact wave function and energy by applying Eqs. (4.8) and (4.11), with Z, and X; defined by Eqs. (4.9) and (4.13), respectively. As pointed out above, neither the operator Z, nor its X, analog can provide the exact description of a many-fermion system for a finite value of t. We should search 77 for the operator X defining |\Ilo) via Eq. (4.3) by minimizing the expectation value expression, E0(X), Eq. (4.14), over all Operators in M. Let us, therefore, examine what the direct minimization of E0(X) in M leads to. According to the Ritz variational principle, E0(X) is bounded from below by the exact, full CI, energy, so that &S%d) mm for all operators X E M. This implies that there should exist a two-body operator X E M that minimizes E0(X). We can write E0(X) = min E0(X). (4.17) XEM Obviously, E0 3 E0(X). (4.18) Let us now consider the energy expression E0,“ Eq. (4.12), for an arbitrary (fixed) value of t. We can immediately write, E0(X) < Eat, (4.19) since E0(X) is a minimum value of E0(X), Eq. (4.14), in a space of all two-body Operators X, whereas E0; = Eo(X¢) is the value Of E0(X) at X = X, (cf. Eqs. (4.14) and (4.12)). As a matter of fact, for a given value of t, one can always find a two-body Operator Y from M such that E0(Y) < Em. An example of such operator might be provided by Xt: with t’ > t, since, as shown in Ref. 113, E0,“ Eq. (4.12), is a monotonically decreasing function Of t. However, since the Operator family M is much larger than the “one-dimensional” manifold of Operators X t, which are multiples of H, there is a chance that there exist two-body Operators Y E M which satisfy E0(Y) < Em: and which are not given by Eq. (4.13). This indicates that the Operator X 78 minimizing E0(X ) may very well be a finite Operator (i.e., defined by finite coefficients x53 and not Obtained by considering the limiting case of the t —+ 00 operators Xt), although we cannot provide a rigorous mathematical proof that this is indeed the case and the existence of finite Operator X may depend on the actual form of the reference state |o), which does not have to be a single Slater determinant (see, e.g. Refs. 108-110). The existence of a finite Operator X E M that minimizes E0(X) according to Eq. (4.17) and that is not of the Hamiltonian form is supported by the numerical calculations for a few many-electron systems105 (see the discussion below). The inequalities (4.18) and (4.19) can be combined into the following result: E0 S 13000 < Eat, (4-20) true for any value of t. In view of the Horn-Weinstein energy expression, Eq. (4.11), by considering the t —> oo limit in Eq. (4.20), we Obtain the identity E0 = E0(X)- (4.21) This means that the two-body operator X, obtained by minimizing the expectation value expression E0(X), Eq. (4.14), gives the exact energy E0 and, by the virtue of the variational principle, the exact ground state |\Ilo), as stated in Eq. (4.3). The above analysis makes the exactness of Eq. (4.3) a real possibility, but one should not treat it as a complete mathematical proof of Eq. (4.3), since we cannot rig- orously prove the existence of the finite coefficients 1:53 that would define the optimum Operator X corresponding to a global minimum of E0(X) for an arbitrary reference |o). We can make, however, several useful Observations. First, the reasoning pre- sented above, which is based on combining the Horn-Weinstein energy formula, Eq. (4.11), with the minimization of E0(X), has an advantage over the arguments given in Refs. 102—104 in that it frees us from necessarily assuming that the operator X can 79 only be obtained by studying the t —+ 00 Operators Xt, Eq. (4.13), or Z,, Eq. (4.9). By minimizing E0(X), Eq. (4.14), in a space of all two-body operators (or, equiva- lently, by minimizing E0(X) in a finite-dimensional space of variables :2”), which is exactly how we obtained Operator X in Ref. 105 (cf. the examples below), we may be able to find finite parameters 113%, defining the exact wave function |‘Ilo), precisely because the operator X is not constrained to be a multiple of the Hamiltonian. If the finite operator X, determined by some numerical procedure for minimizing E0 (X) in M, is a local rather than a global minimum on the E0(X) multi-parameter surface, then the resulting energy E0(X), calculated by substituting X = X into Eq. (4.14), and the resulting wave function |\I!o(X)), Eq. (4.3), do not have to be exact. How- ever, even in this case, the operator X may provide excellent results, opening up a possibility of using the exponential wave functions (4.3), with X defined by Eq. (4.4) (or (4.5)), in high-accuracy ab initio calculations. Second, the mathematical analysis described above does not tell us anything about the specific form of the reference state |o) that should be used in the calculations exploiting the exp(X) ansatz. It is possible, for example, that the finite parameters 2:53 that give the exact state |\IIo(X)) via Eq. (4.3) exist only for certain types of references |o) that do not have to be represented by a single Slater determinant (based on the analyses presented in Refs. 108—110, it is rather unlikely that |o) is a single determinant although there is no proof of this statement). It may happen, however, that highly accurate results are already obtained with the ordinary Hartree—Fock reference Km) and the additional improvements are obtained by using a multi-determinantal reference state |o) in the definition of |\II0(X)), Eq. (4.3). All Of these issues are explored here by preforming numerical calculations for a few small many-electron systems. 80 4.1.3 Extension Of the exp(X) Conjecture to Excited States The exp(X) conjecture was originally proposed for the ground states, but we can easily extend it to the excited states using the standard variational approach. For example, we can construct the trial wave function for the first excited state |\I!1) as WW» = |\I’1(X“’)) — (wolw1(X<”)>Iwo>. (422) Here, In» =Noexw)|‘i’o) (4.23) is the previously obtained normalized ground state (.No is the normalization factor, X (0) is the optimum two-body Operator representing |\Ilo), and I‘io) is the single- or multi-determinantal reference state used in the calculation of |\I'o)) and l‘pl(X(l))):N1€X(l)|(i’l> (4.24) is the normalized exp(X)-like form of the first excited state, where |1) represents the reference state for |\Ill). By minimizing the energy functional :__ (@1(X‘1’)IHI@1(X‘”)> ~ (1) EM ) <@.(X<1>)I®.(X)> (4.25) over all two-body Operators X (I) 6 M, we Obtain the energy of the first excited state and the corresponding wave function |\II1) E |\P1(X(1))), where X (1) is the Optimum value of X (1) obtained by minimizing E1 (X (1)), Eq. (4.25). For the k-th excited state, we can define the trial wave function for variational calculations as k—l likd‘k’» = I‘l'k(X"")> — Z»lwm>. (4.26) m=0 where |\Ilm), m = 1, . . . , k — 1, are the previously computed states and WW) =N.e*‘*’|.> (4.27) 81 is the normalized exp(X)-like form of the k-th excited state I‘IIk), with |k) rep- resenting the reference state for the k-th excited state. By minimizing the energy functional _ (‘i'k(X(k))|Hl‘i’k(X(k))) (‘1’ka "")l\i'k(X (*0) BAX“) (4.28) over all two-body Operators X“) E M, we obtain the energy of the k-th excited state and the corresponding wave function l‘IIk) E |\II,,(X("))), where X (k) is the optimum value of X (k). Since excited states of many-electron systems are almost always very multi-determinantal and since they often have significant singly excited components, we may improve accuracies and accelerate the convergence of the numerical calcula- tions for excited states using Eqs. (4.22)—(4.28) by using Eq. (4.5) (rather than Eq. (4.4)) to represent operator X“), while Obtaining the reference states ling) by diag- onalizing the Hamiltonian H in a small space spanned by a few Slater determinants that we believe are important to describe the ground and excited states of interest. This is what we did in our excited-state calculations described below. 4.2 Numerical Results In our calculations for a few many-electron systems, we first concentrated on the fol- lowing two aspects of theory: (i) the existence of finite coefficients x53 defining the two-body operator X, Eq. (4.4), and (ii) the non-Hamiltonian nature of the Operator X. Initially, we focused on the ground-state theory. We tested the exact ground- state theory, in which we used the unexpanded form Of the exponential operator ex to define the E0(X) energy rather than the truncated power series expansion in X. This was made possible by representing the operators H and X as matrices in the 82 finite-dimensional N-electron Hilbert spaces relevant to a molecular system under consideration (using all symmetry-adapted Slater determinants |o) and @313“), n = 1,. . . ,N, defining the full CI problem, as basis states). In order to calculate the exact value of E0(X), Eq. (4.14), in a given iteration of the numerical proce- dure used to minimize E0(X), we first diagonalized the matrix representing X with some unitary matrix (7 to obtain the diagonal matrix D = 0X0“. Next, we con- structed eb by taking the exponentials of the diagonal elements of D. Finally, after constructing eb, we calculated the matrix representing ex as U‘leDU and applied it to a column vector representing |o) to Obtain |\IIO(X)) according to Eq. (4.3). The value of 13.02) was obtained by calculating (\IIO(X)|H|\II0(X))/(\IIO(X)|\II0(X)), using the matrices representing H and |\IIO(X)), as described above. In order to determine the Optimum Operator X and the corresponding energy E0(X), we used the downhill simplex method82 to minimize the energy E0(X) in M (recall that M is a family of all two-body operators). Typically, our calculations required ~ 100 iterations to Obtain a reasonably converged result, although in some cases we had to iterate much longer, particularly if we wanted to determine high decimal places. We realize that the numerical procedures used to obtain the Optimum Operator X are not suitable for routine calculations or for larger many-electron systems. The main point Of this study is addressing the question if the exact or virtually exact many-fermion wave functions can be represented by Eq. (4.3). In this context, the efficiency of the numerical procedures used to determine X is of the secondary importance. Our test calculations were performed for a few many-electron systems, including, among others, the 8-electron/16-spin-orbital H8 modell”, which consists Of eight hydrogen atoms arranged in a distorted octagonal configuration and described by the minimum basis set (MBS) obtained by placing one 3 function on each hydrogen 83 atom. An example of the H8 model is very important for the discussion of the exactness of Eq. (4.3). This model provides us with a highly nontrivial situation, where the number of independent two-body parameters x1733, defining operator X, or the number of independent two-body integrals hf’g’ defining the Hamiltonian, is considerably smaller than the dimension Of the corresponding N-electron Hilbert space. Indeed, the number of two—body parameters 253 defining operator X is in this case 186, whereas the number of all spin- and symmetry-adapted configurations defining the corresponding full CI problem is 468. The one-body parameter 3;], that enter Eq. (4.5) do not change this situation. The total number of one- and two-body parameters defining X, Eq. (4.5), is 198, which is still a lot less than the number of full CI configurations describing the exact wave function. The H8 model is described by a single parameter a (in bohr), which describes the deviation of the geometry of the Dgh-symmetric H8 system from the regular octagonm. The following values of a were particularly important for testing: a = 1.0 and a = 0.0001. The a = 1.0 H8 system is somewhat less demanding, since in this case the exact ground-state wave function is dominated by the RHF configuration |o). The more demanding a = 0.0001 H8 model is characterized by a strong configurational quasi-degeneracy of the ground electronic state involving the RHF reference MM) and the doubly excited configuration |1) of the HOMO-LUMO typel”. The results of our calculations for the ground-sate of the MBS H8 system are shown in Table 4.1. As we can see, the exp(X) ansatz gives remarkably accurate results for the ground-state of H8, even when the overlap of the reference state is as small as 0.668268, as is the case when a = 0.0001 and we use the RHF determinant as a reference. Independent of the value of a, we obtain microhartree accuracies, which are a lot better than the accuracies resulting from various traditional CI and 84 CC calculations. Our exp(X) calculations for the ground state employing two-body or one- plus two-body Operators X are order of magnitude more accurate than the CISD and CCSD calculations, which also use one- and two-body operators only. In fact, the exp(X) calculations provide a significantly better description Of the ground state than the CISDTQ calculations employing one-, two-, three-, and even four- body excitations. The number of parameters used to describe the CISDTQ wave function is much bigger than the number Of parameters used to define our exp(X) wave functions. The results of our exp(X) calculations employing at most two-body operators are as good as the CCSDTQ results. In fact, by using one- and two-body parameters in X and the two-determinantal reference I430) (see Table 4.1), we Obtain the results which are considerably better than those obtained with CCSDTQ, when the challenging case of the quasi-degenerate a = 0.0001 H8 model is examined. As we can see, the presence of one-body operator X1 = xgcpql in X improves the convergence toward the full CI results. The use of a two-determinantal refer- ence state I450), which reflects the predominantly two-determinantal character of the ground-state wave function at a = 0.0001, has a positive impact on the accuracy of exp(X) calculations. Even if our numerical procedures do not produce the exact (in a mathematical sense) energies, the microhartree accuracies, obtained with the exponential cluster expansions involving up to two—body operators only, are truly in- triguing. The parameters 2:53 or 3% and 2:53 defining the Optimum Operators X are finite. For example, if we Optimize X assuming that X = X2 = é-mggcpcqcsor, the largest values if 1:53 obtained in our ground-state calculations are 0.316845 at a = 1.0 and 0.596623 at a = 0.0001. The corresponding operators X do not commute with the Hamiltonian, so that the Optimum X operators producing the highly accurate results in Table 4.1 are not of the Hamiltonian form. 85 In Table 4.2 we show the results of the exp(X) calculations for the first excited state of the 1A9 symmetry. This state is dominated by the RHF configuration |o) and the doubly excited determinant of the HOMO-LUMO type |1) in the quasi- degenerate, a = 0.0001, region, so that it is natural to choose a two-determinantal state |1) = c01|o) + c11|1) as a reference in our exp(X) calculations (see footnote “a” in Table 4.2). The exp(X) calculations for the first excited 1Ag state were per— formed using the numerical procedure described in Section 4.1.3. To facilitate our numerical effort, we considered the truncated form of the exp(X) expansion, where the exp(X ) series is truncated at the X 5° term. Since the Optimum coefficients 22% and $53 are rather small and X " enters the exp(X) expansion as iX", the truncation of the exp(X) expansion at the 5%X50 term produces essentially no errors (errors that cannot be detected in numerical calculations with the double precision Fortran). As shown in Table 4.2, the exp(X) calculations for the first excited 1A9 state of H8 at a = 0.0001 produce microhartree accuracies. For oz = 1.0 we Obtained a somewhat larger error on the order of 0.4 millihartree, but this can be understood if we real- ize that the reference state |1) is not a very good representation of the exact first excited-state wave function in this region of a. We must realize, however, that this error is on the same order as the error resulting from the equation-of-motion (EOM) CCSDTQ calculations (for basic information about EOMCC methods, see, e.g., Refs. 9, 60, 63), which use a much larger number Of parameters to define the corresponding wave function than the exp(X) expansion employing one- and two-body parameters only. For a = 0.0001, the exp(X) calculations give a microhartree-level accuracy for the first excited 1A9 state of H8. None Of the conventional C1 or CC methods, including CISDTQ or even EOMCCSDTQ can produce the results of this quality. 86 Table 4.1.: A comparison of the ground-state energies of the MBS H8 system obtained with the exp(X)-like wave functions, where X = X2 (a purely two-body operator) or X1 + X2 (a sum of one- and two-body operators), with the exact, full CI, energies and energies obtained in various CI and CC calculations. Full CI energies are total energies in hartree and all other energies are errors relative to full CI (also in hartree). We also give the overlaps of the normalized ground-state wave functions Obtained with the exp(X) ansatz and the full CI approach. The overlaps Of the reference states |o) and I60) with the full CI ground-state wave function |\Ilo) are given for comparison. Number of parameters in the wave function a = 1'0 a = 0000] Method (wave function) Energ'es Full CI 467 —4.352990 —4.204803 Noexp(X2)|o)a 186 0.000008 0.000052 No exp(X1 + )6,.)|<1>0)a 198 0.000002 0.000020 1% exp(X1 + X2)|o)b 198 0.000005 0.000007 CISD 46 0.008396 0.022779 CISDT 146 0.007984 0.016064 CISDTQ 320 0.000254 0.000449 CCSD 46 0.000546 0.005034 CCSDT 146 0.000026 —0.008362 CCSDTQ 320 -0.000001 —0.000035 Overlaps with a full CI wave function: Noexp(X2)I(I’o)a 0.999998 0.999987 No exp(Xl + X2)|<1>o)a 0.999999 0.999995 No exp(Xl + x2)|o)b 0.999998 0.999998 |<1>o)a 0.939657 0.668268 |o)b 0.942804 0.909461 Who) is the ground—state RHF determinant. bléo) = cooléo) +c10|¢I>1), where |o) is the RHF determinant and Mn) is the doubly excited determinant of the (HOMO)2 —+ (LUMO)2 type. The coefficients coo and 010 defining the reference I‘io) were obtained by diagonalizing the Hamiltonian in a space spanned by |o) and |1) and by selecting the lower energy eigenstate of H in this two-dimensional subspace 87 Table 4.2.: Same as Table 4.1 for the first excited state of the 1A9 symmetry. Number of parameters in the wave function a = 1'0 a = 00001 Method (wave function) Energ'es Full CI 467 -3.998978 —4.144027 M exp(X1 + X2)|1)a 198 0.000405 0.000020 CISD 46 0.059314 0.042374 CISDT 146 0.031697 0.009726 CISDTQ 320 0.001608 0.000435 EOMCCSD 46 0.019605 0.015011 EOMCCSDT 146 —0.003174 —0.010505 EOMCCSDTQ 320 —0.000140 -0.000384 Overlaps with a full CI wave function: JV. exp(X1 + X2)I1)a 0.999998 0.999992 |1)a 0.791762 0.902255 3|1) = c01|o) +c11|1), where |o) is the RHF determinant and |1) is the doubly excited determinant of the (HOMO)2 —+ (LUMO)2 type. The coefficients cm and cu defining the reference |1) were obtained by diagonalizing the Hamiltonian in a space spanned by Km) and Mn) and by selecting the higher energy eigenstate of H in this two—dimensional subspace 83 l‘i’il- 88 4.3 Summary Let us summarize the results discussed in this chapter. By combining the theoretical arguments based on the Horn-Weinstein energy expression with variational princi- ple and numerical calculations, we demonstrated that one can obtain the virtually exact description of pairwise interacting many-fermion systems, including molecular systems, by representing their ground- and exited-state wave functions by exponen- tial cluster expansions employing general two-body or one- and two-body Operators. Based on the evidence reported in this chapter, we can conclude that the “Optimum two-body or one- and two-body operators defining these cluster expansions are fi— nite and not of the Hamiltonian form. The results discussed in this chapter confirm that one can tremendously improve the description of many-electron wave functions without using higher—than—two-body cluster operators. All of this implies that there is a lot of flexibility in the exponential cluster expansions, which was not utilized in the past. Based on the formal arguments and the extraordinarily high accuracies obtained in the calculations based on the exp(X) expansions, where X is a sum of one- and two-body components or a purely two-body operator, we can conclude that it is quite likely that one can formulate the exact or virtually exact many-electron theories based on these kinds of expansions. 89 Bibliography 90 9999”!" 10. 11. 12. 13. 14. Bibliography F. COESTER, Nucl. Phys. 7, 421 (1958). F. COESTER and H. KiiMMEL, Nucl. Phys. 17, 477 (1960). J. Cass, J. Chem. Phys. 45, 4256 (1966). J. CiZEK, Adv. Chem. Phys. 14, 35 (1969). J. CI’ZEK and J. PALDUS, Int. J. Quantum Chem. 5, 359 (1971). J. PALDUS, in Methods in Computational Molecular Physics, edited by S. WIL— SON and G. DIERCKSEN, volume 293 of NA T0 Advanced Study Institute, Series B: Physics, pp. 99~194, Plenum Press, New York, 1992. R. J. BARTLETT, in Modern Electronic Structure Theory, Part I, edited by D. R. YARKONY, pp. 1047—1131, World Scientific, Singapore, 1995. T. J. LEE and G. E. SCUSERIA, in Quantum Mechanical Electronic Structure Calculations with Chemical Accuracy, edited by S. R. LANGHOFF, pp. 47—108, Kluwer, Dordrecht, 1995. J. PALDUS and X. L1, Adv. Chem. Phys. 110, 1 (1999). T. D. CRAWFORD and H. F. SCHAEFER III, Rev. Comp. Chem. 14, 33 (2000). P. PIECUCH and K. KOWALSKI, in Computational Chemistry: Reviews of Cur- rent Trends, edited by J. LESZCZYNSKI, volume 5, pp. 1—104, World Scientific, Singapore, 2000. P. PIECUCH, K. KOWALSKI, I. S. O. PIMIENTA, and S. A. KUCHARSKI, in Low-Lying Potential Energy Surfaces, edited by M. R. HOFFMANN and K. G. DYALL, volume 828, pp. 31—64, American Chemical Society, Washington, D.C., 2002. P. PIECUCH, K. KOWALSKI, I. S. O. PIMIENTA, and M. J. MCGUIRE, Int. Rev. Phys. Chem. 21, 527 (2002). G. D. PURVIS III and R. J. BARTLETT, J. Chem. Phys. 76, 1910 (1982). 91 l5. l6. l7. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. M. URBAN, J. NOCA, S. J. COLE, and R. J. BARTLETT, J. Chem. Phys. 83, 4041 (1985). K. RAGHAVACHARI, G. W. TRUCKS, J. A. POPLE, and M. HEAD-GORDON, Chem. Phys. Lett. 157, 479 (1989). G. E. SCUSERIA, A. E. SCHEINER, T. J. LEE, J. E. RICE, and H. F. SCHAEFER 111, J. Chem. Phys. 86, 2881 (1987). G. E. SCUSERIA, C. L. JANSSEN, and H. F. SCHAEFER III, J. Chem. Phys. 89, 7382 (1988). T. J. LEE and J. E. RICE, Chem. Phys. Lett. 150, 406 (1988). P. PIECUCH and J. PALDUS, Int. J. Quantum Chem. 36, 429 (1989). P. PIECUCH and J. PALDUS, Theor. Chim. Acta 78, 65 (1990). P. PIECUCH, R. TOBOLA, and J. PALDUS, Int. J. Quantum Chem. 55, 133 (1995). S. A. KUCHARSKI and R. J. BARTLETT, J. Chem. Phys. 108, 9221 (1998). W. D. LAIDIG, P. SAXE, and R. J. BARTLETT, J. Chem. Phys. 86, 887 (1987). K. B. GHOSE, P. PIECUCH, and L. ADAMOWICZ, J. Chem. Phys. 103, 9331 (1995). P. PIECUCH, V. SPIRKO, A. E. KONDO, and J. PALDUS, J. Chem. Phys. 104, 4699 (1996). L. ADAMOWICZ, P. PIECUCH, and K. B. GHOSE, Mal. Phys. 94, 225 (1998). P. PIECUCH, S. A. KUCHARSKI, and R. J. BARTLETT, J. Chem. Phys. 110, 6103 (1999). P. PIECUCH, S. A. KUCHARSKI, and V. SPIRKO, J. Chem. Phys. 111, 6679 (1999). K. KOWALSKI and P. PIECUCH, J. Chem. Phys. 113, 18 (2000). K. KOWALSKI and P. PIECUCH, J. Chem. Phys. 113, 5644 (2000). K. KOWALSKI and P. PIECUCH, J. Mol. Struct: THEOCHEM 547, 191 (2001). K. KOWALSKI and P. PIECUCH, Chem. Phys. Lett. 344, 165 (2001). P. PIECUCH, S. A. KUCHARSKI, and K. KOWALSKI, Chem. Phys. Lett. 344, 176 (2001). 92 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. P. PIECUCH, S. A. KUCHARSKI, V. SPIRKO, and K. KOWALSKI, J. Chem. Phys. 115, 5796 (2001). P. PIECUCH, K. KOWALSKI, and I. S. O. PIMIENTA, Int. J. Mol. Sci. 3, 475 (2002). M. J. MCGUIRE, K. KOWALSKI, and P. PIECUCH, J. Chem. Phys. 117, 3617 (2002). Y. S. LEE and R. J. BARTLETT, J. Chem. Phys. 80, 4371 (1984). Y. S. LEE, S. A. KUCHARSKI, and R. J. BARTLETT, J. Chem. Phys. 81, 5906 (1984), ibid. 82, 5761 (1985) (Erratum). J. NOCA, R. J. BARTLETT, and M. URBAN, Chem. Phys. Lett. 134, 126 (1987). G. W. TRUCKS, J. NOCA, and R. J. BARTLETT, Chem. Phys. Lett. 145, 548 (1988). S. A. KUCHARSKI and R. J. BARTLETT, Chem. Phys. Lett. 158, 550 (1989). J. NOGA and R. J. BARTLETT, J. Chem. Phys. 86, 7041 (1987), ibid. 89, 3401 (1988) (Erratum). G. E. SCUSERIA and H. F. SCHAEFER III, Chem. Phys. Lett. 152, 382 (1988). S. A. KUCHARSKI and R. J. BARTLETT, Theor. Chim. Acta 80, 387 (1991). S. A. KUCHARSKI and R. J. BARTLETT, J. Chem. Phys. 97, 4282 (1992). N. OLIPHANT and L. ADAMOWICZ, J. Chem. Phys. 95, 6645 (1991). P. PIECUCH and L. ADAMOWICZ, J. Chem. Phys. 100, 5792 (1994). M. MUSIAL, S. A. KUCHARSKI, and R. J. BARTLETT, J. Chem. Phys. 116, 4382 (2002). J. S. ARPONEN, Ann. Phys. 151, 311 (1983). J. S. ARPONEN, R. F. BISHOP, and E. PAJANNE, Phys. Rev. A 36, 2519 (1987). J. S. ARPONEN, R. F. BISHOP, and E. PAJANNE, Condensed Matter Theor. 2, 357 (1987). R. F. BISHOP, J. S. ARPONEN, and E. PAJANNE, in Aspects of Many— Body Effects in Molecules and Extended Systems, edited by D. MUKHERJEE, volume 50 of Lecture Notes in Chemistry, pp. 79—100, Springer, Berlin, 1989. R. F. BISHOP and J. S. ARPONEN, Int. J. Quantum Chem. Symp. 24, 197 (1990). 93 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. J. S. ARPONEN and R. F. BISHOP, Ann. Phys. 207, 171 (1991). R. F. BISHOP, N. I. ROBINSON, and J. S. ARPONEN, Condensed Matter Theor. 5, 37 (1990). R. F. BISHOP, Theor. Chim. Acta 80, 95 (1991). J. S. ARPONEN, Phys. Rev. A 55, 2686 (1997). J. S. ARPONEN, R. F. BISHOP, and E. PAJANNE, Phys. Rev. A 36, 2539 (1987). P. PIECUCH and R. J. BARTLETT, Adv. Quantum Chem. 34, 295 (1999). E. A. SALTER, G. W. TRUCKS, and R. J. BARTLETT, J. Chem. Phys. 90, 1752 (1989). E. A. SALTER and R. J. BARTLETT, J. Chem. Phys. 90, 1767 (1989). J. F. STANTON and R. J. BARTLETT, J. Chem. Phys. 98, 7029 (1993). J. F. STANTON and R. J. BARTLETT, J. Chem. Phys. 99, 5178 (1993). S. PAL, Theor. Chim. Acta 66, 151 (1984). 3. PAL, Phys. Rev. A 33, 2240 (1986). 3. PAL, Phys. Rev. A 34, 2682 (1986). K. B. GHOSE and S. PAL, Phys. Rev. A 36, 1539 (1987). . VAVAL, K. B. GHOSE, and S. PAL, J. Chem. Phys. 101, 4914 (1994). . VAVAL and S. PAL, Phys. Rev. A 54, 250 (1996). . B. KUMAR, N. VAVAL, and S. PAL, Chem. Phys. Lett. 295, 189 (1998). . VAVAL, A. B. KUMAR, and S. PAL, Int. J. Mol. Sci. 2, 89 (2001). . VAVAL, Chem. Phys. Lett. 318, 168 (2000). w22>22 .,PIECUCH K. KOWALSKI, P. -.D FAN, andI. S. O. PIMIENTA, in Advanced Topics in Theoretical Chemical Physics, edited by J. MARUANI, R. LEFEBVRE, and E. BRANDAS, volume 12 of Progress in Theoretical Chemistry and Physics, pp. 119—206, Kluwer, Dordrecht, 2003. P. PIECUCH, K. KOWALSKI, I. S. O. PIMIENTA, P.-D. FAN, M. LODRICU- ITO, M. J. MCGUIRE, S. A. KUCHARSKI, T. KUS, and M. MUSIAL, Theor. Chem. Ace. 112, 349 (2004). P. PIECUCH, I. S. O. PIMIENTA, P.-D. FAN, and K. KOWALSKI, in Re- cent Progress in Electron Correlation Methodology, edited by A. WILSON and K. PETERSON, volume XXX of ACS Symposium Series, pp. XXX—XXXX, American Chemical Society, Washington, D.C., 2005, in press. 94 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. P.-D. FAN, K. KOWALSKI, and P. PIECUCH, Mol. Phys. (2005), in press. J. PALDUS, P. PIECUCH, L. PYLYPOW, and B. JEZIORSKI, Phys. Rev. A 47, 2738 (1993). P. PIECUCH and J. PALDUS, Phys. Rev. A 49, 3479 (1994). P. PIECUCH, R. TOBOLA, and J. PALDUS, Chem. Phys. Lett. 210, 243 (1993). W. J. HEHRE, R. F. STEWART, and J. A. POPLE, J. Chem. Phys. 51, 2657 (1969). Numerical Recipes in Fortran: The Art of Scientific Computing, Cambridge University Press, Cambridge, UK, 1992. S. A. KUCHARSKI and R. J. BARTLETT, J. Chem. Phys. 108, 5243 (1998). S. A. KUCHARSKI and R. J. BARTLETT, J. Chem. Phys. 108, 5255 (1998). T. H. DUNNING, J. Chem. Phys. 53, 2823 (1970). . KOWALSKI and P. PIECUCH, J. Chem. Phys. 115, 2966 (2001). . KOWALSKI and P. PIECUCH, J. Chem. Phys. 116, 7411 (2002). . KOWALSKI and P. PIECUCH, J. Chem. Phys. 120, 1715 (2004). PIECUCH and K. KOWALSKI, Int. J. Mol. Sci. 3, 676 (2002). xwxxx . KOWALSKI and P. PIECUCH, Mal. Phys. 102, 2425 (2004). I. S. O. PIMIENTA, K. KOWALSKI, and P. PIECUCH, J. Chem. Phys. 119, 2951 (2003). P. PIECUCH, K. KOWALSKI, and I. S. O. PIMIENTA, Int. J. Mol. Sci. 3,475 (2002). P. PIECUCH, M. WLOCH, M. LODRIGUITO, and J. R. GOUR, in Recent Ad- vances in the Theory of Chemical and Physical Systems, edited by S. WILSON, J.-P. JULIEN, J. MARUANI, E. BRANDAS, and G. DELCADo-BARRIO, vol- ume XXX of Progress in Theoretical Chemistry and Physics, pp. XXX-XXXX, Springer, Berlin, 2005, in press. P. PIECUCH, S. A. KUCHARSKI, K. KOWALSKI, and M. MUSIAL, Comp. Phys. Commun. 149, 71 (2002). M. J. MCGUIRE, P. PIECUCH, K. KOWALSKI, S. A. KUCHARSKI, and M. MUSIAL, J. Phys. Chem. A 108, 8878 (2004). M. J. MCGUIRE and P. PIECUCH, J. Am. Chem. Soc. 127, 2608 (2005). H. WEYL, The Classical Groups, Their Invariants and Representations, Princeton University Press, Princeton, N.J., 1946. 95 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. C. D. SHERRILL and H. F. SCHAEFER III, Adv. Quantum Chem. 34, 143 (1999). M. NOOIJEN, Phys. Rev. Lett. 84, 2108 (2000). H. NAKATSUJI, J. Chem. Phys. 113, 2949 (2000). H. NAKATSUJI and E. R. DAVIDSON, J. Chem. Phys. 115, 2000 (2001). H. NAKATSUJI, J. Chem. Phys. 115, 2465 (2001). H. NAKATSUJI, J. Chem. Phys. 116, 1811 (2002). T. VAN VOORHIS and M. HEAD-GORDON, J. Chem. Phys. 115, 5033 (2001). P. PIECUCH, K. KOWALSKI, P.-D. FAN, and K. JEDZINIAK, Phys. Rev. Lett. 90, 113001 (2003). M. NOOIJEN and V. LOTRICH, J. Chem. Phys. 113, 4549 (2000). H. NAKATSUJI, Phys. Rev. A 14, 41 (1976). S. RONEN, Phys. Rev. Lett. 91, 123002 (2003). E. R. DAVIDSON, Phys. Rev. Lett. 91, 123001 (2003). D. A. MAZZIOTTI, Phys. Rev. A 69, 012507 (2004). D. MUKHERJEE and W. KUTZELNIGG, Chem. Phys. Lett. 397, 174 (2004). W. KUTZELNIGG and D. MUKHERJEE, Phys. Rev. A 71, 022502 (2005). D. HORN and M. WEINSTEIN, Phys. Rev. D 30, 1256 (1984). K. JANKOWSKI, L. MEISSNER, and J. WASILEWSKI, Int. J. Quantum Chem. 28, 931 (1985). 96 uhyiyyyl((5)