Skip to main content
Log in

Long-Range Charge Order in the Extended Holstein–Hubbard Model

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

This study investigated the extended Holstein–Hubbard model at half-filling as a model for describing the interplay of electron–electron and electron–phonon couplings. When the electron–phonon and nearest-neighbor electron–electron interactions are strong, we prove the existence of long-range charge order in three or more dimensions at a sufficiently low temperature. As a result, we rigorously justify the phase competition between the antiferromagnetism and charge orders.

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.

Similar content being viewed by others

Notes

  1. Namely, he applied the spin reflection positivity to the Holstein–Hubbard and Su–Schrieffer–Heeger models and investigated their ground state properties.

  2. In this paper, we simply assume that the right hand side of (2.6) exists. Alternatively, we choose a subsequence such that the right hand side of (2.6) exists.

  3. To be precise, \(\mathfrak {H}_M\) is defined by

    where \(N=N_{\uparrow }+N_{\downarrow }\) and \(S^3=\frac{1}{2}(N_{\uparrow }-N_{\downarrow })\) with \(N_{\sigma }=\sum _{x\in \Lambda } n_{x\sigma }\). The condition \(N\psi =|\Lambda |\psi \) indicates that we consider the case of half-filling.

  4. Namely, \(\vartheta \) is a bijective antilinear map which satisfies \( \langle \vartheta \varphi |\vartheta \psi \rangle =(\langle \varphi |\psi \rangle )^* \) for all \(\varphi , \psi \in \mathfrak {X}_L.\)

  5. In the Schrödinger representation, \(\Omega _{\mathrm {b}}^L=(\frac{1}{\pi })^{|\Lambda _L|/4} e^{-\sum _{x\in \Lambda _L}\phi _x^2/2} \). \(\Omega _{\mathrm {f}}^L\) is the standard Fock vacuum in \(\mathfrak {F}_L\). Note that \(\Omega _{\mathrm {b}}^L\) is a real-valued function on \(\mathcal {Q}_{\Lambda _L}\).

References

  1. Bardeen, J., Cooper, L.N., Schrieffer, J.R.: Theory of superconductivity. Phys. Rev. 108, 1175–1204 (1957)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  2. Bari, R.A.: Effects of short-range interactions on electron-charge ordering and lattice distortions in the localized state. Phys. Rev. B 3, 2662–2670 (1971)

    Article  ADS  Google Scholar 

  3. Capone, M., Fabrizio, M., Castellani, C., Tosatti, E.: Colloquium: modeling the unconventional superconducting properties of expanded \({\rm A}_{3} C_{ 60} \) fullerides. Rev. Mod. Phys. 81, 943–958 (2009)

    Article  ADS  Google Scholar 

  4. Dyson, F.J., Lieb, E.H., Simon, B.: Phase transitions in quantum spin systems with isotropic and nonisotropic interactions. J. Stat. Phys. 18, 335–383 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  5. Falk, H., Bruch, L.W.: Susceptibility and fluctuation. Phys. Rev. 180, 442–444 (1969)

    Article  ADS  Google Scholar 

  6. Freericks, J.K., Lieb, E.H.: Ground state of a general electron–phonon Hamiltonian is a spin singlet. Phys. Rev. B 51, 2812–2821 (1995)

    Article  ADS  Google Scholar 

  7. Fröhlich, J., Israel, R., Lieb, E.H., Simon, B.: Phase transitions and reflection positivity. I. General theory and long range lattice models. Commun. Math. Phys. 62, 1–34 (1978)

    Article  ADS  MathSciNet  Google Scholar 

  8. Fröhlich, J., Israel, R., Lieb, E.H., Simon, B.: Phase transitions and reflection positivity. II. Lattice systems with short-range and Coulomb interactions. J. Stat. Phys. 22, 297–347 (1980)

    Article  ADS  MathSciNet  Google Scholar 

  9. Fröhlich, J., Simon, B., Spencer, T.: Infrared bounds, phase transitions and continuous symmetry breaking. Commun. Math. Phys. 50, 79–95 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  10. Glimm, J., Jaffe, A., Spencer, T.: Phase transitions for \(\varphi ^4_2\) quantum fields. Commun. Math. Phys. 45, 203–216 (1975)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Glimm, J., Jaffe, A.: Quantum physics. A functional integral point of view, 2nd edn. Springer, New York (1987)

    MATH  Google Scholar 

  12. Giuliani, A., Mastropietro, V., Porta, M.: Universality of conductivity in interacting graphene. Commun. Math. Phys. 311, 317–355 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Jedrzejewski, J.: Electron charge ordering in the extended Hubbard model. Z. Phys. B Conden. Matter 48, 219–225 (1982)

    Article  ADS  Google Scholar 

  14. Kosugi, T., Miyake, T., Ishibashi, S., Arita, R., Aoki, H.: First-principles structural optimization and electronic structure of the superconductor picene for various potassium doping levels. Phys. Rev. B 84, 214506 (2011)

    Article  ADS  Google Scholar 

  15. Kubo, K., Kishi, T.: Rigorous bounds on the susceptibilities of the Hubbard model. Phys. Rev. B 41, 4866–4868 (1990)

    Article  ADS  Google Scholar 

  16. Kubozono, Y., Mitamura, H., Lee, X., He, X., Yamanari, Y., Takahashi, Y., Suzuki, Y., Kaji, Y., Eguchi, R., Akaike, K., Kambe, T., Okamoto, H., Fujiwara, A., Kato, T., Kosugi, T., Aoki, H.: Metal-intercalated aromatic hydrocarbons: a new class of carbon-based superconductors. Phys. Chem. Chem. Phys. 13, 16476–16493 (2011)

    Article  Google Scholar 

  17. Lang, I.G., Firsov, Y.A.: Kinetic theory of semiconductors with low mobility. Sov. Phys. JETP 16, 1301 (1963)

    ADS  MATH  Google Scholar 

  18. Lanzara, A., Bogdanov, P.V., Zhou, X.J., Kellar, S.A., Feng, D.L., Lu, E.D., Yoshida, T., Eisaki, H., Fujimori, A., Kishio, K., Shimoyama, J.-I., Noda, T., Uchida, S., Hussain, Z., Shen, Z.-X.: Evidence for ubiquitous strong electron–phonon coupling in high-temperature superconductors. Nature 412, 510–514 (2001)

    Article  ADS  Google Scholar 

  19. Lieb, E.H.: Two theorems on the Hubbard model. Phys. Rev. Lett. 62, 1201–1204 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  20. Lieb, E.H.: Flux phase of the half-filled band. Phys. Rev. Lett. 73, 2158–2161 (1994)

    Article  ADS  Google Scholar 

  21. Lieb, E.H., Loss, M.: Fluxes, Laplacians, and Kasteleyn’s theorem. Duke Math. J. 71, 337–363 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  22. Lieb, E.H., Nachtergaele, B.: Stability of the Peierls instability for ring-shaped molecules. Phys. Rev. B 51, 4777–4791 (1995)

    Article  ADS  MATH  Google Scholar 

  23. Löwen, H.: Absence of phase transitions in Holstein systems. Phys. Rev. B 37, 8661–8667 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  24. Macris, M., Nachtergaele, B.: On the flux phase conjecture at half-filling: an improved proof. J. Stat. Phys. 85, 745–761 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Miyao, T.: Self-dual cone analysis in condensed matter physics. Rev. Math. Phys. 23, 749–822 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  26. Miyao, T.: Ground state properties of the SSH model. J. Stat. Phys. 149, 519–550 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Miyao, T.: Rigorous results concerning the Holstein–Hubbard model. In: Annales Henri Poincare. arXiv:1402.5202 (in press)

  28. Miyao, T.: Upper bounds on the charge susceptibility of many-electron systems coupled to the quantized radiation field. Lett. Math. Phys. 105, 1119–1133 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Murakami, Y., Werner, P., Tsuji, N., Aoki, H.: Ordered phases in the Holstein-Hubbard model: interplay of strong Coulomb interaction and electron–phonon coupling. Phys. Rev. B 88, 125126 (2013)

    Article  ADS  Google Scholar 

  30. Nomura, Y., Nakamura, K., Arita, R.: Ab initio derivation of electronic low-energy models for \({{\rm C}_{60}}\) and aromatic compounds. Phys. Rev. B 85, 155452 (2012)

    Article  ADS  Google Scholar 

  31. Osterwalder, K., Schrader, R.: Axioms for Euclidean Green’s functions. Comm. Math. Phys. 31 (1973), 83–112. Axioms for Euclidean Green’s functions. II. With an appendix by Stephen Summers. Commun. Math. Phys. 42, 281–305 (1975)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Simon, B.:The Statistical Mechanics of Lattice Gases, Volume I. Princeton Univ Press, Princeton (1993)

  33. Takabayashi, Y., Ganin, A.Y., Jegli, P., Aron, D., Takano, T., Iwasa, Y., Ohishi, Y., Takata, M., Takeshita, N., Prassides, K., Rosseinsky, M.J.: The disorder-free Non-BCS superconductor \({{\rm Cs}_3}{ {\rm C}_{60}}\) emerges from an antiferromagnetic insulator parent state. Science 323, 1585–1590 (2009)

    Article  ADS  Google Scholar 

Download references

Acknowledgments

This work was partially supported by Japan Society for the Promotion of Science (KAKENHI 20554421, KAKENHI 16H03942). I would be grateful to the anonymous referees for useful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tadahiro Miyao.

Appendices

Appendix 1: Proof of (2.5)

We will show that \(\langle q_x\rangle _{\beta , \Lambda }=0\). The hole-particle transformation is a unitary operator u such that

$$\begin{aligned} u c_{x\uparrow } u^{-1} = c_{x\uparrow },\ \ \ u c_{x\downarrow } u^{-1} =(-1)^{\Vert x\Vert }c_{x\downarrow }^*. \end{aligned}$$
(4.1)

We set \(\mathbb {H}=u H_{\Lambda } u^{-1}\). Since \(uq_xu^{-1}=s_x:=n_{x\uparrow }-n_{x\downarrow }\), we obtain \(\mathbb {H}=\mathbb {H}_0+\mathbb {W}\), where

$$\begin{aligned} \mathbb {H}_0&=T+K,\end{aligned}$$
(4.2)
$$\begin{aligned} \mathbb {W}&= U\sum _{x\in \Lambda } s_x^2+V\sum _{\langle x, y\rangle }s_xs_y+g\sum _{x\in \Lambda } s_x(b_x+b_x^*). \end{aligned}$$
(4.3)

Here, T and K are defined by (3.1) and (3.4), respectively. Thus, we have

$$\begin{aligned} \langle q_x\rangle =\langle \!\langle s_x\rangle \!\rangle , \end{aligned}$$
(4.4)

where \(\langle \!\langle \cdot \rangle \!\rangle \) is the thermal expectation associated with \(\mathbb {H}\).

Let D be a unitary operator such that

$$\begin{aligned} Dc_{x\uparrow } D^{-1} =c_{x\downarrow },\ \ D c_{x\downarrow } D^{-1}=c_{x\uparrow }, \ \ Db_x D^{-1}=-b_x. \end{aligned}$$
(4.5)

Since \(D\mathbb {H} D^{-1}=\mathbb {H}\) and \(Ds_x D^{-1} =-s_x\), we have \(\langle q_x\rangle =\langle \!\langle s_x\rangle \!\rangle =0\). This concludes the proof of (2.5). \(\square \)

Appendix 2: The Dyson–Lieb–Simon Inequality

Let \(\mathfrak {X}_L\) and \(\mathfrak {X}_R\) be complex Hilbert spaces and let \(\vartheta \) be an antiunitary transformation from \(\mathfrak {X}_L\) onto \(\mathfrak {X}_R\). Let \(A, B, C_j, D_j, j=1,\dots , n\) be linear operators in \(\mathfrak {X}_L\). Suppose that A and B are self-adjoint and bounded from below and that \(C_j\) and \(D_j\) are bounded. We will study the following Hamiltonian:

$$\begin{aligned} H(A, B, \mathbf {C}, \mathbf {D})&=H_0-V,\end{aligned}$$
(4.6)
$$\begin{aligned} H_0&=A\otimes \mathbbm {1}+\mathbbm {1}\otimes \vartheta B \vartheta ^{-1},\end{aligned}$$
(4.7)
$$\begin{aligned} V&= \sum _{j=1}^n \lambda _j\left( C_j\otimes \vartheta D_j\vartheta ^{-1}+C_j^*\otimes \vartheta D_j^* \vartheta ^{-1}\right) . \end{aligned}$$
(4.8)

\(H(A, B, \mathbf {C}, \mathbf {D})\) is a self-adjoint operator bounded from below acting in \(\mathfrak {X}_L\otimes \mathfrak {X}_R\).

Theorem 3.23

Assume that \(e^{-\beta A}\) and \(e^{-\beta B}\) are trace class operators for all \(\beta >0\) and that \(\lambda _j\ge 0\) for all \(j\in \{1, \dots , n\}\). Let \( Z_{\beta }(A, B, \mathbf {C}, \mathbf {D}) =\mathrm {Tr} \big [ e^{-\beta H(A, B, \mathbf {C}, \mathbf {D})} \big ],\ \beta >0 \). We then have

$$\begin{aligned} Z_{\beta }(A, B, \mathbf {C}, \mathbf {D})^2 \le Z_{\beta }(A, A, \mathbf {C}, \mathbf {C})Z_{\beta }(B, B, \mathbf {D}, \mathbf {D}). \end{aligned}$$
(4.9)

Remark 3.24

  1. (i)

    In [4], all matrix elements of \(A, B, C_j, D_j\) are assumed to be real. However, as noted in [22, 25], this assumption is unnecessary. This point is essential for the present paper.

  2. (ii)

    Suppose that \(\dim \mathfrak {X}_L<\infty \). Therefore, we set \(\mathfrak {X}_L=\mathfrak {X}_R=\mathbb {C}^n\). Let \(\vartheta \) be the standard conjugation: \(\vartheta \psi =\{\overline{\psi _j}\}_{j=1}^n\) for each \( \psi \in \mathfrak {X}_L\). Hence, \(\vartheta B\vartheta ^{-1}\) represents the complex conjugation of the matrix elements of B. Now assume the following: (a) \(C_j=D_j\) for all j; (b) \(C_j\) is self-adjoint for all j (\(C_j^*=C_j\)); (c) \(C_j\) is real for all j (\(\vartheta C_j \vartheta ^{-1}=C_j\)). In this case, we obtain a finite temperature version of [22, Lemma 14]. \(\diamondsuit \)

Proof

While this theorem is proven in [25], we present the proof here for reader’ convenience. It suffices to show the assertion when \(\dim \mathfrak {X}_L<\infty \).

The following property is fundamental:

$$\begin{aligned} \mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R}\Big [ A\otimes \vartheta B\vartheta ^{-1} \Big ]= \mathrm {Tr}_{\mathfrak {X}_L}[A] \big ( \mathrm {Tr}_{\mathfrak {X}_L}[B] \big )^*. \end{aligned}$$
(4.10)

Especially, we have \( \mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R}\Big [ A\otimes \vartheta A\vartheta ^{-1} \Big ]= \big |\mathrm {Tr}_{\mathfrak {X}_L}[A]\big |^2\ge 0 \). The reason for this is as follows. Since \( \mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R}\Big [ A\otimes \vartheta B\vartheta ^{-1} \Big ]= \mathrm {Tr}_{\mathfrak {X}_L}[A] \mathrm {Tr}_{\mathfrak {X}_R}[\vartheta B \vartheta ^{-1}] \), it suffices to show that \( \mathrm {Tr}_{\mathfrak {X}_R}[\vartheta B \vartheta ^{-1}] =(\mathrm {Tr}_{\mathfrak {X}_L}[ B ])^* \). Let \(\{e_n\}_{n=1}^{\infty }\) be a complete orthonormal system (CONS) of \(\mathfrak {X}_R\). Remarking that \(\{\vartheta ^{-1} e_n\}_{n=1}^{\infty }\) is a CONS of \(\mathfrak {X}_L\) and since \(\langle \vartheta ^{-1} \phi |\vartheta ^{-1} \psi \rangle =\langle \psi |\phi \rangle \), we see that

$$\begin{aligned} \mathrm {Tr}_{\mathfrak {X}_R}[\vartheta B \vartheta ^{-1}]&=\sum _{n=1}^N \langle e_n|\vartheta B\vartheta ^{-1}e_n\rangle =\sum _{n=1}^N \langle \vartheta ^{-1}\vartheta B\vartheta ^{-1}e_n|\vartheta ^{-1}e_n\rangle \nonumber \\&=\sum _{n=1}^N \langle B\vartheta ^{-1} e_n|\vartheta ^{-1} e_n\rangle =(\mathrm {Tr}_{\mathfrak {X}_L}[B])^*. \end{aligned}$$
(4.11)

As a first step, we will prove the assertion by assuming that \(C_j\) and \(D_j\) are self-adjoint. For simplicity, assume that \(\lambda _j=1/2\). By the Duhamel formula,

$$\begin{aligned} \mathrm {e}^{-\beta H(A, B; \mathbf {C}, \mathbf {D})}&=\sum _{N\ge 0}\mathscr {D}_{N, \beta }(A, B; \mathbf {C}, \mathbf {D}),\end{aligned}$$
(4.12)
$$\begin{aligned} \mathscr {D}_{N, \beta }(A, B; \mathbf {C}, \mathbf {D})&=\int _{S_N(\beta )}\ e^{-t_1 H_0}V e^{-t_2 H_0}\ldots e^{-t_N H_0}V e^{-(\beta -\sum _{j=1}^N t_j) H_0}, \end{aligned}$$
(4.13)

where \(\int _{S_N(\beta )}=\int _0^{\beta }\mathrm {d}t_1\int _0^{\beta -t_1}\mathrm {d}t_2\ldots \int _0^{\beta -\sum _{j=1}^{N-1}t_j} \mathrm {d}t_N \). Observe that

$$\begin{aligned}&\mathscr {D}_{N, \beta }(A, B; \mathbf {C}, \mathbf {D}) \nonumber \\ =&\sum _{k_1, \dots , k_N\ge 1} \int _{S_N(\beta )}\ \Big [\mathscr {L}_{A; \mathbf {C}}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big ) \Big ] \otimes \vartheta \Big [\mathscr {L}_{B; \mathbf {D}}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big ) \Big ] \vartheta ^{-1}, \end{aligned}$$
(4.14)

where \(\mathbf {k}_{(N)}=(k_1,\dots , k_N)\in \mathbb {N}^N,\) \(\mathbf {t}_{(N)}=(t_1, \dots , t_N)\in \mathbb {R}^N_+\) and

$$\begin{aligned} \mathscr {L}_{X; \mathbf {Y}}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big )=e^{-t_1 X}Y_{k_1}e^{-t_2 X}\ldots e^{-t_N X}Y_{k_N} e^{-(\beta -\sum _{j=1}^N t_j)X} \end{aligned}$$
(4.15)

with \(\mathbf {Y}=\{Y_j\}_j\). By this fact and (4.11), we observe that

$$\begin{aligned}&\mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R}\Big [ \mathscr {D}_{N, \beta }(A, B; \mathbf {C}, \mathbf {D}) \Big ]\nonumber \\ =&\sum _{k_1, \dots , k_N\ge 1}\int _{S_N(\beta )} \Big \{\mathrm {Tr}_{\mathfrak {X}_L}\Big [\mathscr {L}_{A; \mathbf {C}}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)} \big ) \Big ] \Big \} \times \Big \{\mathrm {Tr}_{\mathfrak {X}_L}\Big [\mathscr {L}_{B; \mathbf {D}}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big ) \Big ] \Big \}^*. \end{aligned}$$
(4.16)

Let us introduce an inner product by

$$\begin{aligned} \langle F| G\rangle _{N, \beta }=\sum _{k_1, \dots , k_N\ge 1}\int _{S_N(\beta )}\, F\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big )G\big ( \mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big )^*. \end{aligned}$$
(4.17)

In terms of this inner product, we have

$$\begin{aligned}&\mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R}\Big [ \mathscr {D}_{N, \beta }(A, B; \mathbf {C}, \mathbf {D}) \Big ]=\Big \langle F_{A;\mathbf {C}}^{(N)}\Big | F_{B;\mathbf {D}}^{(N)}\Big \rangle _{N, \beta }, \end{aligned}$$
(4.18)

where

$$\begin{aligned} F_{X;\mathbf {Y}}^{(N)}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big )=\mathrm {Tr}_{\mathfrak {X}_L}\Big [ \mathscr {L}_{X; \mathbf {Y}} \big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big )\Big ]. \end{aligned}$$
(4.19)

By the Schwartz inequality, we have

$$\begin{aligned} \bigg | \mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R}\Big [ e^{-\beta H(A, B, \mathbf {C}, \mathbf {D})} \Big ] \bigg |^2&= \Bigg | \sum _{N\ge 0} \Big \langle F_{A;\mathbf {C}}^{(N)}\Big | F_{B;\mathbf {D}}^{(N)} \Big \rangle _{N, \beta }\Bigg |^2 \nonumber \\&\le \Bigg ( \sum _{N\ge 0}\big \Vert F_{A;\mathbf {C}}^{(N)}\big \Vert ^2_{N, \beta } \Bigg ) \Bigg ( \sum _{N\ge 0}\big \Vert F_{B; \mathbf {D}}^{(N)}\big \Vert ^2_{N, \beta } \Bigg ), \end{aligned}$$
(4.20)

where \(\Vert W \Vert ^2_{N, \beta }:=\langle W| W \rangle _{N, \beta }\). Finally, we remark that

$$\begin{aligned} \sum _{N \ge 0}\big \Vert F_{A; \mathbf {C}}^{(N)}\big \Vert ^2_{N, \beta }&=\sum _{N\ge 0} \sum _{k_1, \dots , k_N\ge 1}\int _{S_N(\beta )} \bigg |\mathrm {Tr}_{\mathfrak {X}}\Big [ \mathscr {L}_{A; \mathbf {C}}\big (\mathbf {k}_{(N)}; \mathbf {t}_{(N)}\big ) \Big ] \bigg |^2\nonumber \\&=\sum _{N\ge 0}\mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R} \Big [ \mathscr {D}_{N, \beta }(A, A; \mathbf {C}, \mathbf {C}) \Big ] =\mathrm {Tr}_{\mathfrak {X}_L\otimes \mathfrak {X}_R} \Big [ e^{-\beta H(A, A; \mathbf {C}, \mathbf {C})} \Big ]. \end{aligned}$$
(4.21)

Combining (4.20) and (4.21), we obtain the assertion for the case where \(C_j\) and \(D_j\) are self-adjoint.

We note that for general \(C_j\) and \(D_j\), these operators can be written as

$$\begin{aligned} C_j=\mathfrak {R}C_j+i \mathfrak {I}C_j,\ \ D_j=\mathfrak {R}D_j+i \mathfrak {I}D_j, \end{aligned}$$
(4.22)

where \(\mathfrak {R}C_j, \mathfrak {R}D_j, \mathfrak {I}C_j\) and \(\mathfrak {I}D_j\) are self-adjoint. Since

$$\begin{aligned} C_j\otimes \vartheta D_j\vartheta ^{-1}+C_j^*\otimes \vartheta D_j^* \vartheta ^{-1} = 2\left( \mathfrak {R}C_j\otimes \vartheta \mathfrak {R}D_j\vartheta ^{-1}+\mathfrak {I}C_j\otimes \vartheta \mathfrak {I}D_j \vartheta ^{-1}\right) , \end{aligned}$$
(4.23)

we can reduce the problem to the case where \(C_j\) and \(D_j\) are self-adjoint. \(\square \)

Appendix 3: A Useful Lemma

Lemma 3.25

Let B and C be self-adjoint operators. Suppose that \(e^{-C}\) is a trace class operator and suppose that B is bounded. We have

$$\begin{aligned} \ln \mathrm {Tr}\big [e^{-(B+C)}\big ] \le \langle -B\rangle +\ln \mathrm {Tr}\big [ e^{-C} \big ], \end{aligned}$$
(4.24)

where \( \langle X \rangle =\mathrm {Tr}\big [ X e^{-(B+C)} \big ]\Big /\mathrm {Tr}\big [ e^{-(B+C)} \big ]\).

Proof

Let X and Y be self-adjoint. We know that \(F(\lambda )=\ln \mathrm {Tr}[e^{\lambda X+(1-\lambda ) Y}]\) is convex, e.g., as in [32]. Thus, we have \(F(1) \ge F'(0)+F(0)\), which implies

$$\begin{aligned} \ln \mathrm {Tr}[e^X]\ge \langle X-Y\rangle _Y+\ln \mathrm {Tr}[e^Y], \end{aligned}$$
(4.25)

where \(\langle L \rangle _Y=\mathrm {Tr}[L\, e^Y]\big /\mathrm {Tr}[e^Y]\). Substituting \(X=-C\) and \(Y=-B-C\), we obtain the desired result. \(\square \)

Appendix 4: Proof of Proposition 3.10

Let

$$\begin{aligned} \mathscr {S}_n^{(0)}= \big \{ (X_1,\dots , X_n)\, \big |\, X_j\in \Lambda \times \{\uparrow , \downarrow \}, \, j=1, \dots , n\ \text{ and }\ X_i\ne X_j,\, \text{ if }\ i\ne j \big \}. \end{aligned}$$
(4.26)

Let \(\mathfrak {S}_n\) be the permutation group on set \(\{1, \dots , n\}\). Let \((X_1, \dots , X_n),\, (Y_1, \dots , Y_n)\in \mathscr {S}_n^{(0)}\). If there exists a \(\pi \in \mathfrak {S}_n\) such that \((X_{\pi (1)}, \dots , X_{\pi (n)})=(Y_1, \dots , Y_n)\), then we write \( (X_1, \dots , X_n) \sim (Y_1, \dots , Y_n) \). The binary relation “\(\sim \)” on \(\mathscr {S}_n^{(0)}\) is an equivalence relation. We denote the quotient set \(\mathscr {S}_n^{(0)} \backslash \sim \) by \(\mathscr {S}_n\) and for the simplicity of notation, we still denote the equivalence class \([(X_1, \dots , X_n)]\) by \((X_1, \dots , X_n)\).

Set \(\displaystyle \mathscr {S}=\bigcup _{n=0}^{2|\Lambda |} \mathscr {S}_n \), where \(\mathscr {S}_{n=0}=\{\emptyset \}\). Let

$$\begin{aligned} \mathscr {S}_L&= \big \{ (X_1,\dots , X_n)\in \mathscr {S}\, \big |\, X_j\in \Lambda _L\times \{\uparrow , \downarrow \}, \, j=1, \dots , n,\ n\in \{0\} \cup \mathbb {N}\big \},\end{aligned}$$
(4.27)
$$\begin{aligned} \mathscr {S}_R&= \big \{ (X_1,\dots , X_n)\in \mathscr {S}\, \big |\, X_j\in \Lambda _L\times \{\uparrow , \downarrow \}, \, j=1, \dots , n,\ n\in \{0\}\cup \mathbb {N}\big \}. \end{aligned}$$
(4.28)

Here, if \(n=0\), then we understand that \((X_1, \dots , X_n)=\emptyset \). For each \(X=(x, \sigma )\in \Lambda \times \{\uparrow , \downarrow \}\), we set \(c_X:=c_{x\sigma }\) and \(a_X:=a_{x\sigma }\). For each \(\mathbf {X}=(X_1, \dots , X_n)\in \mathscr {S}\), we define

$$\begin{aligned} e(\mathbf {X})=c_{X_1}^*\dots c_{X_n}^* \Omega _{\mathrm {f}},\ \ \ f(\mathbf {X})=a_{X_1}^*\ldots a_{X_n}^*\Omega _{\mathrm {f}}, \end{aligned}$$
(4.29)

and \(e(\emptyset )=\Omega _{\mathrm {f}}\), \(f(\emptyset )=\Omega _{\mathrm {f}}\). The definition (4.29) is independent of the choice of the representative up to the sign factor, and trivially, \(\{e(\mathbf {X})\, |\, \mathbf {X}\in \mathscr {S}_R\}\) is a CONS of \(\mathfrak {F}_R\). We note that \(\{a_X\, |X\in \Lambda \times \{\uparrow , \downarrow \}\}\) satisfies the CARs:

$$\begin{aligned} \{a_X, a_Y^*\}=\delta _{XY},\ \ \ \{a_X, a_Y\}=0. \end{aligned}$$
(4.30)

Moreover, it holds that \(a_X\Omega _{\mathrm {f}}=0\) for all \(X\in \Lambda \times \{\uparrow , \downarrow \}\). Thus, \(\{f(\mathbf {X})\, |\, \mathbf {X}\in \mathscr {S}_L\}\) is a CONS of \(\mathfrak {F}_L\).

For each \(X=(x, \sigma )\in \Lambda _R\times \{\uparrow , \downarrow \}\), we set \(r(X):=(r(x), \sigma )\in \Lambda _L\times \{\uparrow , \downarrow \}\), where r in the right-hand side is defined by (3.42). For each \(\mathbf {X}=(X_1, \dots , X_n)\in \mathscr {S}_R\), we further extend the map r as follows:

$$\begin{aligned} r(\mathbf {X}):=(r(X_1), \dots , r(X_n))\in \mathscr {S}_L. \end{aligned}$$
(4.31)

Thus, \(\{f(r(\mathbf {X}))\, |\, \mathbf {X}\in \mathscr {S}_R\}\) is a CONS of \(\mathfrak {F}_L\). For each \(\Psi \in \mathfrak {F}_L\), we have the following expression:

$$\begin{aligned} \Psi =\sum _{\mathbf {X}\in \mathscr {S}_R} \Psi (r(\mathbf {X})) f(r(\mathbf {X})),\ \ \ \ \Psi (r(\mathbf {X}))=\langle f(r(\mathbf {X}))|\Psi \rangle . \end{aligned}$$
(4.32)

Using the expression (4.32), we define an antilinear map \(\xi \) from \(\mathfrak {F}_L\) onto \(\mathfrak {F}_R\) by

$$\begin{aligned} \xi \Psi =\sum _{\mathbf {X}\in \mathscr {S}_R} \overline{\Psi (r(\mathbf {X}))} e(\mathbf {X}) \end{aligned}$$
(4.33)

and \(\xi \Omega _{\mathrm {f}}^L=\Omega _{\mathrm {f}}^R\). \(\xi ^{-1}\) is given by

$$\begin{aligned} \xi ^{-1} \Phi =\sum _{\mathbf {X}\in \mathscr {S}_R} \overline{\Phi (\mathbf {X})} f(r(\mathbf {X})) \end{aligned}$$
(4.34)

for each \(\Phi =\sum _{X\in \mathscr {S}_R} \Phi (\mathbf {X}) e(\mathbf {X})\in \mathfrak {F}_R\). It is not difficult to check that \( \langle \xi \Psi _1|\xi \Psi _2\rangle =\overline{\langle \Psi _1|\Psi _2\rangle } \) for all \(\Psi _1, \Psi _2\in \mathfrak {F}_L\). Hence, \(\xi \) is an antiunitary transformation.

Lemma 3.26

For all \(X\in \Lambda _R\times \{\uparrow , \downarrow \}\), it holds that \(\xi a_{r(X)}\xi ^{-1}=c_X\).

Proof

For each \(\Phi =\sum _{\mathbf {Y}\in \mathscr {S}_R} \Phi (\mathbf {Y})e(\mathbf {Y})\in \mathfrak {F}_R\), we have, by (4.33) and (4.34),

$$\begin{aligned} \xi a_{r(X)}^* \xi ^{-1} \Phi&=\xi a_{r(X)}^* \sum _{\mathbf {Y}\in \mathscr {S}_R} \overline{\Phi (\mathbf {Y})}f(r(\mathbf {Y})) =\xi \sum _{\mathbf {Y}\in \mathscr {S}_R,\ X\notin \mathbf {Y}} \overline{\Phi (\mathbf {Y})}f(r(X, \mathbf {Y}))\nonumber \\&=\sum _{\mathbf {Y}\in \mathscr {S}_R,\ X\notin \mathbf {Y}} \Phi (\mathbf {Y})e(X, \mathbf {Y}) =c_X^* \Phi . \end{aligned}$$
(4.35)

Hence, \(\xi a_{r(X)}^* \xi ^{-1}=c_X^*\). \(\square \)

Recall that \(\mathfrak {H}_L=\mathfrak {F}_L\otimes L^2(\mathcal {Q}_L, d\mu _{\Lambda _L})\). We use the following identification:

$$\begin{aligned} \mathfrak {H}_L=\int _{\mathcal {Q}_L}^{\oplus } \mathfrak {F}_L d\mu _{\Lambda _L}({\varvec{\phi }}). \end{aligned}$$
(4.36)

Thus, each vector \(\Psi \in \mathfrak {H}_L\) is a \(\mathfrak {F}_L\)-valued measurable map on \(\mathcal {Q}_L\), i.e., \({\varvec{\phi }}\mapsto \Psi ({\varvec{\phi }})\). Now, we define an antiunitary transformation \(\vartheta \) from \(\mathfrak {H}_L\) onto \(\mathfrak {H}_R\) by

$$\begin{aligned} (\vartheta \Psi )({\varvec{\phi }})= (\xi \Psi )(r^{-1}({\varvec{\phi }}))\ \ \text{ a.e. } {\varvec{\phi }}\in \mathcal {Q}_R\text{, } \Psi \in \mathfrak {H}_L, \end{aligned}$$
(4.37)

where, for each \({\varvec{\phi }}=\{\phi _x\}_{x\in \Lambda _R} \in \mathcal {Q}_R\), we define \(r^{-1}({\varvec{\phi }})\in \mathcal {Q}_L\) by \( \big (r^{-1}({\varvec{\phi }})\big )_x =\phi _{r^{-1}(x)},\, x\in \Lambda _L \).

Remark 3.27

For each measurable function \(F({\varvec{\phi }})\ ({\varvec{\phi }}\in \mathcal {Q}_L)\) on \(\mathcal {Q}_L\), \(F(r^{-1}({\varvec{\phi }}))\ ({\varvec{\phi }}\in \mathcal {Q}_R)\) can be regarded as a function on \(\mathcal {Q}_R\). Let \(\Psi \in \mathfrak {H}_L\). By (4.32), we have the following expression:

$$\begin{aligned} \Psi ({\varvec{\phi }})= \sum _{\mathbf {X}\in \mathscr {S}_R} \Psi _{r(\mathbf {X})}({\varvec{\phi }}) f(r(\mathbf {X})) \ \ \text{ a.e. } {\varvec{\phi }}\in \mathcal {Q}_L, \end{aligned}$$
(4.38)

where \(\Psi _{r(\mathbf {X})}({\varvec{\phi }})=\langle f(r(\mathbf {X}))|\Psi ({\varvec{\phi }})\rangle \). Using this, we have

$$\begin{aligned} (\vartheta \Psi )({\varvec{\phi }})=\sum _{\mathbf {X}\in \mathscr {S}_R} \overline{\Psi _{r(\mathbf {X})}(r^{-1}({\varvec{\phi }}))} e(\mathbf {X})\ \ \text{ a.e. } {\varvec{\phi }}\in \mathcal {Q}_R. \ \ \ \ \diamondsuit \end{aligned}$$
(4.39)

Proposition 3.28

\(\vartheta \) satisfies all properties in (3.43) and (3.44).

Proof

By Lemma 3.26, it is easy to check that \(\vartheta a_{r(X)}\vartheta ^{-1}=c_X\).

Note that the action of the multiplication operator \(\phi _x\) is as follows: For each \(\Psi \in \mathfrak {H}_L\) and \(x\in \Lambda _L\),

$$\begin{aligned} (\phi _x \Psi )({\varvec{\phi }})= \phi _x \Psi ({\varvec{\phi }}) \ \ \text{ a.e. } {\varvec{\phi }}\in \mathcal {Q}_L. \end{aligned}$$
(4.40)

Thus, we have, for each \(\Psi \in \mathfrak {H}_L\) and \(x\in \Lambda _R\),

$$\begin{aligned} (\vartheta \phi _{r(x)}\Psi )({\varvec{\phi }})=\phi _x \xi \Psi (r^{-1}({\varvec{\phi }})) =(\phi _x\vartheta \Psi )({\varvec{\phi }})\ \ \text{ a.e. } {\varvec{\phi }}\in \mathcal {Q}_R, \end{aligned}$$
(4.41)

which implies \(\vartheta \phi _{r(x)} \vartheta ^{-1}=\phi _x\).

Next, we will prove that \(\vartheta \pi _{r(x)} \vartheta ^{-1}=-\pi _x\). Since \(\pi _x=-i \frac{\partial }{\partial \phi _x}\), we have, for each \(\Psi \in \mathfrak {H}_L\) and \(x\in \Lambda _L\),

$$\begin{aligned} (\pi _x \Psi )({\varvec{\phi }})= (-i)\frac{\partial \Psi }{\partial \phi _x}({\varvec{\phi }}) \ \ \text{ a.e. } {\varvec{\phi }}\in \mathcal {Q}_L. \end{aligned}$$
(4.42)

Hence, we have, for each \(\Psi \in \mathfrak {H}_L\) and \(x\in \Lambda _L\),

$$\begin{aligned} (\vartheta \pi _{x} \Psi )({\varvec{\phi }})=(+i) \overline{\frac{\partial \Psi }{\partial \phi _x}} (r^{-1}({\varvec{\phi }})) =(+i) \frac{\partial (\vartheta \Psi )}{\partial \phi _{r^{-1}(x)}} ({\varvec{\phi }}) =-(\pi _{r^{-1}(x)} \vartheta \Psi )({\varvec{\phi }}) \end{aligned}$$
(4.43)

for a.e. \({\varvec{\phi }}\in \mathcal {Q}_R\). Here, we used the fact that \(\xi \) is antilinear. Thus, we conclude that \(\vartheta \pi _{x} \vartheta ^{-1}=-\pi _{r^{-1}(x)}\) for each \(x\in \Lambda _L\), which implies \(\vartheta \pi _{r(x)} \vartheta ^{-1}=-\pi _{x}\) for each \(x\in \Lambda _R\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Miyao, T. Long-Range Charge Order in the Extended Holstein–Hubbard Model. J Stat Phys 165, 225–245 (2016). https://doi.org/10.1007/s10955-016-1617-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1617-6

Keywords

Navigation