Skip to main content
Log in

On the off-diagonal Wick’s theorem and Onishi formula

Alternative and consistent approach to off-diagonal operator and norm kernels

  • Regular Article - Theoretical Physics
  • Published:
The European Physical Journal A Aims and scope Submit manuscript

Abstract

The projected generator coordinate method based on the configuration mixing of non-orthogonal Bogoliubov product states, along with more advanced methods based on it, require the computation of off-diagonal Hamiltonian and norm kernels. While the Hamiltonian kernel is efficiently computed via the off-diagonal Wick theorem of Balian and Brezin, the norm kernel relies on the Onishi formula (or equivalently the Pfaffian formula by Robledo or the integral formula by Bally and Duguet). Traditionally, the derivation of these two categories of formulae relies on different formal schemes. In the present work, the formulae for the operator and norm kernels are computed consistently from the same diagrammatic method. The approach further offers the possibility to address kernels involving more general states in the future.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Data availibility statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: This is a theoretical work and neither experimental nor numerical data were generated.]

Notes

  1. The simpler projected generator coordinate method based on non-orthogonal Slater determinants is denoted as the non-orthogonal configuration interaction (NOCI) method in quantum chemistry [1].

  2. Some of the Bogoliubov states mixed in the PGCM ansatz may be either manifestly or accidentally orthogonal. This situation can be dealt with at the price of a generalization of the situation discussed in the present work where all pairs of Bogoliubov states entering Eq. (1) are considered to be non-orthogonal.

  3. The collective coordinate is multi dimensional and contains the variable(s) parameterizing the transformations associated with the symmetry(ies) being restored via projection techniques.

  4. The part of the coefficients fixed by the structure of the symmetry group does not have to be determined variationally.

  5. While a generalization is possible, the reference state \(| \varPhi \rangle \) is supposed to be non-orthogonal to both \(|\varPhi (l)\rangle \) and \(|\varPhi (r)\rangle \).

  6. An asymmetric approach to the norm kernel based on finding a unitary transformation between both states exists [17]. This however differs from the present discussion based on non-unitary Thouless transformations.

  7. Given the chosen Bogoliubov reference state \(| \varPhi \rangle \), it is natural to normal order the operator O with respect to that state as is presently done. While this is already very general, one can easily go one step further and express the operator in normal order with respect to yet another product state, e.g. the particle vacuum. The connection between both situations is straightforward.

  8. Nothing in the proof depends on the character, e.g. rank, of the operators \({{\textbf {R}}}\) and \({{\textbf {L}}}\). Thus, the exact factorization of the norm kernel out of the operator kernel constitutes a general result going beyond the scope of the present study that constraints \({{\textbf {R}}}\) (\({{\textbf {L}}}\)) to be a one-body excitation (de-excitation) operator.

  9. The asymmetric approach to the connected operator kernel detailed in Appendix C, including the natural termination of the exponential at play, can be recovered from the results obtained below by setting \({{\textbf {L}}}\equiv 0\) a posteriori.

  10. Were \({{\textbf {L}}}\) and \({{\textbf {R}}}\) of higher rank, e.g. be two-body operators, this property would be lost. Indeed, an operator, e.g., \({{\textbf {L}}}\) belonging to a loop going through an alternate succession of \({{\textbf {L}}}\) and \({{\textbf {R}}}\) operators connecting two quasi-particle operators of \({{\textbf {O}}}^{ij}\) could further entertain a contraction with an operator \({{\textbf {R}}}\) belonging to another closed loop, thus forming a more elaborate closed string eventually involving more than two quasi-particle operators of \({{\textbf {O}}}^{ij}\).

  11. Given that \({{\textbf {O}}}^{ij}\) is in normal-ordered form, no normal \([\beta \beta ^\dagger ]\) string may occur. This would however be the case if the present discussion were extended to the computation of the connected kernel associated with any product, e.g. \({{\textbf {O}}}^{ij}{{\textbf {O}}}^{kl}\), of normal-ordered operators. This happens for example when considering kernels involving elementary excitations of \(\langle \varPhi (l) | \) and/or \(| \varPhi (r) \rangle \) as in the multi-reference perturbation theory based on a PGCM unperturbed state [15]. Such an extension is straightforward.

  12. If \({{\textbf {L}}}\) and/or \({{\textbf {R}}}\) are of higher rank, i.e. if \(| \varPhi (l) \rangle \) and/or \(| \varPhi (r) \rangle \) do not belong to the manifold of Bogoliubov states, the validity of the off-diagonal Wick’s theorem is lost.

  13. The integers i and j always carry the same parity.

  14. If this rule is not fulfilled, the operator cannot be fully contracted, thus providing a vanishing expectation value by virtue of its normal-ordered form.

  15. Instead of considering all possible strings, this is best seen by keeping the contraction pattern fixed and by exchanging the position of the quasi-particle operators within \({{\textbf {O}}}^{ij}\) in all ways consistent with that contraction pattern and by employing the anti-symmetry of the operator matrix elements to recover the original algebraic contribution.

  16. Useful properties of the double factorial are

    $$\begin{aligned} n!&=n!!(n-1)!!\quad \forall n \end{aligned}$$
    (33a)
    $$\begin{aligned} n!!&=2^kk!\quad \text {for }n=2k\,. \end{aligned}$$
    (33b)
  17. The fourth contraction appearing in Eq. (37d) does not presently occur due to the normal-ordered character of \({{\textbf {O}}}^{ij}\).

  18. The connected character of the presently introduced matrix elements is defined with respect to the two operators \(\beta ^{(\dagger )}_{k_2}\) and \(\beta ^{(\dagger )}_{k_1}\) that translate diagrammatically into two external lines; see Appendix B for details.

  19. The sum over k starts from 0 (1) and runs over even (odd) integers whenever i and j are even (odd).

  20. The trivial diagram obtained for \(n=0\) is \(\langle _0 | \, |^0 \rangle = \langle \varPhi | \varPhi \rangle =1\). Since it contains neither vertices nor lines, it does not qualify as a closed diagram.

  21. In the present discussion \(\varGamma _{cl}(n_i)\) equally represents the closed diagram and its algebraic contribution.

  22. Strickly speaking, the Onishi formula, as any formula based on the symmetric approach, can only deliver the phase of the overlap \(\langle \varPhi (l) | \varPhi (r) \rangle \) modulo the knowledge of the phase associated with, i.e. initially fixed for, the overlaps \(\langle \varPhi (l) |\varPhi \rangle \) and \(\langle \varPhi |\varPhi (r) \rangle \) of the two involved states with respect to the reference Bogoliubov state \(|\varPhi \rangle \).

  23. The diagrammatic rules can be straightforwardly generalized to operators \({{\textbf {R}}}\) and \({{\textbf {L}}}\) of higher, and possibly different, ranks.

  24. Note that \(n_o>2\) operators \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) cannot be equivalent if any of them is connected to an operator \({{\textbf {R}}}\) (\({{\textbf {L}}}\)) given that the latter cannot entertain the same contraction pattern with \(n_o>2\) operators \({{\textbf {L}}}\) (\({{\textbf {R}}}\)). More general patterns would however occur if \({{\textbf {L}}}\) and \({{\textbf {R}}}\) were of higher ranks.

  25. There is no general rule to identify them such that the symmetry factor associated with each topologically distinct unlabelled diagram must be identified on a case by case basis.

References

  1. A.J.W. Thom, M. Head-Gordon, J. Chem. Phys. 131, 124113 (2009)

    Article  Google Scholar 

  2. P. Ring, P. Schuck, The Nuclear Many-Body Problem (Springer, New York, 1980)

    Book  Google Scholar 

  3. M. Bender, P.-H. Heenen, P.-G. Reinhard, Self-consistent mean-field models for nuclear structure. Rev. Mod. Phys. 75, 121 (2003). https://doi.org/10.1103/RevModPhys.75.121

    Article  Google Scholar 

  4. T. Niksic, D. Vretenar, P. Ring, Relativistic nuclear energy density functionals: mean-field and beyond. Prog. Part. Nucl. Phys. 66, 519–548 (2011). arXiv:1102.4193, https://doi.org/10.1016/j.ppnp.2011.01.055

  5. L.M. Robledo, T.R. Rodríguez, R.R. Rodríguez-Guzmán, Mean field and beyond description of nuclear structure with the Gogny force: a review. J. Phys. G 46(1), 013001 (2019). arXiv:1807.02518, https://doi.org/10.1088/1361-6471/aadebd

  6. Z.-C. Gao, M. Horoi, Y.S. Chen, Variation after projection with a triaxially deformed nuclear mean field. Phys. Rev. C 92(6), 064310 (2015). arXiv:1509.03058, https://doi.org/10.1103/PhysRevC.92.064310

  7. C.F. Jiao, J. Engel, J.D. Holt, Neutrinoless double-beta decay matrix elements in large shell-model spaces with the generator-coordinate method. Phys. Rev. C 96(5), 054310 (2017). arXiv:1707.03940, https://doi.org/10.1103/PhysRevC.96.054310

  8. B. Bally, A. Sánchez-Fernández, T.R. Rodríguez, Variational approximations to exact solutions in shell-model valence spaces: calcium isotopes in the pf-shell. Phys. Rev. C 100(4), 044308 (2019). arXiv:1907.05493, https://doi.org/10.1103/PhysRevC.100.044308

  9. N. Shimizu, T. Mizusaki, K. Kaneko, Y. Tsunoda, Generator-coordinate methods with symmetry-restored Hartree–Fock–Bogoliubov wave functions for large-scale shell-model calculations. Phys. Rev. C 103(6), 064302 (2021). https://doi.org/10.1103/PhysRevC.103.064302

    Article  Google Scholar 

  10. A. Sánchez-Fernández, B. Bally, T.R. Rodríguez, Variational approximations to exact solutions in shell-model valence spaces: systematic calculations in the \(sd\)-shell. arXiv:2106.08841

  11. J.M. Yao, B. Bally, J. Engel, R. Wirth, T.R. Rodríguez, H. Hergert, \(Ab\)\(Initio\) treatment of collective correlations and the neutrinoless double beta decay of \(^{48}\)Ca. Phys. Rev. Lett. 124(23), 232501 (2020). arXiv:1908.05424, https://doi.org/10.1103/PhysRevLett.124.232501

  12. J.M. Yao, J. Engel, L. J. Wang, C.F. Jiao, H. Hergert, Generator-coordinate reference states for spectra and \(0\nu \beta \beta \) decay in the in-medium similarity renormalization group. Phys. Rev. C 98(5), 054311 (2018). arXiv:1807.11053, https://doi.org/10.1103/PhysRevC.98.054311

  13. M. Frosini, T. Duguet, J.-P. Ebran, B. Bally, H. Hergert, T.R. Rodríguez, R. Roth, J. Yao, V. Somà, Multi-reference many-body perturbation theory for nuclei: III. Ab initio calculations at second order in PGCM-PT. Eur. Phys. J. A 58(4), 64 (2022). arXiv:2111.01461, https://doi.org/10.1140/epja/s10050-022-00694-x

  14. M. Frosini, T. Duguet, J.-P. Ebran, B. Bally, T. Mongelli, T.R. Rodríguez, R. Roth, V. Somà, Multi-reference many-body perturbation theory for nuclei: II. Ab initio study of neon isotopes via PGCM and IM-NCSM calculations. Eur. Phys. J. A 58(4), 63 (2022). arXiv:2111.00797, https://doi.org/10.1140/epja/s10050-022-00693-y

  15. M. Frosini, T. Duguet, J.-P. Ebran, V. Somà, Multi-reference many-body perturbation theory for nuclei: I. Novel PGCM-PT formalism. Eur. Phys. J. A 58(4), 62 (2022). arXiv:2110.15737, https://doi.org/10.1140/epja/s10050-022-00692-z

  16. T. Duguet, A. Signoracci, Symmetry broken and restored coupled-cluster theory. II. Global gauge symmetry and particle number. J. Phys. G 44(1), 015103 (2017). [Erratum: J.Phys.G 44, 049601 (2017)]. arXiv:1512.02878, https://doi.org/10.1088/0954-3899/44/1/015103

  17. B. Bally, T. Duguet, Norm overlap between many-body states: uncorrelated overlap between arbitrary Bogoliubov product states. Phys. Rev. C 97(2), 024304 (2018). arXiv:1704.05324, https://doi.org/10.1103/PhysRevC.97.024304

  18. R. Balian, E. Brezin, Nonunitary Bogoliubov transformations and extension of Wick’s theorem. Il Nuovo Cimento B 64(1), 37–55 (1969)

    Article  MathSciNet  Google Scholar 

  19. N. Onishi, S. Yoshida, Nucl. Phys. 80, 367 (1966)

    Article  Google Scholar 

  20. L.M. Robledo, The sign of the overlap of HFB wave functions. Phys. Rev. C 79, 021302 (2009)

    Article  Google Scholar 

  21. T. Mizusaki, M. Oi, A new formulation to calculate general HFB matrix elements through Pfaffian. Phys. Lett. B 715, 219–224 (2012). arXiv:1204.6531, https://doi.org/10.1016/j.physletb.2012.07.023

  22. T. Mizusaki, M. Oi, F.-Q. Chen, Y. Sun, Grassmann integral and Balian-Brézin decomposition in Hartree–Fock–Bogoliubov matrix elements. Phys. Lett. B 725, 175–179 (2013). arXiv:1305.1682, https://doi.org/10.1016/j.physletb.2013.07.005

  23. D.J. Thouless, Perturbation theory in statistical mechanics and the theory of superconductivity. Ann. Phys. 10, 553 (1960)

    Article  MathSciNet  MATH  Google Scholar 

  24. R.J. Bartlett, J. Noga, Chem. Phys. Lett. 150, 29 (1988)

    Article  Google Scholar 

  25. J.B. Robinson, P.J. Knowles, J. Chem. Phys. 136, 054114 (2012)

    Article  Google Scholar 

  26. A. Marie, F. Kossoski, P.-F. Loos, J. Chem. Phys. 155, 104105 (2021)

    Article  Google Scholar 

  27. A.L. Fetter, J.D. Walecka, Quantum Theory of Many-Particle Systems (Dover Books on Physics), Dover Publications (2003).

  28. J. Blaizot, G. Ripka, Quantum Theory of Finite Systems (MIT Press, Cambridge, 1986)

    Google Scholar 

  29. T. Mizusaki, M. Oi, N. Shimizu, Why does the sign problem occur in evaluating the overlap of HFB wave functions? Phys. Lett. B 779, 237–243 (2018). arXiv:1711.00369, https://doi.org/10.1016/j.physletb.2018.02.012

  30. K. Neergard, E. Wust, Nucl. Phys. A 402, 311 (1983)

    Article  Google Scholar 

  31. M. Oi, T. Mizusaki, N. Shimizu, Y. Sun, Hidden symmetries in the HFB norm overlap functions. EPJ Web Conf. 223, 01047 (2019). https://doi.org/10.1051/epjconf/201922301047

    Article  Google Scholar 

  32. M. Krivoruchenko, Trace identities for skew-symmetric matrices. Math. Comput. Sci. 1(2), 21 (2016)

    Google Scholar 

  33. E.R. Caianiello, Combinatorics and Renormalization in Quantum Field Theory (WA Benjamin Inc, Reading, 1973)

  34. G.C. Wick, The evaluation of the collision matrix. Phys. Rev. 80, 268 (1950). https://doi.org/10.1103/PhysRev.80.268

    Article  MathSciNet  MATH  Google Scholar 

  35. A. Tichai, P. Arthuis, T. Duguet, H. Hergert, V. Somá, R. Roth, Bogoliubov many-body perturbation theory for open-shell nuclei. Phys. Lett. B 786, 195–200 (2018). arXiv:1806.10931, https://doi.org/10.1016/j.physletb.2018.09.044

  36. P. Arthuis, T. Duguet, A. Tichai, R.D. Lasseri, J.P. Ebran, ADG: automated generation and evaluation of many-body diagrams I. Bogoliubov many-body perturbation theory. Comput. Phys. Commun. 240, 202–227 (2019). arXiv:1809.01187, https://doi.org/10.1016/j.cpc.2018.11.023

  37. J. Ripoche, A. Tichai, T. Duguet, Normal-ordered k-body approximation in particle-number-breaking theories. Eur. Phys. J. A 56(2). https://doi.org/10.1140/epja/s10050-020-00045-8. http://dx.doi.org/10.1140/epja/s10050-020-00045-8

Download references

Acknowledgements

The authors wish to thank V. Somà and P. Arthuis for proofreading the manuscript. A.P. is supported by the CEA NUMERICS program, which has received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Skłodowska-Curie grant agreement No 800945.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. Porro.

Additional information

Communicated by Michael Bender.

Appendices

Appendix A: Normal-ordered operator

An arbitrary rank-N particle-number-conserving fermionic operator \(O\) can be written as

$$\begin{aligned} O\equiv \sum _{n=0}^N {O^{nn}} , \end{aligned}$$
(A. 1)

where

$$\begin{aligned} O^{mn}\equiv \frac{1}{m!} \frac{1}{n!} \sum _{\begin{array}{c} a_1\cdots a_m\\ b_1\cdots b_n \end{array}} o^{a_1\cdots a_m}_{b_1\cdots b_n} \, c^\dag _{a_1}\cdots c^\dag _{a_m}c_{b_n}\cdots c_{b_1} , \end{aligned}$$
(A. 2)

contains m(n) particle creation (annihilation) operators. The zero-body part \(O^{00}\) is the scalar obtained as the expectation value of O in the particle vacuum

$$\begin{aligned} O^{00}=\langle 0|O|0 \rangle . \end{aligned}$$
(A. 3)

In Eq. (A. 2), matrix elements are fully anti-symmetric under the exchange of any pair of upper or lower indices, i.e.

$$\begin{aligned} o^{a_1\cdots a_m}_{b_1\cdots b_n} = \epsilon (\sigma _u) \epsilon (\sigma _l) \, o^{\sigma _u(a_1\cdots a_m)}_{\sigma _l(b_1\cdots b_n)} , \end{aligned}$$
(A. 4)

where \(\epsilon (\sigma _u)\) (\(\epsilon (\sigma _l)\)) refers to the signature of the permutation \(\sigma _u(\cdots )\) (\(\sigma _l(\cdots )\)) of the m (n) upper (lower) indices. In case the particle-number conserving operator is hermitian, each term \(O^{nn}\) is hermitian with its matrix elements fulfilling

$$\begin{aligned} o^{a_1\cdots a_n}_{b_1\cdots b_n} = \left( o_{a_1\cdots a_m}^{b_1\cdots b_n}\right) ^*. \end{aligned}$$
(A. 5)

By virtue of standard Wick’s theorem [34], the operator O can be normal ordered with respect to the Bogoliubov vacuum \(| \varPhi \rangle \)

$$\begin{aligned} O&\equiv \sum _{n=0}^N{{\textbf {O}}}^{[2n]} \equiv \sum _{n=0}^N\sum _{\begin{array}{c} i,j=0\\ i+j=2n \end{array}}^{2n}{{\textbf {O}}}^{ij} \end{aligned}$$
(A. 6)

where the component

$$\begin{aligned} {{\textbf {O}}}^{ij}\equiv \frac{1}{i!j!}\sum _{\begin{array}{c} k_1\cdots k_i\\ l_1\cdots l_j \end{array}}{{\textbf {o}}}^{k_1\cdots k_i}_{l_1\cdots l_j}\beta ^\dagger _{k_1}\cdots \beta ^\dagger _{k_i}\beta _{l_j}\cdots \beta _{l_1} \end{aligned}$$
(A. 7)

contains i(j) quasi-particle creation (annihilation) operators. The zero-body part \({{\textbf {O}}}^{[0]}\) is the scalar obtained as the expectation value of O in the Bogoliubov vacuum

$$\begin{aligned} {{\textbf {O}}}^{00}= \langle \varPhi |O|\varPhi \rangle \, . \end{aligned}$$
(A. 8)

In Eq. (A. 7), matrix elements are fully anti-symmetric under the exchange of any pair of upper or lower indices, i.e.

$$\begin{aligned} {{\mathbf {o}}}^{k_1\cdots k_i}_{l_1\cdots l_j} = \epsilon (\sigma _u) \epsilon (\sigma _l) \, {{\mathbf {o}}}^{\sigma _u(k_1\cdots k_i)}_{\sigma _l(l_1\cdots l_j)} \, . \end{aligned}$$
(A. 9)

These matrix elements are functionals of the Bogoliubov matrices \(({\mathcal {U}},{\mathcal {V}})\) associated with \(| \varPhi \rangle \) and of the matrix elements \(\{o^{a_1\cdots a_n}_{b_1\cdots b_n}\}\) initially defining the operator O. As such, the content of each operator \({{\mathbf {O}}}^{ij}\) depends on the rank N of O. For more details about the normal ordering procedure and for explicit expressions of the matrix elements up to \(N=3\), see Refs. [16, 35,36,37].

In case the operator is hermitian, each component \({{\textbf {O}}}^{[2n]}\) is itself hermitian with \({{\textbf {O}}}^{ij} = {{\textbf {O}}}^{ji\dagger }\) such that matrix elements satisfy

$$\begin{aligned} {{\mathbf {o}}}^{k_1\cdots k_i}_{l_1\cdots l_j} = \left( {\mathbf{o}}^{l_1\cdots l_j}_{k_1\cdots k_i}\right) ^*\, . \end{aligned}$$
(A. 10)

Appendix B:Diagrammatic rules

The present appendix is dedicated to setting up the diagrammatic rules allowing one to compute matrix elements of the form

$$\begin{aligned}&\langle _s|^i \, _j| ^t \rangle \equiv \frac{1}{s!} \frac{1}{t!} \langle \varPhi | {{\textbf {L}}}^s {{\textbf {O}}}^{ij} {{\textbf {R}}}^t |\varPhi \rangle \, , \end{aligned}$$
(B.11a)
$$\begin{aligned}&\langle _s| \,|^t \rangle \equiv \frac{1}{s!} \frac{1}{t!} \langle \varPhi | {{\textbf {L}}}^s {{\textbf {R}}}^t |\varPhi \rangle \, , \end{aligned}$$
(B.11b)

where \({{\textbf {O}}}^{ij}\) takes the form given in Eq. (A. 7) and where \({{\textbf {R}}}\) (\({{\textbf {L}}}\)) is a one-bodyFootnote 23 excitation (de-excitation) operator as defined in Eqs. (26a, 26b)).

The diagrammatic rules are also worked out to compute a second category of matrix elements of present interest

$$\begin{aligned} \langle _s | ^{k_2} | \, |_{k_1} |^t \rangle&\equiv \frac{1}{s!} \frac{1}{t!} \langle \varPhi | {{\textbf {L}}}^s \beta ^{\dagger }_{k_2} \beta _{k_1} {{\textbf {R}}}^t |\varPhi \rangle \, , \end{aligned}$$
(B.12a)
$$\begin{aligned} \langle _s | _{k_2} | \, |_{k_1} |^t \rangle&\equiv \frac{1}{s!} \frac{1}{t!} \langle \varPhi | {{\textbf {L}}}^s \beta _{k_2} \beta _{k_1} {{\textbf {R}}}^t |\varPhi \rangle \, , \end{aligned}$$
(B.12b)
$$\begin{aligned} \langle _s | ^{k_2} | \, |^{k_1} |^t \rangle&\equiv \frac{1}{s!} \frac{1}{t!} \langle \varPhi | {{\textbf {L}}}^s \beta ^{\dagger }_{k_2} \beta ^{\dagger }_{k_1} {{\textbf {R}}}^t |\varPhi \rangle \, , \end{aligned}$$
(B.12c)
$$\begin{aligned} \langle _s | _{k_2} | \, |^{k_1} |^t \rangle&\equiv \frac{1}{s!} \frac{1}{t!} \langle \varPhi | {{\textbf {L}}}^s \beta _{k_2} \beta ^{\dagger }_{k_1} {{\textbf {R}}}^t |\varPhi \rangle \, . \end{aligned}$$
(B.12d)

These matrix elements differ from the two introduced in Eq. (B.11) by the presence of two “external/fixed” quasi-particle operators, i.e. quasi-particle operators whose indices are not summed over.

The diagrammatic rules derive from the straight application of standard Wick’s theorem with respect to the Bogoliubov vacuum \(|\varPhi \rangle \). The application of Wick’s theorem delivers the complete set of fully contracted terms associated with the operator product entering the matrix element of interest. Given that the operators at play are all conveniently expressed in the quasi-particle basis associated with the Bogoliubov vacuum \(| \varPhi \rangle \), the four possible elementary contractions take the simplest possible form

$$\begin{aligned} {\mathbf {R}}_{k_1k_2}&= \begin{pmatrix} \frac{\langle \varPhi |\, \beta ^{\dagger }_{k_2} \beta _{k_1} | \varPhi \rangle }{\langle \varPhi | \varPhi \rangle } &{} \quad \frac{\langle \varPhi |\, \beta _{k_2} \beta _{k_1} | \varPhi \rangle }{\langle \varPhi | \varPhi \rangle } \\ \frac{\langle \varPhi |\, \beta ^{\dagger }_{k_2} \beta ^{\dagger }_{k_1} | \varPhi \rangle }{\langle \varPhi | \varPhi \rangle } &{}\quad \frac{\langle \varPhi |\, \beta _{k_2} \beta ^{\dagger }_{k_1} | \varPhi \rangle }{\langle \varPhi | \varPhi \rangle } \end{pmatrix} \nonumber \\&\equiv \begin{pmatrix} R^{+-}_{k_1k_2} &{} \quad R^{--}_{k_1k_2} \\ R^{++}_{k_1k_2} &{} \quad R^{-+}_{k_1k_2} \end{pmatrix} \nonumber \\&= \begin{pmatrix} 0 &{}\quad 0 \\ 0 &{}\quad \delta _{k_1k_2} \end{pmatrix} \ , \end{aligned}$$
(B.13)

such that the sole non-zero contraction \(R^{-+}_{k_1k_2} = \delta _{k_1k_2}\) needs to be considered.

1.1 Appendix B.1: Diagrammatic representation

Fig. 6
figure 6

Diagrammatic representation of the normal-ordered operator \({{\textbf {O}}}^{ij}\) as a fully anti-symmetric Hugenholtz vertex

Fig. 7
figure 7

Diagrammatic representation of the two Thouless operators \({{\textbf {L}}}\) (left) and \({{\textbf {R}}}\) (right)

The diagrammatic representation of the various contributions to the matrix elements of interest relies on the definition of the following building blocks

  1. 1.

    As illustrated in Fig. 6, the normal-ordered operator \({{\textbf {O}}}^{ij}\) entering the matrix element \(\langle _s|^i \, _j| ^t \rangle \) is represented by a Hugenholtz vertex with i (j) lines traveling out of (into) it and representing quasi-particle creation (annihilation) operators. The algebraic factor \({{\textbf {o}}}^{k_1 \cdots k_i}_{l_1 \cdots l_j}\) is associated to the vertex while assigning indices \(k_1 \cdots k_i\) consecutively from the leftmost to the rightmost line above the vertex and indices \(l_{1} \cdots l_{j}\) consecutively from the leftmost to the rightmost line below the vertex.

  2. 2.

    As illustrated in Fig. 7, the operator \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) entering all matrix elements of present interest is represented by a vertex of the \({{\textbf {O}}}^{02}\) (\({{\textbf {O}}}^{20}\)) type and carry the associated algebraic factor \({{\textbf {z}}}^{*}_{l_1 l_2}(l)\) (\({{\textbf {z}}}^{k_1 k_2}(r)\)).

  3. 3.

    The only non-zero contraction \(R^{-+}_{k_1k_2} = \delta _{k_1k_2}\) is represented in Fig. 8 and connects two up-going lines associated with one annihilation and one creation operator, both carrying the same quasi-particle index. For simplicity, one can eventually represent the contraction as a line carrying a single up-going arrow along with one quasi-particle index.

Fig. 8
figure 8

Diagrammatic representation of the single non-zero elementary contraction. The convention is that the left-to-right reading of a matrix element corresponds to the up-down reading of the diagram

1.2 Appendix B.2: Diagrams generation

With these building blocks at hand, one can construct the diagrams gathering all contributions to the matrix elements introduced in Eq. (B.11) and (B.12). The basic rules to do so are as follows

  1. 1.

    Diagrams contain s square vertices (\({{\textbf {L}}}\)) and t triangle vertices (\({{\textbf {R}}}\)), the former being located above the latter. This is consistent with the convention that the left-to-right reading of a matrix element corresponds to the up-down reading of the diagram.

  2. 2.

    Diagrams making up the two matrix elements introduced in Eq. (B.11) are vacuum-to-vacuum diagrams with no line leaving the diagram. In \(\langle _s|^i \, _j| ^t \rangle \), a dot vertex (\({{\textbf {O}}}^{ij}\)) is located in between the square and triangle vertices. This is consistent with the convention that the left-to-right reading of a matrix element corresponds to the up-down reading of the diagram.

  3. 3.

    Diagrams making up the four matrix elements introduced in Eq. (B.12) are linked with two external lines associated with the operators \(\beta ^{(\dagger )}_{k_2}\) and \(\beta ^{(\dagger )}_{k_1}\). The two lines leave the diagram on the same side to the, e.g., left such that (i) both lines are asymptotically in between the square (\({{\textbf {L}}}\)) and triangle (\({{\textbf {R}}}\)) vertices and such that (ii) the line carrying index \(k_2\) is asymptotically located above the line carrying index \(k_1\). This is consistent with the convention that the left-to-right reading of a matrix element corresponds to the up-down reading of the diagram. The arrow carried by each of the two lines points towards the interior (exterior) of the diagram if it is associated with a quasi-particle creation (annihilation) operator.

  4. 4.

    The fact that \(R^{-+}_{k_1k_2}\) is the sole non-zero contraction implies that the number of quasi-particle creation operators involved in a given matrix element is equal to the number of quasi-particle annihilation operators. Given that each operator \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) contains two quasi-particle annihilation (creation) operators, this property require the following conditions to be fulfilled

    1. (a)

      \(\langle _s|^i \, _j| ^t \rangle \) demands \(t = s + (i-j)/2\)  ,

    2. (b)

      \(\langle _s| \, | ^t \rangle \) demands \(t = s\)   ,

    3. (c)

      \(\langle _s | ^{k_2} | \, |_{k_1} |^t \rangle \) demands \(t=s\)   ,

    4. (d)

      \(\langle _s | _{k_2} | \, |_{k_1} |^t \rangle \) demands \(t=s+1\)   ,

    5. (e)

      \(\langle _s | ^{k_2} | \, |^{k_1} |^t \rangle \) demands \(t=s-1\)   ,

    6. (f)

      \(\langle _s | _{k_2} | \, |^{k_1} |^t \rangle \) demands \(t=s\)   .

    such that \(t\ge (i-j)/2\), \(t\ge 1\) and \(s\ge 1\) in case (a), (d) and (e), respectively.

  5. 5.

    Given the above considerations, one must construct all possible topologically distinct unlabelled diagrams from the building blocks; i.e., contract together the lines belonging to the s square (\({{\textbf {L}}}\)) vertices and to the t triangle (\({{\textbf {R}}}\)) vertices, along with those belonging to the dot (\({{\textbf {O}}}^{ij}\)) vertex or to the two external (\(\beta ^{(\dagger )}_{k_2}\) and \(\beta ^{(\dagger )}_{k_1}\)) lines whenever applicable, in all possible ways. Unlabelled diagrams correspond to diagrams in which \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) vertices are not distinguished by a label. Topologically distinct unlabelled diagrams cannot be obtained from one another via a mere displacement, i.e. translation, of the vertices in the plane of the drawing.

  6. 6.

    The above process is constrained by the following properties

    1. (a)

      Because the operators \({{\textbf {L}}}\), \({{\textbf {R}}}\), and \({{\textbf {O}}}^{ij}\) are in normal-ordered form with respect to \(| \varPhi \rangle \), self-contractions must be ignored.

    2. (b)

      Because \(R^{-+}_{k_1k_2}\) is the sole non-zero contraction, \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) operators cannot display contractions among themselves, i.e. they necessarily contract with \({{\textbf {R}}}\) (\({{\textbf {L}}}\)) operators, along with the i (j) quasi-particle creation (annihilation) operators inside \({{\textbf {O}}}^{ij}\) or with \(\beta ^{\dagger }_{k_2}\) (\(\beta _{k_2}\)) and/or \(\beta ^{\dagger }_{k_1}\) (\(\beta _{k_1}\)) whenever applicable.

  7. 7.

    The diagrams making up the various matrix elements of interest display different typical topologies. Indeed, each contribution generated via the application of Wick’s theorem can be expressed as a product of strings of contractions, each of which involves a subset of the \({{\textbf {L}}}\) and \({{\textbf {R}}}\) operators at play. As for \(\langle _s|^i \, _j| ^t \rangle \) with \((i+j) \ge 4\), several such strings actually involve quasi-particle operators belonging to \({{\textbf {O}}}^{ij}\), thus forming an overall closed string that is said to be connected to \({{\textbf {O}}}^{ij}\). Translated into diagrammatic language, closed strings correspond to topologically disjoint closed sub-diagrams. In the case of \(\langle _s|^i \, _j| ^t \rangle \), any given diagram is thus made out of disjoint closed sub-diagrams, one of which is connected. As for the matrix elements introduced in Eq. (B.12), one set of contractions must form a connected string involving the operators \(\beta ^{(\dagger )}_{k_2}\) and \(\beta ^{(\dagger )}_{k_1}\). These two operators cannot belong to two disjoint strings given that any string necessarily involves an even number of quasi-particle operators. Eventually, the diagrams making up the matrix elements introduced in Eq. (B.12) are made out of disjoint closed sub-diagrams, one of which is connected to the external lines. Last but not least, each diagram contributing to \(\langle _s| \, | ^t \rangle \) is made out of disjoint closed sub-diagrams, none of which is connected.

1.3 Appendix B.3: Diagrams evaluation

Once all the diagrams are drawn, one must compute their expressions. The rules to do so are the following

  1. 1.

    Label all quasi-particle lines and associate the appropriate factor to each vertex, i.e. a factor \({{\textbf {z}}}^*_{l_1l_2}(l)\) (\({{\textbf {z}}}^{k_1k_2}(r)\)) to each vertex \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) and a factor \({{\textbf {o}}}^{k_1 \cdots k_i}_{l_1 \cdots l_j}\) to the vertex \({{\textbf {O}}}^{ij}\), respectively.

  2. 2.

    Sum over all internal line labels.

  3. 3.

    Include a factor \((n_e!)^{-1}\) for each set of \(n_e\) equivalent internal lines. Equivalent internal lines are those connecting identical vertices.

  4. 4.

    For any topologically distinct unlabelled diagram \(\varGamma \), a symmetry factor \(S_{\varGamma }^{-1}\) must be considered. Given a labelled version of \(\varGamma \), i.e. a version in which each operator \({{\textbf {L}}}\) and \({{\textbf {R}}}\) carries a specific label, \(S_{\varGamma }\) is equal to the number of permutations of the \({{\textbf {L}}}\) and \({{\textbf {R}}}\) operators delivering a topologically equivalent labelled diagram. The most obvious cases correspond to equivalent subgroups of \({{\textbf {L}}}\) and \({{\textbf {R}}}\) operators whose overall permutations lead to topologically equivalent labelled diagrams. The simplest example concerns two \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) operators that are doubly connected to \({{\textbf {O}}}^{ij}\) or singly connected to \({{\textbf {O}}}^{ij}\) and to the same operator \({{\textbf {R}}}\) (\({{\textbf {L}}}\)). These \(n_o\equiv 2\) operators \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) are equivalent and contribute a factor 2! to \(S_{\varGamma }\). The next simplest example corresponds to \(n_o>2\) operators \({{\textbf {L}}}\) (\({{\textbf {R}}}\)) fully connected to \({{\textbf {O}}}^{ij}\). These \(n_o>2\) operators are indeed equivalentFootnote 24 such that their permutations contribute a factor \(n_o!\) to \(S_{\varGamma }\). Beyond those two examples, a set of \({{\textbf {L}}}\) and \({{\textbf {R}}}\) operators can form a string of contractions that is equivalent to other identical strings. Such \(n_s\) strings are said to be equivalent and contribute a factor \(n_s!\) to \(S_{\varGamma }\). Eventually, less obvious permutations can deliver topologically equivalent labelled diagrams and thus contribute to \(S_{\varGamma }\)Footnote 25.

  5. 5.

    Provide the diagram with a sign \((-1)^{\ell _c}\), where \(\ell _c\) is the number of line crossings in the diagram. For diagrams containing external lines, their potential crossing must be counted.

Appendix C: Asymmetric approach

The asymmetric approach constitutes the standard path to the off-diagonal Wick theorem at play in the computation of the connected operator kernel [2, 18]. Employing the simplified notation

$$\begin{aligned} {{\textbf {R}}}&\equiv {{\textbf {Z}}}^{20}(l,r) =\frac{1}{2}\sum _{k_1k_2}{{\textbf {z}}}^{k_1k_2}(l,r)\beta ^\dagger _{k_1}(l)\beta ^\dagger _{k_2}(l)\,, \end{aligned}$$
(C.14)

for the Thouless operator introduced in Eq. (18b) and satisfying \(\langle \varPhi (l) | {{\textbf {R}}} = 0\), the connected operator kernel reads as

$$\begin{aligned} \frac{\langle \varPhi (l) | O | \varPhi (r) \rangle }{\langle \varPhi (l) | \varPhi (r) \rangle }&\equiv \sum _{n=0}^N\sum _{\begin{array}{c} i,j=0\\ i+j=2n \end{array}}^{2n} \langle \varPhi (l) |{{\textbf {O}}}^{ij} e^{{{\textbf {R}}}} | \varPhi (l) \rangle \, , \nonumber \\&= \sum _{n=0}^N\sum _{\begin{array}{c} i,j=0\\ i+j=2n \end{array}}^{2n} \langle \varPhi (l) | ^{R}{{{\textbf {O}}}^{ij}} | \varPhi (l) \rangle \,, \end{aligned}$$
(C.15)

where the operator

$$\begin{aligned} ^{R}{{{\textbf {O}}}^{ij}}&\equiv e^{-{{\textbf {R}}}} {{\textbf {O}}}^{ij} e^{{{\textbf {R}}}} \nonumber \\&= \frac{1}{i!j!}\sum _{\begin{array}{c} k_1 \cdots k_i\\ l_1 \cdots l_j \end{array}} {{\textbf {o}}}^{k_1\cdots k_i}_{l_1\cdots l_j} \, ^R\beta ^{\dagger }_{k_1} \cdots ^R\beta ^{\dagger }_{k_i} \, ^R\beta _{l_j} \cdots ^R\beta _{l_1} \, , \end{aligned}$$
(C.16)

formally reads as \({{\textbf {O}}}^{ij}\) but with the quasi-particle operators replaced by their similarity-transformed partners

$$\begin{aligned} \begin{pmatrix} ^R\beta _k\\ ^R\beta ^{\dagger }_k \end{pmatrix}&\equiv e^{-{{\textbf {R}}}}\begin{pmatrix} \beta _k\\ \beta ^\dagger _k \end{pmatrix} e^{{{\textbf {R}}}} \, . \end{aligned}$$
(C.17)

Given that the similarity-transformed quasi-particle operators satisfy anticommutation relations

$$\begin{aligned} \{^R\beta ^{\dagger }_{k_1},^R\beta ^{\dagger }_{k_2}\}&= e^{-{{\textbf {R}}}}\{\beta ^{\dagger }_{k_1},\beta ^{\dagger }_{k_2}\}e^{{{\textbf {R}}}} = 0\, , \end{aligned}$$
(C.18a)
$$\begin{aligned} \{^R\beta _{k_1},^R\beta _{k_2}\}&= e^{-{{\textbf {R}}}}\{\beta _{k_1},\beta _{k_2}\}e^{{{\textbf {R}}}} = 0 \, , \end{aligned}$$
(C.18b)
$$\begin{aligned} \{^R\beta _{k_1},^R\beta ^{\dagger }_{k_2}\}&= e^{-{{\textbf {R}}}}\{\beta _{k_1},\beta ^{\dagger }_{k_2}\}e^{{{\textbf {R}}}} = \delta _{k_1k_2}\, , \end{aligned}$$
(C.18c)

standard Wick’s theorem with respect to \(| \varPhi (l) \rangle \) applies and can be used to compute the matrix elements entering the right-hand side of Eq. (C.15). This results into the standard set of fully contracted terms, except that the elementary contractions at play do not involve the original quasi-particle operators but rather the similarity-transformed ones. The latter are related to the former via a non-unitary Bogoliubov transformation that is now detailed to compute the relevant elementary contractions.

Using Baker-Campbell-Hausdorff (BCH) identity, one first evaluates Eq. (C.17) according to

$$\begin{aligned} ^R\beta ^{(\dagger )}_{k_1}&= \beta ^{(\dagger )}_{k_1} - [{{\textbf {R}}},\beta ^{(\dagger )}_{k_1}] + \frac{1}{2!} [{{\textbf {R}}},[{{\textbf {R}}},\beta ^{(\dagger )}_{k_1}]] + \cdots \, . \end{aligned}$$
(C.19)

Given the two elementary commutators

$$\begin{aligned} \Big [ \beta ^{\dagger }_{k}(l) \beta ^{\dagger }_{k'}(l) , \beta _{k_1}(l) \Big ]&= \beta ^{\dagger }_{k}(l) \delta _{k' k_1} - \beta ^{\dagger }_{k'}(l) \delta _{k k_1} \, , \end{aligned}$$
(C.20a)
$$\begin{aligned} \Big [ \beta ^{\dagger }_{k}(l) \beta ^{\dagger }_{k'}(l) , \beta ^{\dagger }_{k_1} (l)\Big ]&= 0 \,, \end{aligned}$$
(C.20b)

it is straightforward to prove

$$\begin{aligned} \Big [ {{\textbf {R}}} , \beta _{k_1} \Big ]&= -\sum _{k_2} \Big [U(l) {{\textbf {z}}}(l,r)\Big ]_{k_1k_2} \beta ^{\dagger }_{k_2} (l)\, , \\ \Big [ {{\textbf {R}}}, \Big [ {{\textbf {R}}} , \beta _{k_1} \Big ] \Big ]&= 0 \, , \\ \Big [ {{\textbf {R}}}, \ldots \Big [ {{\textbf {R}}} , \beta _{k_1} \Big ] \ldots \Big ]&= 0 \, , \\ \Big [ {{\textbf {R}}} , \beta ^{\dagger }_{k_1} \Big ]&= -\sum _{k_2} \Big [V(l) {{\textbf {z}}}(l,r)\Big ]_{k_1k_2} \beta ^{\dagger }_{k_2} (l) \, , \\ \Big [ {{\textbf {R}}}, \Big [ {{\textbf {R}}} , \beta ^{\dagger }_{k_1} \Big ] \Big ]&= 0 \, , \\ \Big [ {{\textbf {R}}}, \ldots \Big [ {{\textbf {R}}} , \beta ^{\dagger }_{k_1} \Big ] \ldots \Big ]&= 0 \, , \end{aligned}$$

such that

$$\begin{aligned} ^R\beta _{k_1} =&\sum _{k_2} U_{k_1k_2}(l) \beta _{k_2} (l) \nonumber \\&+ \Big [V^{*}(l) + U(l) {{\textbf {z}}}(l,r)\Big ]_{k_1k_2} \beta ^{\dagger }_{k_2} (l) \, , \end{aligned}$$
(C.21a)
$$\begin{aligned} ^R\beta ^{\dagger }_{k_1} =&\sum _{k_2} V_{k_1k_2}(l) \beta _{k_2} (l) \nonumber \\&+ \Big [U^{*}(l) + V(l) {{\textbf {z}}}(l,r)\Big ]_{k_1k_2} \beta ^{\dagger }_{k_2} (l) \, , \end{aligned}$$
(C.21b)

which can be compacted in matrix form according to the non-unitary Bogoliubov transformation

$$\begin{aligned} \begin{pmatrix} ^R\beta \\ ^R\beta ^{\dagger } \end{pmatrix}&= \begin{pmatrix} U(l) &{} V^{*}(l) + U(l) {{\textbf {z}}}(l,r) \\ V(l) &{} U^{*}(l) + V(l) {{\textbf {z}}}(l,r) \end{pmatrix} \begin{pmatrix} \beta (l)\\ \beta ^\dagger (l) \end{pmatrix} \, . \end{aligned}$$
(C.22)

As Eqs. (C.21C.22) testify, the infinite expansion in Eq. (C.19), originating from the presence of \(e^{{{\textbf {R}}}}\) in Eq. (23), naturally terminates, i.e. it stops after two terms. Eventually, the four elementary contractions read as

$$\begin{aligned} {\rho }_{k_1k_2}(l,r)&= \langle \varPhi (l) | ^R{\beta ^{\dagger }_{k_2}} \, ^R{\beta _{k_1}} | \varPhi (l) \rangle \nonumber \\&= \left[ V^{*}(l)V^T(l) + U(l){{\textbf {z}}}(l,r)V^T(l)\right] _{k_1k_2} \nonumber \\&= +{\rho }_{k_1k_2}(l,l) + \left[ U(l){{\textbf {z}}}(l,r)V^T(l)\right] _{k_1k_2} \, , \end{aligned}$$
(C.23a)
$$\begin{aligned} {\kappa }_{k_1k_2}(l,r)&= \langle \varPhi (l) | ^R{\beta _{k_2}} \, ^R{\beta _{k_1}} | \varPhi (l) \rangle \nonumber \\&= \left[ V^{*}(l)U^T(l) + U(l){{\textbf {z}}}(l,r)U^T(l)\right] _{k_1k_2} \nonumber \\&= +{\kappa }_{k_1k_2}(l,l) + \left[ U(l){{\textbf {z}}}(l,r)U^T(l)\right] _{k_1k_2} \, \end{aligned}$$
(C.23b)
$$\begin{aligned} -\bar{{\kappa }}^{*}_{k_1k_2}(l,r)&= \langle \varPhi (l) | ^R{\beta ^{\dagger }_{k_2}} \, ^R{\beta ^{\dagger }_{k_1}} | \varPhi (l) \rangle \nonumber \\&= \left[ U^{*}(l)V^T(l) + V(l){{\textbf {z}}}(l,r)V^T(l)\right] _{k_1k_2} \nonumber \\&= -\bar{{\kappa }}^{*}_{k_1k_2}(l,l) + \left[ V(l){{\textbf {z}}}(l,r)V^T(l)\right] _{k_1k_2} \, , \end{aligned}$$
(C.23c)
$$\begin{aligned} -{\sigma }^{*}_{k_1k_2}(l,r)&= \langle \varPhi (l) | ^R{\beta _{k_2}} \, ^R{\beta ^{\dagger }_{k_1}} | \varPhi (l) \rangle \nonumber \\&= \left[ U^{*}(l)U^T(l) + V(l){{\textbf {z}}}(l,r)U^T(l)\right] _{k_1k_2} \nonumber \\&= -{\sigma }^{*}_{k_1k_2}(l,l) + \left[ V(l){{\textbf {z}}}(l,r)U^T(l)\right] _{k_1k_2}, \end{aligned}$$
(C.23d)

where Eqs. (C.21C.22) have been used. This completes the derivation of the off-diagonal Wick theorem where the explicit form of the elementary off-diagonal contractions in Eq. (C.23) reflects the asymmetric character of the approach, i.e. the expressions are anchored on the bra state \(\langle \varPhi (l) |\) and are a functional of the Thouless matrix \({{\textbf {z}}}(l,r)\) associated with the transition Bogoliubov transformation of Eqs. (1517).

Starting from Eq. (C.23) and using repeatedly relations associated with the unitarity of W(l) (Eq. (10)), one can symmetrize the elementary contractions by expressing them in terms of the Thouless matrices \({{\textbf {z}}}(l)\) and \({{\textbf {z}}}(r)\) associated withright states, respectively. Doing so, one recovers exactly Eqs. (39) and (41) obtained directly via the symmetric approach.

Appendix D: Connected kernel of a rank-3 operator

According to Eq. (38), the connected kernel associated with a rank-3 operator \(O\equiv {{\textbf {O}}}^{[0]}+{{\textbf {O}}}^{[2]}+{{\textbf {O}}}^{[4]}+{{\textbf {O}}}^{[6]}\) reads in terms of the off-diagonal elementary contractions as

$$\begin{aligned}&\frac{\langle \varPhi (l)| O |\varPhi (r) \rangle }{\langle \varPhi (l)|\varPhi (r) \rangle } = {{\textbf {O}}}^{[0]}\ + \frac{1}{2} \sum _{k_1k_2} {{\textbf {o}}}^{k_1k_2} \bar{{\kappa }}^{*}_{k_1k_2}(l,r) + \sum _{k_1l_1} {{\textbf {o}}}^{k_1}_{l_1} \rho _{l_1k_1}(l,r) + \frac{1}{2} \sum _{l_1l_2} {{\textbf {o}}}_{l_1l_2} {\kappa }_{l_1l_2}(l,r) \nonumber \\&\quad + \frac{1}{8} \sum _{k_1k_2k_3k_4} {{\textbf {o}}}^{k_1k_2k_3k_4} \bar{{\kappa }}^{*}_{k_1k_2}(l,r) \bar{{\kappa }}^{*}_{k_3k_4} (l,r)+ \frac{1}{2} \sum _{k_1k_2k_3l_1} {{\textbf {o}}}^{k_1k_2k_3}_{l_1} \rho _{l_1k_1}(l,r) \bar{{\kappa }}^{*}_{k_2k_3}(l,r) \nonumber \\&\quad + \frac{1}{2} \sum _{k_1k_2l_1l_2} {{\textbf {o}}}^{k_1k_2}_{l_1l_2} \rho _{l_1k_1}(l,r) \rho _{l_2k_2}(l,r) + \frac{1}{4} \sum _{k_1k_2l_1l_2} {{\textbf {o}}}^{k_1k_2}_{l_1l_2}\bar{{\kappa }}^{*}_{k_1k_2}(l,r){\kappa }_{l_1l_2} (l,r)\nonumber \\&\quad + \frac{1}{2} \sum _{k_1l_1l_2l_3} {{\textbf {o}}}^{k_1}_{l_1l_2l_3} \rho _{l_1k_1} (l,r) {\kappa }_{l_2l_3}(l,r) + \frac{1}{8} \sum _{l_1l_2l_3l_4} {{\textbf {o}}}_{l_1l_2l_3l_4} {\kappa }_{l_1l_2}(l,r) {\kappa }_{l_3l_4}(l,r)\nonumber \\&\quad + \frac{1}{48} \sum _{k_1k_2k_3k_4k_5k_6} {{\textbf {o}}}^{k_1k_2k_3k_4k_5k_6} \bar{{\kappa }}^{*}_{k_1k_2}(l,r) \bar{{\kappa }}^{*}_{k_3k_4} (l,r) \bar{{\kappa }}^{*}_{k_5k_6} (l,r) \nonumber \\&\quad + \frac{1}{8} \sum _{k_1k_2k_3k_4k_5l_1} {{\textbf {o}}}^{k_1k_2k_3k_4k_5}_{l_1} \rho _{l_1k_1}(l,r) \bar{{\kappa }}^{*}_{k_2k_3}(l,r) \bar{{\kappa }}^{*}_{k_4k_5}(l,r) \nonumber \\&\quad + \frac{1}{16} \sum _{k_1k_2k_3k_4l_1l_2} {{\textbf {o}}}^{k_1k_2k_3k_4}_{l_1l_2} \bar{{\kappa }}^{*}_{k_1k_2}(l,r) \bar{{\kappa }}^{*}_{k_3k_4}(l,r) \kappa _{l_1l_2} (l,r)\nonumber \\&\quad + \frac{1}{4} \sum _{k_1k_2k_3k_4l_1l_2} {{\textbf {o}}}^{k_1k_2k_3k_4}_{l_1l_2} \rho _{l_1k_1}(l,r) \rho _{l_2k_2}(l,r) \bar{{\kappa }}^{*}_{k_3k_4} (l,r) \nonumber \\&\quad + \frac{1}{6} \sum _{k_1k_2k_3l_1l_2l_3} {{\textbf {o}}}^{k_1k_2k_3}_{l_1l_2l_3} \rho _{l_1k_1}(l,r) \rho _{l_2k_2}(l,r) \rho _{l_3k_3}(l,r)\nonumber \\&\quad + \frac{1}{4} \sum _{k_1k_2k_3l_1l_2l_3} {{\textbf {o}}}^{k_1k_2k_3}_{l_1l_2l_3} \rho _{l_1k_1}(l,r) \bar{{\kappa }}^{*}_{k_2k_3}(l,r) \kappa _{l_2l_3}(l,r) \nonumber \\&\quad + \frac{1}{16} \sum _{k_1k_2l_1l_2l_3l_4} {{\textbf {o}}}^{k_1k_2}_{l_1l_2l_3l_4} \bar{{\kappa }}^{*}_{k_1k_2}(l,r) \kappa _{l_1l_2} (l,r) \kappa _{l_3l_4} (l,r)\nonumber \\&\quad + \frac{1}{4} \sum _{k_1k_2l_1l_2l_3l_4} {{\textbf {o}}}^{k_1k_2}_{l_1l_2l_3l_4} \rho _{l_1k_1}(l,r) \rho _{l_2k_2}(l,r) \kappa _{l_3l_4}(l,r) \nonumber \\&\quad + \frac{1}{8} \sum _{k_1l_1l_2l_3l_4l_5} {{\textbf {o}}}^{k_1}_{l_1l_2l_3l_4l_5} \rho _{l_1k_1}(l,r) \kappa _{l_2l_3}(l,r) \kappa _{l_4l_5}(l,r) \nonumber \\&\quad +\frac{1}{48} \sum _{l_1l_2l_3l_4l_5l_6} {{\textbf {o}}}_{l_1l_2l_3l_4l_5l_6} \kappa _{l_1l_2}(l,r) \kappa _{l_3l_4}(l,r) \kappa _{l_5l_6}(l,r). \end{aligned}$$
(D.24)

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Porro, A., Duguet, T. On the off-diagonal Wick’s theorem and Onishi formula. Eur. Phys. J. A 58, 197 (2022). https://doi.org/10.1140/epja/s10050-022-00843-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epja/s10050-022-00843-2

Navigation