Skip to main content
Log in

Importance truncation in non-perturbative many-body techniques

Gorkov self-consistent Green’s function calculations

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

Abstract

Expansion many-body methods correspond to solving complex tensor networks. The (iterative) solving of the network and the (repeated) storage of the unknown tensors requires a computing power growing polynomially with the size of basis of the one-body Hilbert space one is working with. Thanks to current computer capabilities, ab initio calculations of nuclei up to mass \(A\sim 100\) delivering a few percent accuracy are routinely feasible today. However, the runtime and memory costs become quickly prohibitive as one attempts (possibly at the same time) (i) to reach out to heavier nuclei, (ii) to employ symmetry-breaking reference states to access open-shell nuclei and (iii) to aim for yet a greater accuracy. The challenge is particularly exacerbated for non-perturbative methods involving the repeated storage of (high-rank) tensors obtained via iterative solutions of non-linear equations. The present work addresses the formal and numerical implementations of so-called importance truncation (IT) techniques within the frame of one particular non-perturbative expansion method, i.e., Gorkov Self-Consistent Green’s Function (GSCGF) theory, with the goal to eventually overcome above-mentioned limitations. By a priori truncating irrelevant tensor entries, IT techniques are shown to reduce the storage to less than 1% of its original cost in realistic GSCGF calculations performed at the ADC(2) level while maintaining a 1% accuracy on the correlation energy. The future steps will be to extend the present development to the next, e.g., ADC(3), truncation level and to SCGF calculations applicable to doubly open-shell nuclei.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

Data Availability Statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: This is a theoretical work and no experimental data were generated. Data associated to the theoretical simulations can be obtained from the authors upon request.]

Notes

  1. For a more complete account of the method, the reader is referred to Refs. [9, 35].

  2. In practice, two different Lagrange parameters are needed to constrain proton and neutron numbers independently, i.e. \(\lambda A\) actually stands for \(\lambda _Z Z + \lambda _N N\).

  3. While \(\mathcal {E}\) is diagonal at the ADC(2) level, it will become full at the ADC(3) level.

  4. Let \(N_b\) be the number of poles resulting from a previous HFB calculation. In such a case \(N_b\) is the size of the \(\varXi ^{(1)}\) matrix appearing in Eq. (20) and provides the size of the quasiparticle basis used to compute higher order contributions in perturbation theory, and in particular the second-order self-energy. From the Lehmann representation of the latter quantity – expressed in terms of three-quasiparticle configurations – the total number of configurations is given by

    $$\begin{aligned} N_s\approx \left( {\begin{array}{c}N_b\\ 3\end{array}}\right) \approx \frac{N_b^3}{6}, \end{aligned}$$
    (21)

    where the first equivalence cannot be considered exact since not all three-quasiparticle configurations are allowed to exist due to selection rules linked to angular momentum coupling [35].

  5. Going from ADC(2) to ADC(3) does not actually increase the dimension of the matrix \(\varXi \) but reduces tremendously the sparsity of the submatrix \(\mathcal {E}\). This leads to a drastic increase of the non-zero matrix elements to compute and to the need for more involved techniques to perform the diagonalisation. Eventually going from ADC(3) to ADC(4) does indeed increase the size of \(\varXi \) with additional sub-blocks running over five-pole indices \(k_1k_2k_3k_4k_5\).

  6. In a spherically symmetric implementation, as the one considered here, one only needs to take into account \(\mathcal {C}\) and \(\mathcal {D}\) (not \(\bar{\mathcal {C}}\) and \(\bar{\mathcal {D}}\)). This assumption is used throughout the present work. See Ref. [9] for details.

  7. The calculations whose results are presented in Sec. 4 are based on the use of the spherical harmonic oscillator basis. In this case, this choice corresponds to selecting the state with the lowest principal quantum number.

  8. In the case of \(\hbox {NNLO}_{\text {sat}}\), SRG-evolution is known to induce substantial four-body (and possibly higher-body) forces that make the corresponding results not equivalent to the ones obtained with the original Hamiltonian. Being interested in the performance of IT relative to a given starting Hamiltonian (and not in generating fully physical results) this drawback does not constitute an issue for the present work.

  9. In principle, the measure \(e_{\,\,_{1C}}^{\,\,_{\text {IT}}}\) is the cheapest to compute among the three. However, the cost of performing the sums over one one-particle index in Eqs. (25b) and (25c) is negligible in practice such that the three \(e_{\,\,_{1}}^{\,\,_{\text {IT}}}\) measures can be considered as equivalent also from the computational point of view.

  10. Other observables will be considered in Sect. 4.8.

  11. In view of the above discussion, one must keep in mind that whenever the energy curve bends down as R decreases, it signals that the perturbative correction dominates and that the one-body propagator solution of the IT Gorkov’s equation tends towards the uncorrelated one.

  12. In Dyson SCGF, elementary excitations correspond to many-particle-many-hole configurations. For instance, at the ADC(2) truncation level, Dyson’s counterpart to three-quasiparticle configurations are two-particle-one-hole and two-hole-one-particle combinations.

  13. In the present j-coupled scheme, such sums eventually boil down to sums over principal quantum numbers only [9].

  14. The position of the peak is driven by the powers of zero. The width is generated by the different values of interaction matrix elements, denominators and additional sums that have been neglected in this qualitative analysis.

  15. We recall that in the present j-scheme implementation Gorkov’s matrix has a block-diagonal structure, with different blocks being labelled by \(\ell , j\) and the isospin projection [9].

  16. In Ref. [32] the extrapolation technique was designed and employed within the frame of no-core shell model calculations as a function of \(\epsilon _\text {min}\). Because of the self-consistent, highly non-linear, character of the ADC(2) equations, \(\epsilon _\text {min}\) or R do not provide suitable enough variables for the construction of a smooth function to be subsequently extrapolated. Instead, \(E_\zeta \) is presently studied as a function of the perturbative correction itself, i.e. the function \(E_\zeta (E_\text {corr})\) is constructed and extrapolated to the limit \(E_\text {corr} \rightarrow 0\). It can indeed be shown that \(E_\text {corr}(\epsilon _\text {min})\) is a monotonous function going smoothly to zero as \(\epsilon _\text {min} \rightarrow 0\) [48]. It can thus play the role of a measure of the error itself, even in absence of the exact solution.

  17. Results obtained with the method-driven measure \(e_{\,\,_{1}}^{\,\,_{\text {IT}}}\) yield analogous results.

  18. One must keep in mind that such states are most probably not converged with respect to the many-body truncation when working at the ADC(2) approximation level.

  19. In reality the ADC(3) matrix has some sparsity, such that the present estimates can be considered as an upper limit.

  20. The term “energy” is used everywhere even if, strictly speaking, the operator presently used is the grand potential and not the Hamiltonian.

  21. Using \(\varOmega (\lambda )\), all quantities of interest (e.g. self-energy, one-body propagator, two-body density matrix...) become themselves \(\lambda \) dependent.

  22. The normalized trace accounts for a 1/2 factor due to the \(2\times 2\) dimension of the quantities defined in Nambu space.

  23. The quantity \(\varvec{\varSigma }_n(\lambda )\) indirectly depends on the coupling constant \(\lambda \) given that it is itself a functional of the propagator \(\varvec{\mathcal {G}}(\lambda )\). Consequently, \(\varvec{\varSigma }_n(\lambda )\) is not the \(n^\text {th}\)-order contribution in a perturbative expansion, but simply the part of the exact functional \(\varvec{\varSigma }(\lambda )\) containing \((2n-1)\) propagators \(\varvec{\mathcal {G}}(\lambda )\) and n interaction vertices [54].

  24. The ratio appearing in Eq. (A.13) must be read as the matrix product \(\varvec{\mathcal {G}}(\omega ,\lambda )\varvec{\mathcal {G}}^{-1}_0(\omega )\), where \(\varvec{\mathcal {G}}^{-1}_0(\omega )\) is the inverse matrix of \(\varvec{\mathcal {G}}_0(\omega )\) in both the one-body Hilbert space \({{\mathcal {H}}}_1\) and the \(2\times 2\) Nambu space.

  25. While not obvious at first sight, the expression in Eq. (A.19) is fully symmetric under the exchange of any pair among the four quasi-particle indices such that one can arbitrarily isolate any triplet to define the IT measure.

References

  1. J. Langhammer, R. Roth, C. Stumpf, Phys. Rev. C 86, 054315 (2012). https://doi.org/10.1103/PhysRevC.86.054315

    Article  ADS  Google Scholar 

  2. A. Tichai, J. Langhammer, S. Binder, R. Roth, Phys. Lett. B 756, 283 (2016). https://doi.org/10.1016/j.physletb.2016.03.029

    Article  ADS  Google Scholar 

  3. B. Hu, F. Xu, Z. Sun, J.P. Vary, T. Li, Phys. Rev. C 94, 014303 (2016). https://doi.org/10.1103/PhysRevC.94.014303

    Article  ADS  Google Scholar 

  4. A. Tichai, E. Gebrerufael, K. Vobig, R. Roth, Phys. Lett. B 786, 448 (2018a). https://doi.org/10.1016/j.physletb.2018.10.029

    Article  ADS  Google Scholar 

  5. A. Tichai, P. Arthuis, T. Duguet, H. Hergert, V. Somà, R. Roth, Phys. Lett. B 786, 195 (2018b). https://doi.org/10.1016/j.physletb.2018.09.044

    Article  ADS  Google Scholar 

  6. P. Arthuis, T. Duguet, A. Tichai, R.-D. Lasseri, J.-P. Ebran, Comput. Phys. Commun. 240, 202 (2019). https://doi.org/10.1016/j.cpc.2018.11.023

    Article  ADS  Google Scholar 

  7. A. Tichai, R. Roth, T. Duguet, Front. Phys. 8, 164 (2020). https://doi.org/10.3389/fphy.2020.00164

    Article  Google Scholar 

  8. W.H. Dickhoff, C. Barbieri, Prog. Part. Nucl. Phys. 52, 377 (2004). https://doi.org/10.1016/j.ppnp.2004.02.038

    Article  ADS  Google Scholar 

  9. V. Somà, T. Duguet, C. Barbieri, Phys. Rev. C 84, 064317 (2011). https://doi.org/10.1103/PhysRevC.84.064317

    Article  ADS  Google Scholar 

  10. A. Carbone, A. Cipollone, C. Barbieri, A. Rios, A. Polls, Phys. Rev. C 88, 054326 (2013). https://doi.org/10.1103/PhysRevC.88.054326

    Article  ADS  Google Scholar 

  11. V. Somà, A. Cipollone, C. Barbieri, P. Navrátil, T. Duguet, Phys. Rev. C 89, 061301 (2014). https://doi.org/10.1103/PhysRevC.89.061301

    Article  ADS  Google Scholar 

  12. V. Lapoux, V. Somà, C. Barbieri, H. Hergert, J.D. Holt, S. Stroberg, Phys. Rev. Lett. 117, 052501 (2016). https://doi.org/10.1103/PhysRevLett.117.052501

    Article  ADS  Google Scholar 

  13. T. Duguet, V. Somà, S. Lecluse, C. Barbieri, P. Navrátil, Phys. Rev. C 95, 034319 (2017). https://doi.org/10.1103/PhysRevC.95.034319

    Article  ADS  Google Scholar 

  14. F. Raimondi, C. Barbieri, Phys. Rev. C 99, 054327 (2019). https://doi.org/10.1103/PhysRevC.99.054327

    Article  ADS  Google Scholar 

  15. P. Arthuis, C. Barbieri, M. Vorabbi, P. Finelli, Phys. Rev. Lett. 125, 182501 (2020). https://doi.org/10.1103/PhysRevLett.125.182501

    Article  ADS  Google Scholar 

  16. V. Somà, Front. Phys. 8, 340 (2020). https://doi.org/10.3389/fphy.2020.00340

    Article  Google Scholar 

  17. G. Hagen, M. Hjorth-Jensen, G. Jansen, R. Machleidt, T. Papenbrock, Phys. Rev. Lett. 108, 242501 (2012). https://doi.org/10.1103/PhysRevLett.108.242501

    Article  ADS  Google Scholar 

  18. S. Binder, P. Piecuch, A. Calci, J. Langhammer, P. Navrátil, R. Roth, Phys. Rev. C 88, 054319 (2013). https://doi.org/10.1103/PhysRevC.88.054319

    Article  ADS  Google Scholar 

  19. S. Binder, J. Langhammer, A. Calci, R. Roth, Phys. Lett. B 736, 119 (2014). https://doi.org/10.1016/j.physletb.2014.07.010

    Article  ADS  Google Scholar 

  20. G. Hagen, T. Papenbrock, M. Hjorth-Jensen, D.J. Dean, Rept. Prog. Phys. 77, 096302 (2014). https://doi.org/10.1088/0034-4885/77/9/096302

    Article  ADS  Google Scholar 

  21. A. Signoracci, T. Duguet, G. Hagen, G. Jansen, Phys. Rev. C 91, 064320 (2015). https://doi.org/10.1103/PhysRevC.91.064320

    Article  ADS  Google Scholar 

  22. T.D. Morris, J. Simonis, S.R. Stroberg, C. Stumpf, G. Hagen, J.D. Holt, G.R. Jansen, T. Papenbrock, R. Roth, A. Schwenk, Phys. Rev. Lett. 120, 152503 (2018). https://doi.org/10.1103/PhysRevLett.120.152503

    Article  ADS  Google Scholar 

  23. K. Tsukiyama, S.K. Bogner, A. Schwenk, Phys. Rev. Lett. 106, 222502 (2011). https://doi.org/10.1103/PhysRevLett.106.222502

    Article  ADS  Google Scholar 

  24. H. Hergert, S.K. Bogner, T.D. Morris, A. Schwenk, K. Tsukiyama, Phys. Rept. 621, 165 (2016). https://doi.org/10.1016/j.physrep.2015.12.007

    Article  ADS  Google Scholar 

  25. H. Hergert, S.K. Bogner, J.G. Lietz, T.D. Morris, S. Novario, N.M. Parzuchowski, F. Yuan, Lect. Notes Phys. 936, 477 (2017). https://doi.org/10.1007/978-3-319-53336-0_10

    Article  ADS  Google Scholar 

  26. N.M. Parzuchowski, S.R. Stroberg, P. Navrátil, H. Hergert, S.K. Bogner, Phys. Rev. C 96, 034324 (2017). https://doi.org/10.1103/PhysRevC.96.034324

    Article  ADS  Google Scholar 

  27. M. Heinz, A. Tichai, J. Hoppe, K. Hebeler, A. Schwenk, Phys. Rev. C 103, 044318 (2021). https://doi.org/10.1103/PhysRevC.103.044318

    Article  ADS  Google Scholar 

  28. T. Duguet, J. Phys. G: Nucl. Part. Phys. 42, 025107 (2015). https://doi.org/10.1088/0954-3899/42/2/025107

    Article  ADS  Google Scholar 

  29. T. Duguet, A. Signoracci, J. Phys. G: Nucl. Part. Phys. 44, 015103 (2017). https://doi.org/10.1088/1361-6471/aa5d3e

    Article  ADS  Google Scholar 

  30. Y. Qiu, T.M. Henderson, J. Zhao, G.E. Scuseria, J. Chem. Phys. 147, 064111 (2017). https://doi.org/10.1063/1.4991020

    Article  ADS  Google Scholar 

  31. Y. Qiu, T. Henderson, T. Duguet, G. Scuseria, Phys. Rev. C 99, 044301 (2019). https://doi.org/10.1103/PhysRevC.99.044301

    Article  ADS  Google Scholar 

  32. R. Roth, Phys. Rev. C 79, 064324 (2009). https://doi.org/10.1103/PhysRevC.79.064324

    Article  ADS  Google Scholar 

  33. R. Roth, P. Navrátil, Phys. Rev. Lett. 99, 092501 (2007). https://doi.org/10.1103/PhysRevLett.99.092501

    Article  ADS  Google Scholar 

  34. A. Tichai, J. Ripoche, T. Duguet, Eur. Phys. J. A 55, 90 (2019). https://doi.org/10.1140/epja/i2019-12758-6

    Article  ADS  Google Scholar 

  35. V. Somà, C. Barbieri, T. Duguet, Phys. Rev. C 89, 024323 (2014). https://doi.org/10.1103/PhysRevC.89.024323

    Article  ADS  Google Scholar 

  36. J. Ripoche, A. Tichai, T. Duguet, Eur. Phys. J. A 56, 40 (2020). https://doi.org/10.1140/epja/s10050-020-00045-8

    Article  ADS  Google Scholar 

  37. M. Frosini, T. Duguet, B. Bally, Y. Beaujeault-Taudière, J.P. Ebran, V. Somà, Eur. Phys. J. A 57, 151 (2021). https://doi.org/10.1140/epja/s10050-021-00458-z

    Article  ADS  Google Scholar 

  38. P. Ring, P. Schuck, The Nuclear Many-Body Problem (Theoretical and Mathematical Physics) (Springer, New York, 2005). https://www.xarg.org/ref/a/354021206X/

  39. J. Schirmer, L.S. Cederbaum, O. Walter, Phys. Rev. A 28, 1237 (1983). https://doi.org/10.1103/PhysRevA.28.1237

    Article  ADS  Google Scholar 

  40. A. Cipollone, C. Barbieri, P. Navrátil, Phys. Rev. Lett. 111, 062501 (2013). https://doi.org/10.1103/PhysRevLett.111.062501

    Article  ADS  Google Scholar 

  41. F. Raimondi, C. Barbieri, Phys. Rev. C 97, 054308 (2018). https://doi.org/10.1103/PhysRevC.97.054308

    Article  ADS  Google Scholar 

  42. C. Barbieri, M. Hjorth-Jensen, Phys. Rev. C 79, 064313 (2009). https://doi.org/10.1103/PhysRevC.79.064313

    Article  ADS  Google Scholar 

  43. V. Somà, P. Navrátil, F. Raimondi, C. Barbieri, T. Duguet, Phys. Rev. C 101, 014318 (2020). https://doi.org/10.1103/PhysRevC.101.014318

    Article  ADS  Google Scholar 

  44. A. Cipollone, C. Barbieri, P. Navrátil, Phys. Rev. C 92, 014306 (2015). https://doi.org/10.1103/PhysRevC.92.014306

    Article  ADS  Google Scholar 

  45. A. Ekström, G.R. Jansen, K.A. Wendt, G. Hagen, T. Papenbrock, B.D. Carlsson, C. Forssén, M. Hjorth-Jensen, P. Navrátil, W. Nazarewicz, Phys. Rev. C 91, 051301 (2015). https://doi.org/10.1103/PhysRevC.91.051301

    Article  ADS  Google Scholar 

  46. A. Porro,Importance-Truncation techniques in Gorkov-Green’s functions calculations for atomic nuclei. Master’s thesis, Università degli Studi di Milano (2020)

  47. V. Somà, C. Barbieri, T. Duguet, P. Navrátil, Eur. Phys. J. A 57, 135 (2021). https://doi.org/10.1140/epja/s10050-021-00437-4

    Article  ADS  Google Scholar 

  48. R. Buenker, S. Peyerimhoff, Theoret. Chim. Acta 39, 217 (1975). https://doi.org/10.1007/BF00555301

    Article  Google Scholar 

  49. H.-W. Hammer, S. König, U. van Kolck, Rev. Mod. Phys. 92, 025004 (2020). https://doi.org/10.1103/RevModPhys.92.025004

    Article  ADS  Google Scholar 

  50. M. Hjorth-Jensen, M. P. Lombardo, U. van Kolck, eds., An Advanced Course in Computational Nuclear Physics (Springer, New York, 2017). https://doi.org/10.1007/978-3-319-53336-0

  51. K.A. Brueckner, Phys. Rev. 100, 36 (1955). https://doi.org/10.1103/PhysRev.100.36

    Article  ADS  Google Scholar 

  52. J. Goldstone, Proc. R. Soc. Lond. Ser. A. Math. Phys. Sci. 239, 267 (1957). https://doi.org/10.1098/rspa.1957.0037

  53. R.P. Feynman, Phys. Rev. 56, 340 (1939). https://doi.org/10.1103/PhysRev.56.340

    Article  ADS  Google Scholar 

  54. P. Nozières, Le problème à N corps (Dunod, 1963). https://books.google.fr/books?id=EsPQAAAAMAAJ

  55. W. Kohn, J.M. Luttinger, Phys. Rev. 118, 41 (1960). https://doi.org/10.1103/PhysRev.118.41

    Article  ADS  MathSciNet  Google Scholar 

  56. J.M. Luttinger, J.C. Ward, Phys. Rev. 118, 1417 (1960). https://doi.org/10.1103/PhysRev.118.1417

    Article  ADS  MathSciNet  Google Scholar 

  57. N.E. Dahlen, U. von Barth, J. Chem. Phys. 120, 6826 (2004). https://doi.org/10.1063/1.1650307

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The authors wish to thank C. Barbieri and G. Colò for fruitful discussions, P. Arthuis for useful remarks on the manuscript and P. Navrátil for providing the interaction matrix elements used in the numerical calculations. 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. A.T. was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 279384907 – SFB 1245 and by the Max Planck Society. Calculations were performed using HPC resources from GENCI-TGCC (Contracts Nos. A007057392, A009057392).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. Porro.

Additional information

Communicated by Arnau Rios Huguet.

Appendix A: Energy-driven measure

Appendix A: Energy-driven measure

1.1 Appendix A.1: Interaction and correlation ground-state energies

The objective is to define an energy-driven measure based on perturbation theory and expressed in terms of the tensors at play in ADC(2) Gorkov SCGF calculations. Let us thus introduce

$$\begin{aligned} \varOmega (\lambda )\equiv \varOmega _0+\lambda \varOmega _1, \end{aligned}$$
(A.1)

where \(\lambda \) is the perturbation parameter leading the system from the uncorrelated state (\(\lambda =0\)) to the fully correlated (interacting) one (\(\lambda =1\)). The unperturbed grand potential \(\varOmega _0\) is taken to be of one-body character such that its ground-state is the Bogoliubov product state \(| \varPhi _0 \rangle \), i.e.,

$$\begin{aligned} \varOmega _0 | \varPhi _0 \rangle = \varOmega ^{0}_0| \varPhi _0 \rangle . \end{aligned}$$
(A.2)

Correspondingly, the ground state of \(\varOmega (\lambda )\) fulfils Schrödinger’s equation

$$\begin{aligned} \varOmega (\lambda )| \varPsi _0(\lambda ) \rangle =\varOmega _0(\lambda )| \varPsi _0(\lambda ) \rangle . \end{aligned}$$
(A.3)

From Eqs. (A.2) and (A.3), one obtains

$$\begin{aligned} \varOmega _0(\lambda )&=\langle \varPsi _0(\lambda ) |\varOmega (\lambda )| \varPsi _0(\lambda ) \rangle \nonumber \\&=\langle \varPsi _0(\lambda ) |\varOmega _0| \varPsi _0(\lambda ) \rangle +\lambda \langle \varPsi _0(\lambda ) |\varOmega _1| \varPsi _0(\lambda ) \rangle \nonumber \\&\equiv \varOmega _{\text {noint}} (\lambda )+ \varOmega _{\text {int}}(\lambda ) . \end{aligned}$$
(A.4)

as well as

$$\begin{aligned} \varOmega _0(\lambda )&= \frac{\langle \varPhi _0|\varOmega |\varPsi _0(\lambda )\rangle }{\langle \varPhi _0|\varPsi _0(\lambda )\rangle } \nonumber \\&=\frac{\langle \varPhi _0|\varOmega _0|\varPsi _0(\lambda )\rangle }{\langle \varPhi _0|\varPsi _0(\lambda )\rangle } + \lambda \frac{\langle \varPhi _0|\varOmega _1|\varPsi _0(\lambda )\rangle }{\langle \varPhi _0|\varPsi _0(\lambda )\rangle } \nonumber \\&\equiv \varOmega _0^0 + \varDelta \varOmega _0(\lambda ) . \end{aligned}$$
(A.5)

The two above equations define the so-called interaction and correlation energiesFootnote 20, respectively, such that \(\varOmega _{\text {int}}(1)\equiv \varOmega _{\text {int}}\) and \(\varDelta \varOmega _0(1)\equiv \varDelta \varOmega _0\) for the fully interacting system.

1.2 Appendix A.2: Perturbative expansion of \(\varDelta \varOmega _0\)

The correlation energy is amenable to a perturbative expansion on the basis of Goldstone’s theorem [51, 52]

$$\begin{aligned} \varDelta \varOmega _0=\langle \varPhi _0 |\varOmega _1\sum _{n=0}^{\infty }\bigg (\frac{1}{\varOmega _0^0-\varOmega _0}H_1\bigg )^n| \varPhi _0 \rangle _{C}, \end{aligned}$$
(A.6)

where the index C denotes the fact that only so-called connected terms are retained. However, this expression does not explicitly relate the correlation energy to the propagator, the self energy and thus the many-body tensors at play in Gorkov SCGF calculations.

1.3 Appendix A.3: Linking \(\varDelta \varOmega _0\) to the self energy

To achieve the needed connection, one must first exploit Hellmann-Feynman’s theorem [53] to relate the correlation energy to the interaction energy according to

$$\begin{aligned} \varDelta \varOmega _0(\lambda )=\int _0^{\lambda }\frac{d\lambda '}{\lambda '}\varOmega _{\text {int}}(\lambda '), \end{aligned}$$
(A.7)

and further exploit the link the interaction energy entertains with the one-body propagator.

Expressing \(\varOmega _{\text {int}}(\lambda )\) in terms of the two-body density matrix and exploiting the definition of the one-body self-energy connecting the one-body Green’s function to the two-times reduction of the two-body Green’s functionFootnote 21, one easily obtains

$$\begin{aligned} \varOmega _{\text {int}}(\lambda )= -\frac{1}{2} \text {Tr}\,\bigl \{\varvec{\varSigma }(\lambda )\,\varvec{\mathcal {G}}(\lambda )\bigr \}, \end{aligned}$$
(A.8)

where the triple trace over the one-body Hilbert space, Nambu space and the frequency defined throughFootnote 22

$$\begin{aligned} \text {Tr} \,\equiv -\text {Tr}_{\mathcal {H}_1}\frac{1}{2}\text {Tr}_{\mathcal {G}}\oint \frac{d\omega }{2\pi i}, \end{aligned}$$
(A.9)

has been used to shorten the expression. Substituting Eq. (A.8) into Eq. (A.7) one obtains

$$\begin{aligned} \varDelta \varOmega _0(\lambda )=-\frac{1}{2}\int _0^{\lambda }\frac{d\lambda '}{\lambda '}\text {Tr}\,\bigl \{\varvec{\varSigma }(\lambda ')\,\varvec{\mathcal {G}}(\lambda ')\bigr \}. \end{aligned}$$
(A.10)

Next, the self-energy is formally expanded in powers of the residual interaction according toFootnote 23

$$\begin{aligned} \varvec{\varSigma }(\lambda )=\sum _{n=1}^{\infty }\lambda ^n\varvec{\varSigma }_n(\lambda ), \end{aligned}$$
(A.11)

such that Eq. (A.10) becomes

$$\begin{aligned} \varDelta \varOmega _0(\lambda )=-\frac{1}{2}\sum _{n=1}^{\infty }\int _0^{\lambda }d\lambda '\,\lambda '^{n-1}\text {Tr}\, \varvec{\varSigma }_n(\lambda ')\varvec{\mathcal {G}}(\lambda '). \end{aligned}$$
(A.12)

Following standard steps [54], the integration over the coupling strength can be performed in Eq. (A.12) to eventually obtainFootnote 24

$$\begin{aligned} \varDelta \varOmega _0(\lambda )=\varPhi [\varvec{\mathcal {G}}(\lambda )]+ \text {Tr}\,\bigg \{\frac{\varvec{\mathcal {G}}(\lambda )}{\varvec{\mathcal {G}}_0}-1\bigg \}-\text {Tr}\,\bigg \{\ln \frac{\varvec{\mathcal {G}}(\lambda )}{\varvec{\mathcal {G}}_0}\bigg \}, \nonumber \\ \end{aligned}$$
(A.13)

where the Luttinger-Ward functional [55, 56] adapted to Gorkov’s formalism was introduced

$$\begin{aligned} \varPhi [\varvec{\mathcal {G}}(\lambda )]\equiv -\frac{1}{2}\sum _{n=1}^{\infty }\frac{\lambda ^n}{n}\text {Tr}\,\varvec{\varSigma }_n(\lambda )\varvec{\mathcal {G}}(\lambda ). \end{aligned}$$
(A.14)

Taking \(\lambda = 1\) in Eq. (A.13), one obtains the fully interacting correlation energy \(\varDelta \varOmega _0\) under the form of a functional of the fully interacting propagator \(\varvec{\mathcal {G}}(\lambda )\). The \(\varPhi \)-functional \(\varPhi [\varvec{\mathcal {G}}(\lambda )]\) invokes the convolution of \(\varvec{\varSigma }_n(\lambda )\), itself a functional of \(\varvec{\mathcal {G}}(\lambda )\), with \(\varvec{\mathcal {G}}(\lambda )\). As a result, \(\varPhi [\varvec{\mathcal {G}}(\lambda )]\) is made out of the complete set of two-particle irreducible skeleton vacuum-to-vacuum diagrams whose lines denote the fully interacting propagator \(\varvec{\mathcal {G}}(\lambda )\).

1.4 Appendix A.4: Perturbative connection

In the present context, we are interested in a perturbative connection between the correlation energy and the one-body self energy. Starting from Eq. (A.13), it is easy to prove that \(\varDelta \varOmega _0\) is stationary with respect to any variation of the propagator, i.e., \(\varvec{\mathcal {G}}(\omega )\) solution of Gorkov’s equation delivers an extremum of \(\varDelta \varOmega _0\).

Concretely, this variational property translates into the possibility to alter the employed propagator without significantly affecting the correlation energy. This property was numerically validated in Ref. [57]. Exploiting this feature, the correlation energy is presently approximated by evaluating its expression at the uncorrelated propagator rather at the interacting one, i.e., using

$$\begin{aligned} \varDelta \varOmega _0\simeq \varDelta \varOmega _0[\varvec{\mathcal {G}}_0]=\varPhi [\varvec{\mathcal {G}}_0]. \end{aligned}$$
(A.15)

Based on this approximation, a perturbation-theory-like series is obtained for the correlation energy such that the \(n^\text {th}\) order reads as

$$\begin{aligned} \varDelta \varOmega _0^{(n)}[\varvec{\mathcal {G}}_0]=\varPhi ^{(n)}[\varvec{\mathcal {G}}_0]=-\frac{1}{2n}\, \text {Tr}\,\varvec{\varSigma }_n[\varvec{\mathcal {G}}_0]\,\varvec{\mathcal {G}}_0. \end{aligned}$$
(A.16)

One must note that Eqs. (A.6) and (A.15) do not provide, in general, the same expansion, i.e., while the first one is exact when summed to all orders, the second one is approximate. However, employing an uncorrelated propagator \(\varvec{\mathcal {G}}_0\) associated with the solution of the HFB mean-field equations, one can prove that the two expansions match exactly up to second-order. This exact matching has been validated numerically and is used in the following to compute the second-order correlation energy within Gorkov SCGF formalism.

The last step of the derivation thus consists of exploiting Eq. (A.16) for \(n=2\) and express the result in terms of the many-body tensors making up the ADC(2) self energy. For a spherically symmetric implementation, the following identities hold

$$\begin{aligned} \text {Tr}\{\varSigma ^{12}\mathcal {G}^{21}\}&=\text {Tr}\{\varSigma ^{21}\mathcal {G}^{12}\}, \end{aligned}$$
(A.17a)
$$\begin{aligned} \text {Tr}\{\varSigma ^{11}\mathcal {G}^{11}\}&=\text {Tr}\{\varSigma ^{22}\mathcal {G}^{22}\}, \end{aligned}$$
(A.17b)

eventually leading to the second-order correlation energy expression

$$\begin{aligned} \varDelta \varOmega _0^{(2)}[\varvec{\mathcal {G}}_0]=&\frac{1}{4}\sum _{ab}\oint \frac{d\omega }{2\pi 1}\bigl \{\varSigma ^{11(2)}_{ab}(\omega )\mathcal {G}_{0\,ba}^{11}(\omega ) \nonumber \\&+\varSigma ^{12(2)}_{ab}(\omega )\mathcal {G}_{0\,ba}^{21}(\omega )\bigr \} . \end{aligned}$$
(A.18)

Inserting Eqs. (9) and (13) into Eq. (A.18), the integration over the frequency can be performed to generate the needed form

$$\begin{aligned} \varDelta \varOmega _0^{(2)}[\varvec{\mathcal {G}}_0]&=\frac{1}{4}\sum _{\begin{array}{c} k_1k_2k_3k_4\\ ab \end{array}}\bigg \{\frac{{\mathcal {V}}_b^{k_4}{\mathcal {V}}_a^{k_4}\mathcal {C}_a^{k_1k_2k_3}\mathcal {C}_b^{k_1k_2k_3}}{E_{k_1k_2k_3}+\omega _{k_4}} \nonumber \\&\quad +\frac{\mathcal {U}_b^{k_4}\mathcal {U}_a^{k_4}{\mathcal {D}}_a^{k_1k_2k_3}{\mathcal {D}}_b^{k_1k_2k_3}}{E_{k_1k_2k_3}+\omega _{k_4}} \nonumber \\&\quad +\frac{{\mathcal {U}}_b^{k_4}{\mathcal {V}}_a^{k_4}\mathcal {C}_a^{k_1k_2k_3}\mathcal {D}_b^{k_1k_2k_3}}{E_{k_1k_2k_3}+\omega _{k_4}} \nonumber \\&\quad +\frac{\mathcal {V}_b^{k_4}\mathcal {U}_a^{k_4}{\mathcal {D}}_a^{k_1k_2k_3}{\mathcal {C}}_b^{k_1k_2k_3}}{E_{k_1k_2k_3}+\omega _{k_4}}\bigg \} \nonumber \\&=\frac{1}{4}\sum _{k_1k_2k_3k_4}\frac{\Big [\sum _a\big (\mathcal {V}_a^{k_4}\mathcal {C}_a^{k_1k_2k_3}+\mathcal {U}_a^{k_4}\mathcal {D}_a^{k_1k_2k_3}\big )\Big ]^2}{E_{k_1k_2k_3k_4}} \nonumber \\&=\sum _{k_1k_2k_3} e_{\,\,_{2}}^{\,\,_{\text {IT}}}(k_1k_2k_3) , \end{aligned}$$
(A.19)

where the last equality simply amounts to isolating three-pole contributionsFootnote 25\((k_1k_2k_3)\) and proceeds to a trivial rewriting. The above formula thus makes explicit the relation between the correlation energy and the energy-driven IT measure introduced in Eq. (26).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Porro, A., Somà, V., Tichai, A. et al. Importance truncation in non-perturbative many-body techniques. Eur. Phys. J. A 57, 297 (2021). https://doi.org/10.1140/epja/s10050-021-00606-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epja/s10050-021-00606-5

Navigation