Skip to main content
Log in

Bold Feynman Diagrams and the Luttinger–Ward Formalism via Gibbs Measures: Perturbative Approach

  • Published:
Archive for Rational Mechanics and Analysis Aims and scope Submit manuscript

Abstract

Many-body perturbation theory (MBPT) is widely used in quantum physics, chemistry, and materials science. At the heart of MBPT is the Feynman diagrammatic expansion, which is, simply speaking, an elegant way of organizing the combinatorially growing number of terms of a certain Taylor expansion. In particular, the construction of the ‘bold Feynman diagrammatic expansion’ involves the partial resummation to infinite order of possibly divergent series of diagrams. This procedure demands investigation from both the combinatorial (perturbative) and the analytical (non-perturbative) viewpoints. In this paper, we illustrate how the combinatorial properties of Feynman diagrams in MBPT can be studied in the simplified setting of Gibbs measures (known as the Euclidean lattice field theory in the physics literature) and provide a self-contained explanation of Feynman diagrams in this setting. We prove the combinatorial validity of the bold diagrammatic expansion, with methods generalizable to several types of field theories and interactions. Our treatment simplifies the presentation and numerical study of known techniques in MBPT such as the self-consistent Hartree–Fock approximation (HF), the second-order Green’s function approximation (GF2), and the GW approximation. The bold diagrams are closely related to the Luttinger-Ward (LW) formalism, which was proposed in 1960 but whose analytic properties have not been rigorously established. The analytical study of the LW formalism in the setting of Gibbs measures will be the topic of an accompanying paper.

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.

Institutional subscriptions

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
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27
Fig. 28
Fig. 29
Fig. 30

Similar content being viewed by others

Notes

  1. The Luttinger-Ward formalism is also known as the Kadanoff-Baym formalism [4] depending on the context. In this paper we always use the former.

  2. Intuitively speaking, these data specify a recipe for linking up half-edges to form a connected undirected graph of degree 4, but the previously specified data are a more natural representation of the diagram, especially once labels are introduced.

  3. So \(\{h_1,h_2\}\) is contained in the pairing \(\Pi _{\Gamma }\) of half-edges associated with \(\Gamma \).

  4. The proof of Proposition 4.4 fails for closed diagrams in that the case \(\vert E^{(j,k)} \vert =0 \) cannot be ruled out. (In the original proof, this case could be ruled out because it implied the existence of a closed subdiagram of a connected self-energy diagram, which is impossible.)

  5. For instance, if \(\alpha _1 = 1, \alpha _2 = 1, \alpha _3 = 2\), there are only three distinct multiset permutations: (1, 1, 2), (1, 2, 1), and (2, 1, 1).

  6. Interestingly, if one were to try to construct a unique skeleton decomposition for closed connected diagrams, that is, free energy diagrams, this is the place where the argument would fail; see Section 4.9.

  7. By such a ‘path’ we mean a sequence of half-edges \((h_1,h_2,\ldots ,h_{2m-1},h_{2m})\) in \(\Gamma ^{(k)}\) such that \(h_1 = h\); \(h_{2m} = h'\); \(h_l,h_{l+1}\) share an interaction line for l odd; and \(h_l,h_{l+1}\) are paired by \(\Gamma ^{(k)}\).

References

  1. Altland, A., Simons, B.D.: Condensed Matter Field Theory. Cambridge University Press, Cambridge 2010

    Book  Google Scholar 

  2. Amit, D.J., Martin-Mayor, V.: Field Theory, The Renormalization Group, and Critical Phenomena: Graphs to Computers. World Scientific Publishing Co Inc, Singapore 2005

    Book  Google Scholar 

  3. Aryasetiawan, F., Gunnarsson, O.: The GW method. Rep. Prog. Phys. 61, 237, 1998

    Article  ADS  Google Scholar 

  4. Baym, G., Kadanoff, L.P.: Conservation laws and correlation functions. Phys. Rev. 124, 287, 1961

    Article  ADS  MathSciNet  Google Scholar 

  5. Benlagra, A., Kim, K.-S., Pépin, C.: The Luttinger–Ward functional approach in the Eliashberg framework: a systematic derivation of scaling for thermodynamics near the quantum critical point. J. Phys. Condens. Matter 23, 145601, 2011

    Article  ADS  Google Scholar 

  6. Bernardi, M., Palummo, M., Grossman, J.C.: Extraordinary sunlight absorption and one nanometer thick photovoltaics using two-dimensional monolayer materials. Nano Lett. 13, 3664–3670, 2013

    Article  ADS  Google Scholar 

  7. Biermann, S., Aryasetiawan, F., Georges, A.: First-principles approach to the electronic structure of strongly correlated systems: combining the GW approximation and dynamical mean-field theory. Phys. Rev. Lett. 90, 086402, 2003

    Article  ADS  Google Scholar 

  8. Bulla, R., Costi, T.A., Pruschke, T.: Numerical renormalization group method for quantum impurity systems. Rev. Mod. Phys. 80, 395–450, 2008

    Article  ADS  Google Scholar 

  9. Dahlen, N.E., Van Leeuwen, R., Von Barth, U.: Variational energy functionals of the green function tested on molecules. Int. J. Quantum Chem. 101, 512–519, 2005

    Article  Google Scholar 

  10. Deligne, P., Etingof, P., Freed, D.S., Jeffrey, L.C., Kazhdan, D., Morgan, J.W., Morrison, D.R., Witten, E. (eds.): Quantum Fields and Strings: A Course for Mathematicians, 1, 2 edn. Amer Math Soc, Providence 1999

    Google Scholar 

  11. Elder, R.: Comment on “Non-existence of the Luttinger-Ward functional and misleading convergence of skeleton diagrammatic series for Hubbard-like models”. arXiv:1407.6599, 2014

  12. Fernández, R., Frölich, J., Sokal, A.: Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory. Springer, Berlin 1991

    Google Scholar 

  13. Fetter, A.L., Walecka, J.D.: Quantum Theory of Many-Particle Systems. Courier Corp, New York 2003

    MATH  Google Scholar 

  14. Georges, A., Kotliar, G., Krauth, W., Rozenberg, M.J.: Dynamical mean-field theory of strongly correlated fermion systems and the limit of infinite dimensions. Rev. Mod. Phys. 68, 13, 1996

    Article  ADS  MathSciNet  Google Scholar 

  15. Glimm, J., Jaffe, A.: Quantum Physics: A Functional Integral Point of View. Springer, Berlin 1987

    Book  Google Scholar 

  16. Gunnarsson, O., Rohringer, G., Schäfer, T., Sangiovanni, G., Toschi, A.: Breakdown of traditional many-body theories for correlated electrons. Phys. Rev. Lett. 119, 056402, 2017

    Article  ADS  Google Scholar 

  17. Hedin, L.: New method for calculating the one-particle Green’s function with application to the electron–gas problem. Phys. Rev. 139, A796, 1965

    Article  ADS  Google Scholar 

  18. Ismail-Beigi, S.: Correlation energy functional within the GW-RPA: exact forms, approximate forms, and challenges. Phys. Rev. B 81, 1–21, 2010

    Article  Google Scholar 

  19. Isserlis, L.: On a formula for the product-moment coefficient of any order of a normal frequency distribution in any number of variables. Biometrika 12, 134–139, 1918

    Article  Google Scholar 

  20. Kotliar, G., Savrasov, S.Y., Haule, K., Oudovenko, V.S., Parcollet, O., Marianetti, C.A.: Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 78, 865, 2006

    Article  ADS  Google Scholar 

  21. Kozik, E., Ferrero, M., Georges, A.: Nonexistence of the Luttinger–Ward functional and misleading convergence of skeleton diagrammatic series for hubbard-like models. Phys. Rev. Lett. 114, 156402, 2015

    Article  ADS  Google Scholar 

  22. Li, Y., Lu, J.: Bold diagrammatic Monte Carlo in the lens of stochastic iterative methods. arXiv:1710.00966, 2017

  23. Lin, L., Lindsey, M.: Variational structure of Luttinger–Ward formalism and bold diagrammatic expansion for Euclidean lattice field theory. Proc. Natl. Acad. Sci. 115, 2282, 2018

    Article  MathSciNet  Google Scholar 

  24. Luttinger, J.M., Ward, J.C.: Ground-state energy of a many-fermion system. II. Phys. Rev. 118, 1417, 1960

    Article  ADS  MathSciNet  Google Scholar 

  25. Martin, R.M., Reining, L., Ceperley, D.M.: Interacting Electrons. Cambridge University Press, Cambridge 2016

    Book  Google Scholar 

  26. Negele, J.W., Orland, H.: Quantum many-particle systems, Westview, 1988

  27. Onida, G., Reining, L., Rubio, A.: Electronic excitations: density-functional versus many-body Green’s-function approaches. Rev. Mod. Phys. 74, 601, 2002

    Article  ADS  Google Scholar 

  28. Peskin, M.E., Schroeder, D.V.: An introduction to quantum field theory, 1995

  29. Polyak, M.: Feynman diagrams for pedestrians and mathematicians. In: Graphs and Patterns in Mathematics and Theoretical Physics, pp. 1–27, 2004

  30. Potthoff, M.: Non-perturbative construction of the Luttinger–Ward functional. Condens. Matter Phys. 9, 557–567, 2006

    Article  Google Scholar 

  31. Prokof’ev, N.V., Svistunov, B.V.: Bold diagrammatic Monte Carlo: a generic sign-problem tolerant technique for polaron models and possibly interacting many-body problems. Phys. Rev. B 77, 125101, 2008

    Article  ADS  Google Scholar 

  32. Rentrop, J.F., Meden, V., Jakobs, S.G.: Renormalization group flow of the Luttinger–Ward functional: conserving approximations and application to the Anderson impurity model. Phys. Rev. B 93, 195160, 2016

    Article  ADS  Google Scholar 

  33. Salmhofer, M.: Renormalization: An Introduction. Springer, Berlin 2007

    MATH  Google Scholar 

  34. Staar, P., Maier, T., Schulthess, T.C.: Dynamical cluster approximation with continuous lattice self-energy. Phys. Rev. B 88, 115101, 2013

    Article  ADS  Google Scholar 

  35. Stefanucci, G., Van Leeuwen, R.: Nonequilibrium Many-Body Theory of Quantum Systems: A Modern Introduction. Cambridge University Press, Cambridge 2013

    Book  Google Scholar 

  36. Sun, Q., Chan, G.K.-L.: Quantum embedding theories. Acc. Chem. Res. 49, 2705–2712, 2016

    Article  Google Scholar 

  37. Tarantino, W., Romaniello, P., Berger, J.A., Reining, L.: Self-consistent Dyson equation and self-energy functionals: an analysis and illustration on the example of the Hubbard atom. Phys. Rev. B 96, 045124, 2017

    Article  ADS  Google Scholar 

  38. Zinn-Justin, J.: Quantum Field Theory and Critical Phenomena. Clarendon Press, Oxford 2002

    Book  Google Scholar 

Download references

Acknowledgements

This work was partially supported by the Department of Energy under grant DE-AC02-05CH11231 (L.L., M.L.), by the Department of Energy under grant No. DE-SC0017867 and by the Air Force Office of Scientific Research under award number FA9550-18-1-0095 (L.L.), and by the NSF Graduate Research Fellowship Program under Grant DGE-1106400 (M.L.). We thank Fabien Bruneval, Garnet Chan, Alexandre Chorin, Lek-Heng Lim, Nicolai Reshetikhin, Chao Yang and Lexing Ying for helpful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael Lindsey.

Additional information

Communicated by G. Friesecke.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A. Proof of Theorem 2.2 (Isserlis-Wick)

Proof

For any \(b\in {\mathbb {R}}^{N}\), define

$$\begin{aligned} Z_{0}(b) = \int _{{\mathbb {R}}^{N}} e^{-\frac{1}{2} x^{T} Ax + b^{T}x} \,\mathrm {d}x. \end{aligned}$$
(A.1)

Clearly \(Z_{0}=Z_{0}(0)\). Performing the change of variable \(\widetilde{x}=x+A^{-1}b\), we find

$$\begin{aligned} Z_{0}(b) = e^{\frac{1}{2} b^{T} A^{-1} b} Z_{0}. \end{aligned}$$
(A.2)

Observe that, for integers \(1\le \alpha _{1},\ldots ,\alpha _{m}\le N\),

$$\begin{aligned} \frac{\partial ^{m} Z_{0}(b)}{\partial b_{\alpha _{1}}\cdots \partial b_{\alpha _{m}}}\Big \vert _{b=0} = \int _{{\mathbb {R}}^{N}} x_{\alpha _{1}}\cdots x_{\alpha _{m}} e^{-\frac{1}{2} x^{T} Ax} \,\mathrm {d}x. \end{aligned}$$
(A.3)

Combining with Eq. (A.2), we have that

$$\begin{aligned} \left\langle x_{\alpha _{1}}\cdots x_{\alpha _{m}}\right\rangle _{0} = \frac{\partial ^{m} }{\partial b_{\alpha _{1}}\cdots \partial b_{\alpha _{m}}} e^{\frac{1}{2} b^{T} A^{-1} b} \Big \vert _{b=0}. \end{aligned}$$
(A.4)

Now write the Taylor expansion

$$\begin{aligned} e^{\frac{1}{2} b^{T} A^{-1} b} = \sum _{n=0}^{\infty } \frac{1}{n! \, 2^{n}} \left( b^{T}A^{-1} b \right) ^{n}. \end{aligned}$$
(A.5)

Since the expansion contains no odd powers of the components of b, the right-hand side of Eq. (A.4) vanishes whenever m is odd. If m is an even number, only the term \(n=m/2\) in the Taylor expansion will contribute to the right-hand side of Eq. (A.4). This gives

$$\begin{aligned} \frac{\partial ^{m} }{\partial b_{\alpha _{1}}\cdots \partial b_{\alpha _{m}}} e^{\frac{1}{2} b^{T} A b} \Big \vert _{b=0} = \frac{1}{2^{m/2} (m/2)!} \frac{\partial ^{m} }{\partial b_{\alpha _{1}}\cdots \partial b_{\alpha _{m}}} \left( b^{T}A^{-1} b \right) ^{m/2}. \end{aligned}$$
(A.6)

Now one can write

$$\begin{aligned} \frac{\partial ^{m} }{\partial b_{\alpha _{1}}\cdots \partial b_{\alpha _{m}}} = \frac{\partial ^{m} }{\partial ^{m_1} b_{\beta _{1}}\cdots \partial ^{m_p} b_{\beta _{p}}}, \end{aligned}$$

where the indices \(\beta _1,\ldots ,\beta _p\) are distinct and \(\sum _{j=1}^p m_j = m\). Then to further simplify the right-hand side of Eq. (A.6), we are interested in computing the coefficient of the \(b_{\beta _1}^{m_1} \cdots b_{\beta _p}^{m_p}\) term of the polynomial \((b^T A^{-1} b)^{m/2}\). Expand \((b^T A^{-1} b)^{m/2}\) into a sum of monomials as

$$\begin{aligned} (b^T A^{-1} b)^{m/2} = \sum _{j_1, k_1, \ldots , j_{m/2}, k_{m/2} } \prod _{i=1}^{m/2} (A^{-1})_{j_i, k_i} b_{j_i} b_{k_i}. \end{aligned}$$

Thus each distinct permutation \((j_1, k_1, \ldots , j_{m/2}, k_{m/2})\) of the multiset \(\{\alpha _1, \ldots ,\alpha _m \}\)Footnote 5 contributes \(\prod _{i=1}^{m/2} (A^{-1})_{j_i, k_i}\) to our desired coefficient.

Since

$$\begin{aligned} \frac{\partial ^{m} }{\partial ^{m_1} b_{\beta _{1}}\cdots \partial ^{m_p} b_{\beta _{p}}} b_{\beta _1}^{m_1} \cdots b_{\beta _p}^{m_p} = m_1! \cdots m_p!, \end{aligned}$$

we obtain, from Eq. (A.4) and Eq. (A.6), that

$$\begin{aligned} \left\langle x_{\alpha _{1}}\cdots x_{\alpha _{m}}\right\rangle _{0} = \frac{m_1! \cdots m_p!}{2^{m/2} (m/2)!} \sum _{(j_1, k_1, \ldots , j_{m/2}, k_{m/2})} \prod _{i=1}^{m/2} (A^{-1})_{j_i, k_i}, \end{aligned}$$
(A.7)

where the sum is understood to be taken over multiset permutations of \(\{\alpha _1, \ldots ,\alpha _m \}\).

Now every permutation of \(\mathcal {I}_m\) can be thought of as inducing a permutation of the multiset \(\{\alpha _1, \ldots ,\alpha _m \}\) via the map \((l_1, \ldots , l_m) \mapsto (\alpha _{l_1}, \ldots , \alpha _{l_m})\). By this map each multiset permutation of \(\{\alpha _1, \ldots ,\alpha _m \}\) is associated with \(m_1! \cdots m_p!\) different permutations of \(\mathcal {I}_m\).

Moreover, every permutation of \(\mathcal {I}_m\) can be thought of as inducing a pairing on \(\mathcal {I}_m\) by pairing the first two indices in the permutation, then the next two, etc. For example, if \(m=4\), then the permutation (1, 4, 3, 2) induces the pairing (1, 4)(3, 2). By this map, each pairing on \(\mathcal {I}_m\) is associated with \(2^{m/2} (m/2)!\) permutations of \(\mathcal {I}_m\) since a pairing does not distinguish the ordering of the pairs, nor the order of the indices within each pair.

Finally, observe that if we take any two permutations of \(\mathcal {I}_m\) associated with a pairing on \(\mathcal {I}_m\) and then consider the (possibly distinct) multiset permutations \((j_1, k_1, \ldots , j_{m/2}, k_{m/2})\) and \((j_1', k_1', \ldots , j_{m/2}', k_{m/2}')\) of \(\{\alpha _1, \ldots ,\alpha _m \}\) associated to these permutations, then in fact \(\prod _{i=1}^{m/2} (A^{-1})_{j_i, k_i} = \prod _{i=1}^{m/2} (A^{-1})_{j_i', k_i'}\).

Thus we can replace the sum over multisets in Eq. (A.7) with a sum over pairings on \(\mathcal {I}_m\). Each term must be counted \(\frac{2^{m/2} (m/2)!}{m_1! \cdots m_p!}\) times because each pairing on \(\mathcal {I}_m\) induces \(2^{m/2} (m/2)!\) permutations of \(\mathcal {I}_m\), each of which is redundant by a factor of \(m_1! \cdots m_p!\). This factor cancels with the factor in Eq. (A.7) to yield

$$\begin{aligned} \left\langle x_{\alpha _{1}}\cdots x_{\alpha _{m}}\right\rangle _{0} = \sum _{\sigma \in \Pi (\mathcal {I}_{m})} \prod _{i \in \mathcal {I}_{m}/\sigma } (A^{-1})_{\alpha _{i},\alpha _{\sigma (i)} }. \end{aligned}$$

Recalling that \(G^{0}=A^{-1}\), this completes the proof of Theorem 2.2. \(\quad \square \)

Remark A.1

In field theories, the auxiliary variable b is often interpreted as a coupling external field.

Appendix B. Proof of Proposition 4.4 (Skeleton decomposition)

Proof

As suggested in the statement of the proposition, let \(\Gamma ^{(k)}\) be the maximal insertions admitted by \(\Gamma \), where we do not count separately any insertions that differ only by their external labeling (see Remark 4.5). Let \(h_1^{(k)}\) and \(h_2^{(k)}\) be half edges such that \(\Gamma \) admits the insertion \(\Gamma ^{(k)}\) at \((h_1^{(k)},h_2^{(k)})\). Let \(e_1^{(k)},e_2^{(k)}\) be the external half-edges of the truncated Green’s function diagram \(\Gamma ^{(k)}\) paired with \(h_1^{(k)},h_2^{(k)}\), respectively, in the overall diagram \(\Gamma \).

First we aim to show that the \(\Gamma ^{(k)}\) are disjoint, that is, share no half-edges. In this case we say that the \(\Gamma ^{(k)}\) do not overlap. In fact we will show additionally that the external half-edges of the \(\Gamma ^{(k)}\) do not touch one another (that is, are not paired in \(\Gamma \)), and accordingly we say that the \(\Gamma ^{(k)}\) do not touch.

To this end, let \(j \ne k\). The idea is to consider the diagrammatic structure formed by collecting all of the half-edges appearing in \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) and then arguing by maximality that \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) cannot overlap or touch, let this structure admit \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) as insertions.

We now formalize this notion. Let \(H^{(j)}\) and \(H^{(k)}\) be the half-edge sets of \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\), respectively, and consider the union \(H^{(j,k)} := H^{(j)} \cup H^{(k)}\), together with a partial pairing \(\Pi ^{(j,k)}\) on \(H^{(j,k)}\), that is, a collection of disjoint pairs in \(H^{(j,k)}\), defined by the rule that \(\{h_1,h_2\} \in \Pi ^{(j,k)}\) if and only if \(\{h_1,h_2\} \in \Pi _{\Gamma }\) and \(h_1,h_2 \in H^{(j,k)}\). In other words, the structure \((H^{(j,k)},\Pi ^{(j,k)})\) is formed by taking all of the half-edges appearing in \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) and then pairing the ones that are paired in the overall diagram \(\Gamma \).

Let \(E^{(j,k)}\) be the subset of unpaired half-edges in \(H^{(j,k)}\), that is, those half-edges that do not appear in any pair in \(\Pi ^{(j,k)}\). Since all half-edges in \(\Gamma ^{(k)}\) besides \(e_1^{(k)},e_2^{(k)}\) are paired in the diagram \(\Gamma ^{(k)}\), we must have \(E^{(j,k)} \subset \{ e_1^{(j)},e_2^{(j)}, e_1^{(k)},e_2^{(k)} \}\). \(\quad \square \)

Lemma B.1

\(\vert E^{(j,k)} \vert = 4\).

Proof

We claim that \(\vert E^{(j,k)} \vert \) is even. To see this, we first establish that \(\vert H^{(j,k)} \vert \) is even. Notice that a truncated Green’s function diagram (in particular, \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\)) contains an even number of half-edges (more specifically 4n, where n is the order of the diagram). Thus \(\vert H^{(j)} \vert , \vert H^{(k)}\vert \) are even. Moreover, \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) share a half-edge if and only if they share the vertex (or interaction line) associated with this half-edge, in which case \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) share all four half-edges emanating from this vertex. Thus the number \(\vert H^{(j)} \cap H^{(k)} \vert \) of half-edges common to \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) is four times the number of common interaction lines, in particular an even number. So \(\vert H^{(j,k)} \vert = \vert H^{(j)} \vert + \vert H^{(k)} \vert - \vert H^{(j)} \cap H^{(k)} \vert \) is even, as desired. Now any partial pairing on \(H^{(j,k)}\) includes an even number of distinct elements, so the number of leftover elements, that is, \(\vert E^{(j,k)} \vert \), must be even as well, as claimed.

Next we rule out the cases \(\vert E^{(j,k)} \vert \in \{ 0, 2 \}\).

Suppose that \(\vert E^{(j,k)} \vert = 0\). Then the structure \((H^{(j,k)},\Pi ^{(j,k)})\) defines a closed sub-diagram of \(\Gamma \), disconnected from the rest of \(\Gamma \). Since \(\Gamma \) is not closed, the sub-diagram cannot be all of \(\Gamma \). This conclusion contradicts the connectedness of \(\Gamma \).Footnote 6

Next suppose that \(\vert E^{(j,k)} \vert = 2\). Then the structure \((H^{(j,k)},\Pi ^{(j,k)})\) has two unpaired half-edges, hence defines a truncated Green’s function diagram \(\Gamma ^{(j,k)}\) that contains both \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) and admits both as insertions. But since \(\Gamma ^{(j)} \ne \Gamma ^{(k)}\), \(\Gamma ^{(j,k)}\) is neither \(\Gamma ^{(j)}\) nor \(\Gamma ^{(k)}\) (that is, the containment is proper). This conclusion contradicts the maximality of \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\).

From Lemma B.1 it follows that \(E^{(j,k)} = \{ e_1^{(j)},e_2^{(j)}, e_1^{(k)},e_2^{(k)} \}\). (And moreover \( e_1^{(j)},e_2^{(j)}, e_1^{(k)},e_2^{(k)}\) are distinct, which was not assumed!) This establishes one of our original claims, that is, that the external half-edges of \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) do not touch (that is, are not paired.)

We need to establish the other part of our original claim, that is, that the half-edge sets \(H^{(j)}\) and \(H^{(k)}\) are disjoint. To see this, suppose otherwise, so \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}\) share some half-edge h. Since \(\Gamma ^{(j)} \ne \Gamma ^{(k)}\), one of \(H^{(j)}\) and \(H^{(k)}\) does not contain the other, so assume without loss of generality that \(H^{(j)}\) does not contain \(H^{(k)}\), so there is some half-edge \(h' \in H^{(k)} \backslash H^{(j)}\). Since \(\Gamma ^{(k)}\) is connected there must be some path in \(\Gamma ^{(k)}\) connecting h with \(h'\).Footnote 7 Now \(\Gamma ^{(j)}\) can be disconnected from the rest of \(\Gamma \) by deleting the links \(\{e_1^{(j)},h_1^{(j)}\}\) and \(\{e_2^{(j)},h_2^{(j)}\}\) from the pairing \(\Pi _{\Gamma }\), so evidently our path in \(\Gamma ^{(k)}\) connecting h and \(h'\) must contain either \((e_1^{(j)},h_1^{(j)})\) or \((e_2^{(j)},h_2^{(j)})\) as a ‘subpath’. But then either \(e_1^{(j)}\) or \(e_2^{(j)}\) is paired in \(\Gamma ^{(k)}\), hence also by \(\Pi ^{(j,k)}\), contradicting its inclusion in \(E^{(j,k)}\).

In summary we have shown that the \(\Gamma ^{(k)}\) are disjoint, that is, share no half-edges, and that the external half-edges of the \(\Gamma ^{(k)}\) do not touch one another, that is, are not paired in \(\Gamma \). We can define a truncated Green’s function diagram \(\Gamma _{\mathrm {s}}\) by considering \(\Gamma \), then omitting all half-edges and vertices appearing in the \(\Gamma ^{(k)}\). Since the \(\Gamma ^{(k)}\) are disjoint and do not touch, this leaves behind the half-edges \(h_1^{(k)},h_2^{(k)}\) for all k, which are now left unpaired. Then we complete the construction of \(\Gamma _{\mathrm {s}}\) by adding the pairings \(\{h_1^{(k)},h_2^{(k)}\}\). In short, we form \(\Gamma _{\mathrm {s}}\) from \(\Gamma \) by replacing each insertion \(\Gamma ^{(k)}\) with the edge \(\{h_1^{(k)},h_2^{(k)}\}\). Eq. (4.2) then holds by construction, for a suitable external labeling of the \(\Gamma ^{(k)}\).

Moreover, we find that \(\Gamma _{\mathrm {s}}\) is 2PI. It is not hard to check first that \(\Gamma _{\mathrm {s}}\) is 1PI. Indeed, the unlinking of any half-edge pair in \(\Gamma _{\mathrm {s}}\) that is not one of the \(\{h_1^{(k)},h_2^{(k)}\}\) can be lifted to the unlinking of the same half-edge pair in the original diagram \(\Gamma \). Since \(\Gamma \) is 1PI, this unlinking does not disconnect \(\Gamma \). Re-collapsing the maximal insertions once again does not affect the connectedness of the result, so \(\Gamma _{\mathrm {s}}\) does not become disconnected by the unlinking. On the other hand, the unlinking of a half-edge pair \(\{h_1^{(k)},h_2^{(k)}\}\) were to disconnect \(\Gamma _{\mathrm {s}}\), then necessarily the unlinking of either \(\{e_1^{(k)},h_1^{(k)}\}\) or \(\{e_2^{(k)},h_2^{(k)}\}\) would disconnect \(\Gamma \), which contradicts the premise that \(\Gamma \) is 1PI.

Thus \(\Gamma _{\mathrm {s}}\) is 1PI, and two-particle irreducibility is equivalent to the absence of any insertions. But if \(\Gamma _{\mathrm {s}}\) admits an insertion containing either \(h_1^{(k)}\) or \(h_2^{(k)}\), then this contradicts the maximality of the insertion \(\Gamma ^{(k)}\) in \(\Gamma \). Moreover, if \(\Gamma _{\mathrm {s}}\) admits an insertion containing none of the \(h_1^{(k)},h_2^{(k)}\), then this insertion lifts to an insertion in the original diagram \(\Gamma \), hence this insertion (that is, all of its interaction lines and half-edges) must have been omitted from \(\Gamma _{\mathrm {s}}\) (contradiction). We conclude that \(\Gamma _{\mathrm {s}}\) admits no insertions, hence is 2PI.

Finally it remains to prove the uniqueness of the decomposition of Eq. (4.2). To this end, let \(\Gamma _{\mathrm {s}} \in \mathfrak {F}_2^{\mathrm {2PI}}\) and \(\Gamma ^{(k)} \in \mathfrak {F}_2^{\mathrm {c,t}}\) for \(k=1,\ldots ,K\), and let \(\left\{ h_1^{(k)},h_2^{(k)}\right\} \) be distinct half-edge pairs in \(\Gamma _{\mathrm {s}}\) for \(k=1,\ldots ,K\). Then define \(\Gamma \) via Eq. (4.2). The uniqueness claim then follows if we can show that the \(\Gamma ^{(k)}\) are the maximal insertions in \(\Gamma \).

Suppose that \(\Gamma ^{(k)}\) is not maximal for some k. Then by definition the diagram \(\Gamma '\) formed from \(\Gamma \) by collapsing the insertion \(\Gamma ^{(k)}\) admits an insertion containing \(h_1^{(k)}\) or \(h_2^{(k)}\) (assume \(h_1^{(k)}\) without loss of generality). In fact let \(\Gamma ^{(k)}_{\mathrm {m}}\) be a maximal insertion containing \(h_1^{(k)}\). Note also that \(\Gamma '\) still admits the \(\Gamma ^{(j)}\) for \(j\ne k\) as insertions.

Then for \(j\ne k\) form the structure \((H^{(j,k)},\Pi ^{(j,k)})\) by ‘merging’ \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}_{\mathrm {m}}\) via the same construction as above (now within the overall diagram \(\Gamma '\)). By the same reasoning as in Lemma B.1, the set of unpaired half-edges \(E^{(j,k)}\) must be of even cardinality, and we can rule out the case \(\vert E^{(j,k)} \vert =0\).

In the case \(\vert E^{(j,k)} \vert = 2\), the structure \((H^{(j,k)},\Pi ^{(j,k)})\) once again defines a truncated Green’s function diagram \(\Gamma ^{(j,k)}\) admitting both \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}_{\mathrm {m}}\) as insertions. But since \(\Gamma ^{(k)}_{\mathrm {m}}\) is maximal, \(\Gamma ^{(j,k)} = \Gamma ^{(k)}_{\mathrm {m}}\), and \(\Gamma ^{(j)}\) is contained in \(\Gamma ^{(k)}_{\mathrm {m}}\).

In the case \(\vert E^{(j,k)} \vert = 4\), since \(\Gamma ^{(j)}\) does not contain \(\Gamma ^{(k)}_{\mathrm {m}}\) (that is, the half-edge set of the former does not contain that of the latter), the same reasoning as above guarantees that \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}_{\mathrm {m}}\) do not overlap or touch.

Then consider the diagram \(\Gamma ''\) formed from \(\Gamma '\) by collapsing the insertion \(\Gamma ^{(j)}\). In both of our cases (namely, that \(\Gamma ^{(j)}\) is contained in \(\Gamma ^{(k)}_{\mathrm {m}}\) and that \(\Gamma ^{(j)}\) and \(\Gamma ^{(k)}_{\mathrm {m}}\) do not overlap or touch), the insertion \(\Gamma ^{(k)}_{\mathrm {m}}\) descends to an insertion in \(\Gamma ''\) containing \(h_1^{(k)}\).

Iteratively repeating this procedure for all \(j\ne k\) (that is, collapsing all of the insertions \(\Gamma ^{(j)}\) to obtain the original skeleton diagram \(\Gamma _{\mathrm {s}}\)), we find that \(\Gamma ^{(k)}_{\mathrm {m}}\) descends to an insertion in \(\Gamma _{\mathrm {s}}\) containing \(h_1^{(k)}\), contradicting the fact that \(\Gamma _{\mathrm {s}}\) is 2PI.

\(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lin, L., Lindsey, M. Bold Feynman Diagrams and the Luttinger–Ward Formalism via Gibbs Measures: Perturbative Approach. Arch Rational Mech Anal 242, 581–642 (2021). https://doi.org/10.1007/s00205-021-01692-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00205-021-01692-x

Navigation