Skip to main content
Log in

On Derivation of the Poisson–Boltzmann Equation

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

Abstract

Starting from the microscopic reduced Hartree–Fock equation, we derive the macroscopic linearized Poisson–Boltzmann equation for the electrostatic potential associated with the electron density.

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

Similar content being viewed by others

Notes

  1. The REHF obtained from the Hartree–Fock equation (HFE) by omitting the exchange term, see below.

  2. The decomposition \(L^2+L^2_{\mathrm{per}}\) is unique: if \(f\in L^2+L^2_{\mathrm{per}}\), then the periodic part, \(f_{\mathrm{per}}\), of f is given by the Fourier coefficients

    $$\begin{aligned}{\hat{f}}_{\mathrm{per}}(k):=\lim _{n\rightarrow \infty } \frac{1}{|\Lambda _n|} (2\pi )^{-d/2}\int _{\Lambda _n}e^{ik\cdot x}f(x) dx,\ k\in \mathcal {L}^*,\end{aligned}$$

    where \(\Lambda _n:=\cup _{\lambda \in \mathcal {L}_n}(\Omega +\lambda )\), with \(\mathcal {L}_n:=\mathcal {L}\cap [-n, n]^d\) and \(\Omega \) an arbitrary fundamental cell of \(\mathcal {L}\), and \(\mathcal {L}^*\) is the reciprocal lattice. Hence \(L^2+L^2_{\mathrm{per}}\) is a Hilbert space with the inner product which is sum of the inner products in \(L^2\) and \(L^2_{\mathrm{per}}\). The operator \(\Delta \) on \(L^2+L^2_{\mathrm{per}}\) is self-adjoint on the natural domain (i.e. \(H^2+H^2_{\mathrm{per}}\)) and is invertible on the subspace \(L^2+(L^2_{\mathrm{per}})^\perp \).

References

  1. Anantharaman, A., Cancès, E.: Existence of minimizers for Kohn-Sham models in quantum chemistry. Ann. I. H. Poincaré - AN 26, 2425–2455 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  2. Bach, V., Breteaux, S., Chen, Th., Fröhlich, J.M., Sigal, I.M.: The time-dependent Hartree-Fock-Bogoliubov equations for bosons. J. Evol. Equ. 2020 (to appear). arXiv:1602.05171v2

  3. Bach, V., Lieb, E.H., Solovej, J.P.: Generalized Hartree-Fock theory and the Hubbard model. J. Stat. Phys. 76, 3–89 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  4. Brislawn, C.: Kernels of trace class operators, Proceedings AMS 104. No. 4 (1988)

  5. Cancès, E., Deleurence, A., Lewin, M.: A new approach to the modeling of local defects in crystals: the reduced Hartree-Fock case. Commun. Math. Phys. 281(1), 129–177 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  6. Cancès, E., Deleurence, A., Lewin, M.: Non-perturbative embedding of local defects in crystalline materials. J. Phys. 20, 294213 (2008)

    Google Scholar 

  7. Cancès, E., Lewin, M.: The dielectric permittivity of crystals in the reduced Hartree-Fock approximation. Arch. Ration. Mech. Anal. 197(1), 139–177 (2010)

    Article  MathSciNet  Google Scholar 

  8. Cancès, E., Lewin, M., Stoltz, G.: The Microscopic Origin of the Macroscopic Dielectric Permittivity of Crystals. Lecture Notes in Computational Science and Engineering, vol. 82. Springer (2011)

  9. Cancès, E., Stoltz, G.: A mathematical formulation of the random phase approximation for crystals. Ann. I. H. Poincaré - AN 29(6), 887–925 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  10. Catto, I., Le Bris, C., Lions, P.-L.: On the thermodynamic limit for Hartree-Fock type models. Ann. I. H. Poincaré - AN 18(6), 687–760 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  11. Catto, I., Le Bris, C., Lions, P.-L.: On some periodic Hartree type models. Ann. I. H. Poincaré - AN 19(2), 143–190 (2002)

    Article  ADS  MathSciNet  Google Scholar 

  12. Chenn, I., Sigal, I.M.: On the Bogolubov-de Gennes equations. arXiv:1701.06080v2 (2019)

  13. Chenn, I., Sigal, I.M.: On effective PDEs of quantum physics. In: D’Abbicco, M., et al. (eds.) New Tools for Nonlinear PDEs and Application, Birkhäuser Series. Trends in Mathematics.

  14. Cycon, H., Froese, R., Kirsch, W., Simon, B.: Schrödinger Operators (with Applications to Quantum Mechanics and Global Geometry). Springer, Berlin (1987)

    Book  Google Scholar 

  15. E, W., Lu, J.: Electronic structure of smoothly deformed crystals Cauchy-Born Rule for the nonlinear tight-binding model. Commun. Pure Appl. Math 63(11), 1432–1468 (2010)

    Article  MathSciNet  Google Scholar 

  16. E, W., Lu, J.: The electronic structure of smoothly deformed crystals: Wannier functions and the Cauchy-Born rule. Arch. Ration. Mech. Anal. 199(2), 407–433 (2011)

    Article  MathSciNet  Google Scholar 

  17. E, W., Lu, J.: The Kohn-Sham equation for deformed crystals. Mem. AMS (2013)

  18. E, W., Lu, J., Yang, X.: Effective Maxwell equations from time-dependent density functional theory. Acta. Math. Sin.-English Ser 27, 339–368 (2011)

    ADS  MathSciNet  MATH  Google Scholar 

  19. Fogolari, F., Brigo, A., Molinari, H.: The Poisson-Boltzmann equation for biomolecular electrostatics: a tool for structural biology. J. Mol. Recognit. 15, 377–392 (2002)

    Article  Google Scholar 

  20. Gustafson, S.J., Sigal, I.M.: Mathematical Concepts of Quantum Mechanics, 2nd edn. Universitext, Springer (2011)

    Book  Google Scholar 

  21. Hainzl, Ch., Lewin, M., Seré, E.: Existence of atoms and molecules in the mean-field approximation of no-photon quantum electrodynamics. Arch. Ration. Mech. Anal. 192(3), 453–499 (2009)

    Article  MathSciNet  Google Scholar 

  22. Le Bris, C., Lions, P.-L.: From atoms to crystals: a mathematical journey. J. Bull AMS 42, 291–363 (2005)

    Article  MathSciNet  Google Scholar 

  23. Levitt, A.: Screening in the finite-temperature reduced Hartree-Fock model. arXiv:1810.03342v1 (2018)

  24. Lieb, E., Loss, M.: Analysis, 2nd edn. AMS Press, Providence, RI (2001)

    MATH  Google Scholar 

  25. Lieb, E.H.: The stability of matter: from atoms to stars. Bull. AMS 22, 1–49 (1990)

    Article  MathSciNet  Google Scholar 

  26. Lieb, E.H., Simon, B.: The Hartree-Fock theory for Coulomb systems. Commun. Math. Phys. 53(3), 185–194 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  27. Lindblad, G.: Expectations and entropy inequalities for finite quantum systems. Commun. Math. Phys. 39, 111–119 (1974)

    Article  ADS  MathSciNet  Google Scholar 

  28. Lions, P.L.: Hartree-Fock and Related Equations. Nonlinear Partial Differential Equations and Their Applications. Collège de France Seminar, vol. IX. Pitman Res. Notes Math. Ser. vol. 181, pp. 304–333 (1988)

  29. Lions, P.L.: Solutions of Hartree-Fock equations for Coulomb systems. Commun. Math. Phys. 109, 33–97 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  30. Markowich, P.A., Rein, G., Wolansky, G.: Existence and nonlinear stability of stationary states of the Schrödinger-Poisson system. J. Stat. Phys. 106(5–6), 1221–1239 (2002)

    Article  Google Scholar 

  31. Mermin, N.D.: Thermal properties of the inhomogeneous electron gas. Phys. Rev. 137, A1441 (1965)

    Article  ADS  MathSciNet  Google Scholar 

  32. Nier, F.: A variational formulation of Schrödinger-Poisson systems in dimension \(d \le 3\). Commun. PDEs 18(7 and 8), 1125–1147 (1993)

  33. Prodan, E., Nordlander, P.: On the Kohn-Sham equations with periodic background potential. J. Stat. Phys. 111(3–4), 967–992 (2003)

    Article  MathSciNet  Google Scholar 

  34. Reed, M., Simon, B.: Method of Modern Mathematical Physics IV: Analysis of Operators. Academic Press, London (1978)

    MATH  Google Scholar 

  35. Sevik, C., Bulutay, C.: Theoretical study of the insulating oxides and nitrate: \(SiO_2\), \(GeO_2\), \(Al_2 O_3\), \(Si_3 N_4\), and \(Ge_3 N_4\). arXiv:cond-mat/0610176v2 (2008)

  36. Simon, B.: Trace Ideals and Their Applications, 2nd edn. AMS Press, Providence, RI (2005)

    MATH  Google Scholar 

Download references

Acknowledgements

The authors thank Rupert Frank, Jürg Fröhlich, Gian Michele Graf, Christian Hainzl and Jianfeng Lu for stimulating discussions and to the anonymous referee for many pertinent remarks. The correspondence with Antoine Levitt played a crucial role in steering the research at an important junction. The second author is also grateful to Volker Bach, Sébastien Breteaux, Thomas Chen and Jürg Fröhlich for enjoyable collaboration on related topics. The research on this paper is supported in part by NSERC Grant No.NA7901. The first author is also in part supported by NSERC CGS D graduate scholarship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to I. M. Sigal.

Additional information

Communicated by Ivan Corwin.

Publisher's Note

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

To Joel with friendship and admiration.

Appendices

Appendix A: \(\epsilon (T) \rightarrow \epsilon (0)\) as \(T \rightarrow 0\)

Lemma A 1

Let \(\text {xc}= 0\). Then \(\epsilon \equiv \epsilon (T) \rightarrow \epsilon (0)\) as \(T \rightarrow 0\), where \(\epsilon (0)\) is the dielectric constant for \(T=0\) obtained in [7].

Proof

We see from (1.35) below that \(\epsilon (T), T=1/\beta ,\) is of the form

$$\begin{aligned} \epsilon (T) = \frac{1}{2\pi i} \int _\Gamma f_T(z-\mu ) X(z) \end{aligned}$$
(A.1)

where X(z) is some holomorphic function on \(\mathbb {C}\backslash \mathbb {R}\), independent of \(\beta \), and remains holomorphic on the real axis where the gap of \(h_{\mathrm{per}}\) occurs. On \(\mathbb {R}\), we note that \(f_{\mathrm{FD}}(\beta x)\) converges to the indicator function \(\chi _{(-\infty , 0)}\) as \(\beta \rightarrow \infty \). If we take \(\beta \rightarrow \infty \), the integral

$$\begin{aligned} \frac{1}{2\pi i} \int _\Gamma f_T(z-\mu ) X(z) \end{aligned}$$
(A.2)

converges to \(\frac{1}{2\pi i} \int _{G_1} X(z)\) where \(G_1\) is any contour around the part of the spectrum of \(h_{\mathrm{per}}\) that is less than \(\mu _{\mathrm{per}}\). This is the same expression as in [7] after inserting \(1 = \sum _i |\varphi _i \rangle \langle \varphi _i|\) for each resolvent of \(h_{\mathrm{per}}\) in X(z) where the \(\varphi _i\)’s are eigenvectors of \(h_{\mathrm{per}}\). \(\square \)

Appendix B: Bounds on m and V

In this section, we prove bounds on m and V given (1.33) and (4.28). Note that \(m=\Vert V\Vert _{L^1_{\mathrm{per}}}\). Since \(f'_{T} < 0\) (\(T= 1/\beta \)), (4.28) implies that \(V > 0\) and therefore, by (4.28), \(\Vert V\Vert _{L^1_{\mathrm{per}}} = \int _\Omega V\), where \(\Omega \) is a fundamental domain of \(\mathcal {L}\) (see Sect. 1.5), which yields

$$\begin{aligned} m=\int _\Omega V = - \mathrm {Tr}_{L^2_{\mathrm{per}}} f'_{T}(h_{\mathrm{per}, 0}-\mu ),\ \qquad \mu =\mu _{\mathrm{per}}. \end{aligned}$$
(B.1)

Lemma B.1

Let Assumption [A1] hold and \(\eta _0\) be given in (1.27). Then

$$\begin{aligned} m=\Vert V\Vert _{L^1_{\mathrm{per}}} \ge \frac{1}{4} \beta e^{-\beta \eta _0}, \end{aligned}$$
(B.2)

where \(\eta _0\) is given in (1.27).

Proof

Using that \(\eta _0\) is the smallest distance between \(\mu = \mu _{\mathrm{per}}\) and the spectrum of \(h_{\mathrm{per}, 0}\) (see (1.27)) and Eq. (B.1) and replacing \(\mathrm {Tr}_{L^2_{\mathrm{per}}} f'_{T}(h_{\mathrm{per}, 0}-\mu )\) by the contribution of the eigenvalue of \(h_{\mathrm{per}, 0}\) closest to \(\mu \), we find

$$\begin{aligned} m \ge&- f'_{T}(\eta _0) = \beta \frac{e^{\beta \eta _0}}{(1+e^{\beta \eta _0})^2} \ge \frac{1}{4} \beta e^{-\beta \eta _0}. \end{aligned}$$
(B.3)

This gives (B.2). \(\square \)

Lemma B.2

Let Assumption [A1] hold. Then, for \(1 \le p \le \infty \),

$$\begin{aligned} \Vert V\Vert _{L^p_{\mathrm{per}}} \lesssim \beta e^{-\eta _0\beta }. \end{aligned}$$
(B.4)

Proof

We do the case for \(p=1\) and \(p=\infty \), and conclude the lemma by interpolation. By Assumption [A1], the potential \(\phi _{\mathrm{per}}\) is bounded. Thus, \(h_{\mathrm{per}, 0}\) has only discrete spectrum on \(L^2_{\mathrm{per}}(\mathbb {R}^3)\) and

$$\begin{aligned} \frac{1}{\beta }\Vert V\Vert _{L^1_{\mathrm{per}}}&= \sum _{\lambda \in \sigma (h_{\mathrm{per}, 0})} \frac{ e^{\beta (\lambda -\mu )}}{(1+e^{\beta (\lambda -\mu )})^2} \end{aligned}$$
(B.5)
$$\begin{aligned}&\le \sum _{\mu > \lambda \in \sigma (h_{\mathrm{per}, 0})}e^{\beta (\lambda -\mu )}+ \sum _{\mu < \lambda \in \sigma (h_{\mathrm{per}, 0})}e^{-\beta (\lambda -\mu )} \end{aligned}$$
(B.6)
$$\begin{aligned}&= \sum _{\lambda \in \sigma (h_{\mathrm{per}, 0})}e^{\beta |\lambda -\mu |} . \end{aligned}$$
(B.7)

Again, we use that \(\eta _0\) is the smallest distance between \(\mu = \mu _{\mathrm{per}}\) and the spectrum of \(h_{\mathrm{per}, 0}\) (see (1.27)). Peeling the eigenvalue(s) closest to \(\mu \) and letting \(\eta _0+\xi \) stand for the distance between \(\mu \) and the rest of the spectrum \(\sigma (h_{\mathrm{per}, 0})\), we find, for some constant c,

$$\begin{aligned} \sum _{\lambda \in \sigma (h_{\mathrm{per}, 0})}&e^{\beta |\lambda -\mu |} =ce^{\beta \eta _0}+ \sum _{\lambda \in \sigma (h_{\mathrm{per}, 0}), |\lambda -\mu |\ge \eta _0+\xi }e^{\beta |\lambda -\mu |}. \end{aligned}$$
(B.8)

We estimate the sum on the r.h.s. by an integral as follows. Since the potential \(\phi _{\mathrm{per}}\) is infinitesimally bounded with respect to \(-\Delta \), the eigenvalues of \(h_{\mathrm{per}, 0}\) go to infinity at a similar rate as those of \(-\Delta \) (on \(L^2_{\mathrm{per}}(\mathbb {R}^3)\)), i.e. as \(n^2\). Thus, assuming that for \(\lambda \) sufficiently large, the nth eigenvalue \(\lambda _n\approx n^2\) has the degeneracy of the order \(O(n^k), k\ge 0\), we conclude that

$$\begin{aligned}&\sum _{\mu < \lambda \in \sigma (h_{\mathrm{per}, 0})}e^{-\beta (\lambda -\mu )}\lesssim \int _{x^2 \ge \mu +\eta _0+\xi } x^k e^{-\beta (x^2 -\mu )} d x \nonumber \\&\quad =\frac{1}{2} \int _{y\ge \eta _0+\xi } (y+\mu )^{\frac{k-1}{2}} e^{-\beta y} d y \lesssim \frac{1}{\beta }\mu ^{\frac{k-1}{2}} e^{-\beta (\eta _0+\xi )}. \end{aligned}$$
(B.9)

For the first sum in (B.6), we consider separately the cases \(\mu \lesssim 1\) and \(\mu \gg 1\) and, in the 2nd case, break the sum into the sums over \(\lambda \lesssim 1\) and \(\lambda \gg 1\). In the first three situations, the estimate is straightforward and in the last one, we proceed as in (B.9) to obtain

$$\begin{aligned} \sum _{\mu -\eta _0-\xi \ge \lambda \in \sigma (h_{\mathrm{per}, 0})}e^{\beta (\lambda -\mu )}\lesssim \frac{1}{\beta }\mu ^{\frac{k-1}{2}} e^{-\beta \eta _0}. \end{aligned}$$

This proves the lemma for \(p=1\).

Let \(W^{4,1}_{\mathrm{per}}\) be the usual Sobolev space associated to \(L^1_{\mathrm{per}}\) involving up to 4 derivatives. For the case \(p=\infty \), we use the Sobolev inequality

$$\begin{aligned} \Vert f\Vert _\infty \lesssim \Vert f\Vert _{W^{4,1}_{\mathrm{per}}} \end{aligned}$$
(B.10)

for \(f \in W^{4,1}_{\mathrm{per}}\). Thus, it suffices for us to estimate \(\Vert \nabla ^j V\Vert _{L^1_{\mathrm{per}}}, j=0,\ldots ,4\). To this end, we note that

$$\begin{aligned} \nabla {\text {den}}(A) = {\text {den}}( [\nabla , A] ) \end{aligned}$$
(B.11)

for an operator A on \(L^2_{\mathrm{per}}(\mathbb {R}^3)\). Thus, it suffices that we estimate the trace 1-norm of \(\nabla ^s f_{T}'(h_{\mathrm{per}, 0}-\mu ) \nabla ^{4-s}\) on \(L^2_{\mathrm{per}}(\mathbb {R}^3)\) for \(s=0,\ldots ,4\). Since the potential \(\phi _{\mathrm{per}}\) is bounded together with all its derivatives, we have, for \(s=0, \dots , 4\),

$$\begin{aligned} \Vert \nabla ^s h^{-s/2}\Vert \lesssim 1, \end{aligned}$$
(B.12)

where \(h:=h_{\mathrm{per}, 0}+c\), with \(c>0\) s.t. \(h_{\mathrm{per}, 0}+c>0\). Indeed, to fix ideas, consider one of the terms, say, \(\Vert \nabla ^3 h^{-3/2}\Vert \). We have \(\Vert \nabla ^3f\Vert ^2\le \Vert (-\Delta )^{3/2}f\Vert ^2=\langle f, (h+\phi _{\mathrm{per}}-c)^3f\rangle \). Taking \(f = h^{-3/2}u\), expanding the binomial \((h+\phi _{\mathrm{per}}-c)^3\) and commuting the operator h in the resulting terms \(h^2\phi _{\mathrm{per}}\) and \(\phi _{\mathrm{per}} h^2\) to the right and left, respectively, and estimating the resulting commutators, \([h, \phi ]\) and \([\phi _{\mathrm{per}}, h]=-[h, \phi _{\mathrm{per}}]\), we arrive at the estimate \(\Vert \nabla ^3 h^{-3/2}\Vert \lesssim 1\) as claimed. (B.12) implies also that \( \Vert h^{-2+s/2}\nabla ^{4-s}\Vert \lesssim 1\), for \(j=0, \dots , 4\). As the result, we have

$$\begin{aligned} \Vert \nabla ^s f_{T}'(h_{\mathrm{per}, 0}-\mu ) \nabla ^{4-s}\Vert _{S^1}\lesssim \Vert g(h_{\mathrm{per}, 0})\Vert _{S^1}, \end{aligned}$$

where \(g(x):=-(x+c)^{s/2} f_{T}'(x-\mu ) (x+c)^{2-s/2}\ge 0\). Hence, it suffices to estimate \(\Vert g(h_{\mathrm{per}, 0})\Vert _{S^1}=\mathrm {Tr}[g(h_{\mathrm{per}, 0})]\). The latter can be done the same way as the case for \(p=1\) by summing eigenvalues of \(h_{\mathrm{per}, 0}\) and the lemma is proved. \(\square \)

Appendix C: Bound on \(M_\delta \)

In an analogy to \(L_{\mathrm{per}}^2\equiv L_{\mathrm{per}}^2(\mathbb {R}^3)\) given in (1.19), we let

$$\begin{aligned} L^2_{\mathrm{per}, \delta }\equiv L^2_{\mathrm{per}, \delta }(\mathbb {R}^3) := \{ f \in L^2_{\mathrm{loc}}(\mathbb {R}^3) : f \text { is } \mathcal {L}_\delta \text {-periodic } \}. \end{aligned}$$
(C. 1)

Moreover, we recall \(\nabla ^{-1} := \nabla (-\Delta )^{-1}\) (see (3.56)). The main result of this appendix is the following

Proposition C.1

Let Assumption [A1] hold. Then the operator \(M_\delta \) can be decomposed as

$$\begin{aligned} M_\delta = M_\delta ' + M_\delta '' \, , \end{aligned}$$
(C. 2)

with the operator \(M_\delta '\) and \(M_\delta ''\) satisfying the estimates

$$\begin{aligned}&\Vert {\bar{P_r}} \nabla ^{-1} M_\delta ' P_r\varphi \Vert _{L^2} \lesssim \delta ^{-1} \Vert V\Vert _{L^2_{\mathrm{per}}} \Vert P_r\varphi \Vert _{L^2}, \end{aligned}$$
(C. 3)
$$\begin{aligned}&\Vert {\bar{P_r}} \nabla ^{-1} M_\delta '' P_r\nabla ^{-1} \varphi \Vert _{L^2} \lesssim \Vert P_r\varphi \Vert _{L^2}. \end{aligned}$$
(C. 4)

Proof of Proposition C.1

Proposition 4.2 and the rescaling relation (3.34) imply the explicit form for the k-fibers of \(M_\delta \):

Lemma C.2

Then \(M_\delta \) has a Bloch–Floquet decomposition (2.34) with \(\mathcal {L}= \mathcal {L}_\delta \), whose \(k-\)fiber \(M_{\delta , k}\) acting on \(L^2_{\mathrm{per}, \delta }\) is given by

$$\begin{aligned} M_{\delta ,k} f =&- \delta {\text {den}}\left[ \oint r^\delta _{\mathrm{per, 0}}(z)f r^\delta _{\mathrm{per}, k}(z)\right] \end{aligned}$$
(C. 5)

where \(f \in L^2_{\mathrm{per}, \delta }\) and, on \(L^2_{\mathrm{per}, \delta }\),

$$\begin{aligned}&r^\delta _{\mathrm{per}, k}= (z- h^\delta _{\mathrm{per}, k})^{-1}, \quad h^\delta _{\mathrm{per}, k}= \delta ^2 (-i\nabla -k)^2 + \delta \phi _{\mathrm{per}}^\delta . \end{aligned}$$
(C. 6)

We decompose the operator \(M_{\delta , k}\) acting on \(L^2_{\mathrm{per}, \delta }\) as

$$\begin{aligned} M_{\delta , k}&=: M_{\delta , 0} + M_{\delta , k}'' \, , \end{aligned}$$
(C. 7)

where \(M_{\delta , 0} = M_{\delta , k=0}\) and \(M_{\delta , k}'\) is defined by the expression (C. 7). We define operators \(M_\delta '\) and \(M_\delta ''\) on \(L^2(\mathbb {R}^2)\) via

$$\begin{aligned} M_\delta '&:= \int _{\Omega _\delta ^*}^\oplus d{\hat{k}} \, M_{\delta , 0} \varphi , \end{aligned}$$
(C. 8)
$$\begin{aligned} M_\delta ''&:= \int _{\Omega _\delta ^*}^\oplus d{\hat{k}} \, M''_{\delta , k} \varphi \, , \end{aligned}$$
(C. 9)

where \(\Omega _\delta ^*\) is a fundamental cell of the reciprocal lattice to \(\mathcal {L}_\delta \) and \(d{\hat{k}} = |\Omega _\delta ^*|^{-1} dk\). By Lemma C.2 and definition (C. 7), the latter operators satisfy (C. 2).

Lemma C.3

\(M_\delta '\) (see (C. 8)) restricted to the range of \(P_r\) is a multiplication operator given by

$$\begin{aligned} (M_\delta ' P_r\varphi )(x) = V_\delta (x) (P_r\varphi )(x), \end{aligned}$$
(C. 10)

where

$$\begin{aligned} V_\delta (x) = -\delta ^{-2} {\text {den}}\left[ f_{T}'(h_{\mathrm{per}, 0}-\mu ) \right] (\delta ^{-1}x) \, , \end{aligned}$$
(C. 11)

with \(h_{\mathrm{per}, 0}\) given in (4.6) (with \(k = 0\)).

Proof

By (C. 5) and definition of \(M_\delta '\) in (C. 8), we see that

$$\begin{aligned} M_\delta '&P_r \varphi = - \int ^\oplus _{\Omega _\delta ^*} d{\hat{k}} \, \delta {\text {den}}\left[ \oint r^\delta _{\mathrm{per, 0}}(z)(P_r \varphi )_k r^\delta _{\mathrm{per, 0}}(z)\right] . \end{aligned}$$
(C. 12)

By Corollary 2.4 and the Cauchy integral formula,

$$\begin{aligned} M_\delta ' P_r\varphi&= - \int ^\oplus _{\Omega _\delta ^*} d{\hat{k}} \, \delta {\text {den}}\left[ \oint (r^\delta _{\mathrm{per, 0}}(z))^2 \right] |\Omega _\delta |^{-1} {\hat{\varphi }}(k) \end{aligned}$$
(C. 13)
$$\begin{aligned}&= - \int ^\oplus _{\Omega _\delta ^*} d{\hat{k}} \, {\text {den}}[f_{T}'(h^\delta _{\mathrm{per}, 0} - \mu )] |\Omega _\delta |^{-1} {\hat{\varphi }}(k) . \end{aligned}$$
(C. 14)

where \(r^\delta _{\mathrm{per, 0}}(z)\) and \(h^\delta _{\mathrm{per}, 0}\) are given in (C. 6). Applying the inverse Bloch–Floquet transform (2.24), (C. 14) implies

$$\begin{aligned} M_\delta '&P_r\varphi = -\delta {\text {den}}[f_{T}'(h^\delta _{\mathrm{per}, 0}-\mu ) ] \int _{\Omega _\delta ^*} d{\hat{k}} \, e^{-ikx} |\Omega _\delta |^{-1}{\hat{\varphi }}(k). \end{aligned}$$
(C. 15)

Since \(d{\hat{k}}\) is normalized by the volume \(|\Omega _\delta ^*|\) (which is independent of the choice of the cell), (C. 15) shows

$$\begin{aligned} M_\delta ' P_r\varphi =&- \delta {\text {den}}\left[ f_{T}'(h^\delta _{\mathrm{per}, 0}-\mu ) \right] P_r\varphi . \end{aligned}$$
(C. 16)

By Lemma 2.7 and recalling the definition of \(U_{\delta }\) from (2.45), we see that

$$\begin{aligned} \delta {\text {den}}\left[ f_{T}'(h^\delta _{\mathrm{per}, 0}-\mu ) \right]&= \delta {\text {den}}\left[ U_{\delta }f_{T}'(h_{\mathrm{per}, 0}-\mu ) U_{\delta }^* \right] \end{aligned}$$
(C. 17)
$$\begin{aligned}&= \delta ^{-2} {\text {den}}\left[ f_{T}'(h_{\mathrm{per}, 0}-\mu ) \right] (\delta ^{-1}x) \, , \end{aligned}$$
(C. 18)

where \(h_{\mathrm{per}, 0} = h_{\mathrm{per}, 0}^{\delta = 1}\), which together with (C. 16) gives (C. 10)–(C. 11). \(\square \)

Proof of (C. 3)

Let \(V_\delta \) be given in (C. 11). Since the Bloch–Floquet decomposition is unitary, we see, by Lemma C.3 and Corollary 2.4, that

$$\begin{aligned} \Vert M_\delta ' P_r\varphi \Vert _{L^2}^2 =&\Vert V_\delta P_r\varphi \Vert _{L^2}^2 = \int _{\Omega _\delta ^*} d{\hat{k}} \Vert V_\delta |{\hat{\varphi }}(k)| |\Omega _\delta |^{-1} \Vert _{L^2_{\mathrm{per}, \delta }}^2, \end{aligned}$$
(C. 19)

where \(L^2_{\mathrm{per}, \delta }\) is given in (C. 1). Using the fact that \(d{\hat{k}} = |\Omega _\delta ^*|^{-1} dk\) and \(|\Omega _\delta | = \delta ^3 |\Omega |\), (C. 19) implies

$$\begin{aligned} \Vert M_\delta ' P_r\Vert _{L^2}^2 =&\delta ^{-3} |\Omega |\Vert V_\delta \Vert _{L^2_{\mathrm{per}, \delta }}^2 \Vert P_r\varphi \Vert _{L^2}^2. \end{aligned}$$
(C. 20)

By a change of variable, we see that \(\Vert V_\delta \Vert _{L^2_{\mathrm{per}, \delta }} = \delta ^{-1/2}\Vert V \Vert _{L^2_{\mathrm{per}}},\) where V is given by (4.28). Combing with (C. 20), the fact \({\bar{P_r}} (-i\nabla )^{-1} \lesssim r^{-1}\) (where \(\nabla ^{-1}\) is given in (3.56)) and \(r^{-1}= a^{-1}\delta \lesssim \delta \) (see (3.38)) yields Eq. (C. 3). \(\square \)

Proof of (C. 4)

Let \(M_\delta ''\) be given by (C. 9) and \(k^{-1} := k/|k|^2\). Let \(\varphi \in L^2(\mathbb {R}^3)\). By Corollary 2.4, we have

$$\begin{aligned} (P_r\nabla ^{-1} \varphi )_k= k^{-1} {\hat{\varphi }}(k) |\Omega _\delta |^{-1} \chi _{B(r)}(k). \end{aligned}$$
(C. 21)

This gives \(M_\delta '' P_r\nabla ^{-1} \varphi = |\Omega _\delta |^{-1}\int ^\oplus _{B(r)} d{\hat{k}} M''_{\delta , k} k^{-1} {\hat{\varphi }}(k)\). Since the Bloch–Floquet decomposition is unitary, we see, using (C. 21), that

$$\begin{aligned} \Vert M_\delta '' P_r\nabla ^{-1}&\varphi \Vert _{L^2}^2 \nonumber \\&= |\Omega _\delta |^{-2}\int _{B_r} d{\hat{k}} \, \Vert M''_{\delta , k} 1\Vert _{L^2_{\mathrm{per}, \delta }}^2 |k|^{-2} |{\hat{\varphi }}(k)|^2. \end{aligned}$$
(C. 22)

Since \(d {\hat{k}} = |\Omega ^*|^{-1} dk = |\Omega | dk\) and \(|\Omega _\delta | = \delta ^3|\Omega |\), (C. 22) is bounded as

$$\begin{aligned} \Vert M_\delta '' P_r\nabla ^{-1}&\varphi \Vert _{L^2}^2 \nonumber \\&\lesssim \delta ^{-3} \sup _{k \in B_r} \left( \Vert M''_{\delta , k} 1\Vert _{L^2_{\mathrm{per}, \delta }}^2 |k|^{-2} \right) \Vert P_r\varphi \Vert _{L^2}^2, \end{aligned}$$
(C. 23)

where \(1 \in L^2_{\mathrm{per}, \delta }\) is the constant function 1 and \(L^2_{\mathrm{per}, \delta }\) is given in (C. 1).

By (C. 5) and (C. 7), we have, for \(M''_{\delta , k}\) given in (C. 7), that

$$\begin{aligned} M''_{\delta , k} \varphi =&- \delta {\text {den}}\left[ \oint r_{\mathrm{per}, 0}(z) \varphi (r^\delta _{\mathrm{per}, k}(z)-r^\delta _{\mathrm{per}, 0}(z)) \right] . \end{aligned}$$
(C. 24)

Since \(r^\delta _{\mathrm{per}, k}(z)-r_{\mathrm{per}, 0}(z)=r^\delta _{\mathrm{per}, 0}(z)A_k r^\delta _{\mathrm{per}, k}(z)\), where \(A_k:=-2 (-i\nabla ) \delta k+\delta ^2 |k|^2\), this gives

$$\begin{aligned} M''_{\delta , k} \varphi =&- \delta {\text {den}}\oint \left[ r^\delta _{\mathrm{per}}(z) \varphi r^\delta _{\mathrm{per}}(z) A_k r^\delta _{\mathrm{per}, k}(z) \right] . \end{aligned}$$
(C. 25)

By the rescaling relation (3.34) and (C. 25), we see that

$$\begin{aligned} \Vert M''_{\delta , k} 1\Vert _{L^2_{\mathrm{per}, \delta }}&= \Vert U_\delta ^* M''_{\delta , k}U_\delta \cdot U_\delta ^* 1\Vert _{L^2_{\mathrm{per}}} = \delta ^{3/2-2}\Vert M''_{1, k} 1\Vert _{L^2_{\mathrm{per}}}\nonumber \\&= \delta ^{-1/2} \big \Vert {\text {den}}\big [ \oint r_{\mathrm{per}}^2(z) A_kr_{\mathrm{per}, \delta k}(z) \big ] \big \Vert _{L^2_{\mathrm{per}}}. \end{aligned}$$
(C. 26)

By (C. 26), notation \(A_k:=-2 (-i\nabla ) \delta k+\delta ^2 |k|^2\) and inequality (4.32), with \(\alpha =0,1/2\), we obtain, for \(|k| \le r\),

$$\begin{aligned} \Vert M''_{\delta , k}1\Vert _{L^2_{\mathrm{per}}} |k|^{-1}&\lesssim \delta ^{-1/2+1} + \delta ^{-1/2+2} r \nonumber \\&= \delta ^{1/2} + \delta ^{3/2} r \, . \end{aligned}$$
(C. 27)

By (3.38) and (C. 23), Eq. (C. 27) shows that

$$\begin{aligned} \Vert M_\delta '' P_r\nabla ^{-1} \varphi \Vert _{L^2} \lesssim \delta ^{-1} . \end{aligned}$$
(C. 28)

This bound, the observation that \(\Vert {\bar{P_r}} \nabla ^{-1}\Vert _\infty \lesssim r^{-1}\) (see (3.37)) and the definition \(r =a/\delta > rsim 1/\delta \) imply Eq. (C. 4). \(\square \)

This completes the proof of Proposition C.1. \(\square \)

We use Proposition C.1 to prove the following

Proposition C.4

Let Assumption [A1] hold and let \(\beta e^{-\eta _0\beta }\lesssim 1\) (which is weaker than Assumption [A3]). Then the operator \(M_\delta \) is bounded as

$$\begin{aligned} \Vert \nabla ^{-1}{\bar{P_r}}&M_\delta f\Vert _{\dot{H}^1} \lesssim \Vert f\Vert _{\delta }. \end{aligned}$$
(C. 29)

Proof of Proposition C.4

Decomposing \(M_\delta \) according to (C. 2) and using bounds Eqs. (C. 3) and (C. 4) of Proposition C.1, we see that

$$\begin{aligned} \Vert \nabla ^{-1}{\bar{P_r}} M_\delta P_r&\varphi \Vert _{L^2} \le \Vert \nabla ^{-1}{\bar{P_r}} M_\delta ' P_r\varphi \Vert _{L^2} + \Vert \nabla ^{-1}{\bar{P_r}} M_\delta '' P_r\varphi \Vert _{L^2}\nonumber \\&\lesssim \delta ^{-1} \Vert V \Vert _{L^2_{\mathrm{per}}} \Vert f\Vert _{L^2} + \Vert \nabla f\Vert _{L^2}, \end{aligned}$$
(C. 30)

where V is given in (4.28). Using \(\Vert V\Vert _{L^2_{\mathrm{per}}} \le \Vert V\Vert _{L^\infty }^{1/2} \Vert V\Vert _{L_{\mathrm{per}}^1}^{1/2}\) and the definition \(m:=\Vert V\Vert _{L^1_{\mathrm{per}}}\) in (C. 30) gives

$$\begin{aligned} \Vert \nabla ^{-1}{\bar{P_r}}&M_\delta f\Vert _{\dot{H}^1} \lesssim \Vert V \Vert _{L^\infty }^{1/2} (\delta ^{-1} m^{1/2}\Vert f\Vert _{L^2}) + \Vert \nabla f\Vert _{L^2} \, . \end{aligned}$$
(C. 31)

Lemma B.2 and definition (3.45) imply (C. 29). \(\square \)

Appendix D: Refined Nonlinear Estimates

Let \(N_\delta \) be given implicitly by (3.31) and recall the definition of the \(B_{s,\delta }\) norm from (3.45). Let \(\dot{H}^{0}\equiv L^2\). In this section we prove estimates on \(N_\delta \).

Proposition D.1

Let Assumption [A1] hold. If \(\Vert \varphi _1\Vert _{B_{s,\delta }}, \Vert \varphi _2\Vert _{B_{s,\delta }} = o(\delta ^{-1/2})\), then we have the estimate

$$\begin{aligned} \Vert N_\delta&(\varphi _1) - N_\delta (\varphi _2)\Vert _{L^2} \nonumber \\&\lesssim e^{-\beta }m^{ -1/3} \delta ^{-1/2}(\Vert \varphi _1\Vert _{B_{s,\delta }} + \Vert \varphi _2\Vert _{B_{s, \delta }})\Vert \varphi _1 - \varphi _2\Vert _{\dot{H}^1} . \end{aligned}$$
(D.1)

We derive Proposition D.1 from its version with \(\delta = 1\) by rescaling. For \(\delta = 1\), we have the following result.

Proposition D.2

Let Assumption [A1] hold \(\Vert \psi \Vert _{L^2} = o(1)\). Then \(N:=N_{\delta =1}\) satisfies the estimate

$$\begin{aligned} \Vert N(\psi _1) - N(\psi _2)\Vert _{L^2}&\lesssim \sum _{j=1}^2\bigg [ (\Vert \psi _j\Vert _{\dot{H}^1})\Vert \psi _1 - \psi _2\Vert _{\dot{H}^1} \nonumber \\&\quad + e^{-\beta } \big (\Vert \psi _j\Vert _{\dot{H}^1}^{1/3} \Vert \psi _j\Vert _{L^2}^{2/3} \Vert \psi _1 - \psi _2\Vert _{\dot{H}^{1}} \nonumber \\&\quad + \Vert \psi _j\Vert _{\dot{H}^{1}} \Vert \psi _1 - \psi _2\Vert _{\dot{H}^{1}}^{1/3}\Vert \psi _1 - \psi _2\Vert _{L^2}^{2/3}\big )\bigg ]. \end{aligned}$$
(D.2)

The derivation of Proposition D.1 from Proposition D.2 is same as that of Proposition 5.1 from Proposition 5.2 and we omit it here.

Proof of Proposition D.2

Let \(h_{\mathrm{per}}\) and \(r_{\mathrm{per}}(z)\) be given in (3.8). We use the relations (5.6)–(5.12) in the proof of Proposition 5.2. Following the latter proof we see that it suffices to improve the estimate of \(N_k(\psi )\) in Proposition 5.3, to which we proceed. \(\square \)

Proposition D.3

Let Assumption [A1] hold and let \(N_k\) be given by (5.12). Assume that \(\Vert \nabla \psi \Vert _{L^2} = o(1)\), then, for any \(k\ge 2\), we have the estimate

$$\begin{aligned}&\Vert {\text {den}}[N_k(\psi )]\Vert _{L^2} \lesssim \Vert \nabla \psi \Vert _{L^2}^k + e^{-\beta } \Vert \nabla \psi \Vert _{L^2}^{4/3} \Vert \psi \Vert _{L^2}^{2/3}\delta _{k, 2}, \end{aligned}$$
(D.3)

where the constants associated with \(\lesssim \) are independent of \(\beta \) and \(\delta _{k, 2}\) is the Kronecker delta.

Proof

We begin with \(k=2\). To improve upon estimate (5.13), we, following [8], use the partition of unity

$$\begin{aligned} P_1+ P_2= \mathbf {1}, \text { with } P_1 := \chi _{h_{\mathrm{per}} < \mu } \text { and } P_2 := \chi _{h_{\mathrm{per}} \ge \mu }. \end{aligned}$$
(D.4)

Let \(R_i \equiv R_i (z)= r_{\mathrm{per}}(z) P_i\) where \(i=1, 2,\)\(P_i\) and \(r_{\mathrm{per}}(z)\) are given in (D.4) and (3.8). Recalling definition (5.12) of \(N_k(\psi )\) and inserting the partition of unity, \(P_1 + P_2 = \mathbf {1}\), after each R in the integrand of (5.12), we arrive at

$$\begin{aligned} N_2(\psi ) = \sum _{a,b,c = 1,2}N^{abc}_2(\psi ) \, , \end{aligned}$$
(D.5)

where the \(N_2^{abc}(\psi )\), for \(a,b,c =1, 2\), denote the operators

$$\begin{aligned} N^{abc}_2(\psi ) =&\oint R_a \phi R_b \psi R_c \, . \end{aligned}$$
(D.6)

We estimate the terms individually. Below, we use the estimate (see (4.32))

$$\begin{aligned} \Vert (1-\Delta )^{\alpha } R_i(z)\Vert \lesssim \&d^{\alpha -1}\lesssim 1,\ i=1, 2, \end{aligned}$$
(D.7)

for \(\alpha \in [0, 1]\) and \(z\in \Gamma \), where

$$\begin{aligned} d \equiv d(z):=\mathrm {dist}(z, \sigma (h_{\mathrm{per}}))\ge \frac{1}{4} . \end{aligned}$$
(D.8)

Case 1 (121) and (212). We estimate the case for (121), the other case is done similarly. Since \(P_1P_2 = 0\), we write

$$\begin{aligned} N^{(121)}_2(\psi )&= \oint R_1 \psi R_2 P_2 \psi R_1 \end{aligned}$$
(D.9)
$$\begin{aligned}&= \oint R_1 [P_1, \psi ]P_2 R_2 P_2 [\psi , P_1] R_1 \, . \end{aligned}$$
(D.10)

Applying Lemma 2.1 and Eq. (D.7) to the r.h.s. and using that the operator norm is bounded by the \(I^2\) norm, we find

$$\begin{aligned} \Vert {\text {den}}[N^{(121)}_2(\psi )] \Vert _{L^2}&\lesssim \Vert (1-\Delta )^{3/4+\epsilon } N^{(121)}_2(\phi )\Vert _{S^2} \end{aligned}$$
(D.11)
$$\begin{aligned}&\lesssim \left| \oint \right| d^{-3} \Vert [P_1, \psi ]P_2 \Vert _{S^2}^2. \end{aligned}$$
(D.12)

where, recall,

$$\begin{aligned} \left| \oint \right| :=&\frac{1}{2\pi } \int _\Gamma dz |f_{T} (z-\mu )|. \end{aligned}$$
(D.13)

A key observation allowing us to obtain an improved estimate is that the commutators lead to gradient estimates:

Lemma D.4

Let Assumption [A1] hold, we have the estimate

$$\begin{aligned} \Vert [P_i, \psi ] \Vert _{S^2} \lesssim \Vert \nabla \psi \Vert _{L^2},\ i=1, 2 . \end{aligned}$$
(D.14)

Proof of Lemma D.4

Since the identity commutes with any operator and \(P_2 = \mathbf {1}- P_1\) (see (D.4)), we prove the lemma for \(P_1\) only. Since \(h_{\mathrm{per}}\) (see (3.8)) has a gap at \(\mu \), the Cauchy integral formula implies

$$\begin{aligned} P_1 = \frac{1}{2\pi i} \int _{\Gamma _1} (z-h_{\mathrm{per}})^{-1} = \frac{1}{2\pi i} \int _{\Gamma _1} r_{\mathrm{per}}(z) \end{aligned}$$
(D.15)

where \(\Gamma _1\) is the contour \(\{t + i ; -c\le t< \mu \} \cup \{t - i ; -c\le t < \mu \} \cup \{ -c -it + (1-t)i : t \in [0,1] \} \cup \{ \mu -it + (1-t)i : t \in [0,1] \}\), where \(c > 0\) is any constant such that \(h_{\mathrm{per}} > -c +1\), and the contour is traversed counter-clockwise. We see that

$$\begin{aligned}{}[P_1, \psi ]&= \frac{1}{2\pi i} \int _{\Gamma _1} [ r_{\mathrm{per}}(z) ,\psi ] \end{aligned}$$
(D.16)
$$\begin{aligned}&= \frac{1}{2\pi i} \int _{\Gamma _1} r_{\mathrm{per}}(z) [\nabla \cdot , \nabla \psi ] r_{\mathrm{per}}(z) \nonumber \\&\quad + \frac{1}{2\pi i} \int _{\Gamma _1} r_{\mathrm{per}}(z) (2\nabla \psi \cdot \nabla ) r_{\mathrm{per}}(z). \, \end{aligned}$$
(D.17)

Lemma D.4 is now proved by an application of the Kato–Seiler–Simon inequality ((2.14)) to (D.17) and noting that \(\Gamma _1\) is compact and has length O(1). \(\square \)

Using Lemma D.4 and estimates (3.6) and (D.8) in (D.12) yields that

$$\begin{aligned} \Vert {\text {den}}[N^{(121)}_2(\psi )] \Vert _{L^2} \lesssim \Vert \nabla \psi \Vert _{L^2}^2 \, . \end{aligned}$$
(D.18)

Case 2: (112), (211), (122), (221) We estimate the case for (112), the other cases are done similarly. Again, since \(P_1P_2 = 0\), we write

$$\begin{aligned} N^{(112)}_2(\psi )&= \oint R_1 \psi R_1 \psi R_2 \end{aligned}$$
(D.19)
$$\begin{aligned}&= \oint R_1 \psi R_1 [\psi , P_1] R_2 \, . \end{aligned}$$
(D.20)

Using Lemma 2.1 as in with \(N^{(121)}_2(\psi )\) in (D.11), we estimate (D.20) as

$$\begin{aligned} \Vert {\text {den}}[N^{(112)}_2(\psi )] \Vert _{L^2}&\lesssim \left| \oint \right| d^{-1} \Vert \psi R_1 \Vert \Vert [\psi , P_1]R_2 \Vert _{S^2} . \end{aligned}$$
(D.21)

where \(\left| \oint \right| \) is defined in (D.13). By the inequality \(\Vert A\Vert \le \Vert A\Vert _{I^p}\), for any \(p < \infty \) for any operator A on \(L^2(\mathbb {R}^3)\), and the Kato–Seiler–Simon inequality (2.14), we find \(\Vert \psi R_1 \Vert \le \Vert \psi R_1 \Vert _{S^6}\lesssim \Vert \psi \Vert _{L^6}\). Using this, together with Lemma D.4, in (D.21), we obtain

$$\begin{aligned} \Vert {\text {den}}[N^{(112)}_2(\psi )] \Vert _{L^2} \lesssim&\left| \oint \right| d^{-2} \Vert \psi \Vert _{L^6} \Vert \nabla \psi \Vert _{L^2} . \end{aligned}$$
(D.22)

Combining this with (3.6), (D.8) and Hardy–Littlewood’s inequality (2.18) gives

$$\begin{aligned} \Vert {\text {den}}[N^{(112)}_2(\psi )] \Vert _{L^2} \lesssim&\Vert \nabla \psi \Vert _{L^2}^2. \end{aligned}$$
(D.23)

Case 3 (111) and (222). We use the \(L^2\)\(L^2\) duality to estimate the \(L^2\) norm of \({\text {den}}[N^{(qqq)}_2(\psi )], q=1, 2\). We have, by (2.5) and definition (5.12),

$$\begin{aligned} \Vert {\text {den}}[N^{(qqq)}_2(\psi )]\Vert _{L^2}&=\sup _{ \Vert f\Vert _{L^2}=1}\left| \int f {\text {den}}[N^{(qqq)}_2(\psi )]\right| \nonumber \\&=\sup _{ \Vert f\Vert _{L^2}=1}\left| \mathrm {Tr}[f N^{(qqq)}_2(\psi )]\right| \nonumber \\&=\sup _{ \Vert f\Vert _{L^2}=1}\left| \oint \mathrm {Tr}(f R_q \psi R_q \psi R_q)\right| . \end{aligned}$$
(D.24)

(In the last two lines, f is considered as a multiplication operator.) To show that the integral on the r.h.s. converges absolutely, we follow the arguments in (5.17)–(5.20) to prove, for \(q=0, 1\),

$$\begin{aligned} \big |\mathrm {Tr}( f R_q (\psi R_q)^2)\big |\lesssim&d^{-4/3}\Vert f\Vert _{L^2} \Vert \nabla \psi \Vert _{L^2}^{4/3} \Vert \psi \Vert _{L^2}^{2/3}. \end{aligned}$$
(D.25)

Due to definition (D.8) of \(d \equiv d(z)\), this shows that the integral on the r.h.s. of (D.24) converges absolutely.

Lemma D.5

For \(q=1, 2\), we have

$$\begin{aligned} \oint&\mathrm {Tr}[f R P_q g R P_q h P_q] \nonumber \\&= \frac{1}{2\pi i}\int _{\Gamma _q} (f_{T}(z) - 1) \mathrm {Tr}[f R P_q g R P_q h P_q]. \end{aligned}$$
(D.26)

Proof

Note that the contour \(\Gamma \) in Fig. 2 is the union of two disjoint contours, \(\Gamma =\Gamma _1 \cup \Gamma _2\), with \(\Gamma _1\) being the closed contour and \(\Gamma _1\) unbounded one (i.e. the parts of \(\Gamma \) with \(\mathrm {Re}\, z < \mu \) and \(\mathrm {Re}\, z > \mu \)). We first note that, by Bloch’s theory,

$$\begin{aligned} \int _{\Gamma _q} dz \,&\mathrm {Tr}[f R P_q g R P_q h P_q] =\int _{\Gamma _q} dz\, \int _{(\Omega ^*)^3} d{\hat{k}} d{\hat{k}}_1 d {\hat{k}}_2 \end{aligned}$$
(D.27)
$$\begin{aligned}&\times \mathrm {Tr}_{L^2_{\mathrm{per}}} f_{k-k_1} (RP_q)_{k_1} g_{k_1-k_2} (RP_q)_{k_2} h_{k_2-k} (RP_q)_k . \end{aligned}$$
(D.28)

Computing the trace in the complete orthonormal basis of eigenvectors \(\varphi _{m,k}\) of \((RP_q)_{k}\) (with eigevalues \(\lambda _{m,k}\)) and inserting the complete orthonormal bases of eigenvectors \(\varphi _{n,k_1}\) and \(\varphi _{r, k_2}\) of \((RP_q)_{k_1}\) and \((RP_q)_{k_2}\) (with eigevalues \(\lambda _{n,k_1}\) and \(\lambda _{r,k_2}\)) into (D.28), we see that

$$\begin{aligned} \int _{\Gamma _1}&\mathrm {Tr}[f R P_q g R P_q h P_q] = \sum _{m,n,r} \int _{(\Omega ^*)^3} d{\hat{k}} d{\hat{k}}_1 d {\hat{k}}_2 \end{aligned}$$
(D.29)
$$\begin{aligned}&\times \langle \varphi _{m,k}, f_{k-k_1} \varphi _{n, k_1} \rangle \langle \varphi _{n, k_1}, g_{k_1-k_2} \varphi _{r, k_2} \rangle \langle \varphi _{r, k_2}, h_{k_2-k} \varphi _{m, k} \rangle \end{aligned}$$
(D.30)
$$\begin{aligned}&\times \int _{\Gamma _q} dz\, \frac{1}{(z-\lambda _{m,k})(z-\lambda _{n,k_1})(z-\lambda _{r, k_2})} . \end{aligned}$$
(D.31)

Since \(P_1\) projects to the spectrum of \(h_{\mathrm{per}}\) on the left of \(\mu \), we see that \(\lambda _{m,k}, \lambda _{n,k_1}, \lambda _{r, k_2} < \mu \). In particular, these eigenvalues are in the left closed contour in Fig. 2. Consequently, Cauchy’s integral formula shows that the term in the large bracket in (D.31) is identically zero. Similar argument applies to \(P_2\). This shows that

$$\begin{aligned} \frac{1}{2\pi i}\int _{\Gamma _q} \mathrm {Tr}[f R P_q g R P_q h P_q] = 0. \end{aligned}$$
(D.32)

Thus (D.26) follows. \(\square \)

Using the explicit form of the Fermi–Diract distribution \(f_{T}\) in (1.2), we see that

$$\begin{aligned} |f_{T}(z-\mu ) - 1| = \frac{e^{\beta (\mathrm {Re}\, z-\mu )}}{|1+e^{\beta (z-\mu )}|} . \end{aligned}$$
(D.33)

By condition (1.26) and by the choice of the contour, \(\Gamma \), in Fig. 2, we see that the if \(z \in \Gamma \), then \(\mathrm {Re}\, z\) is at least at the distance \(\ge 1\) from \(\mu \). Hence, for z in a contour \(\Gamma \), (D.33) implies that

$$\begin{aligned} |f_{T}(z-\mu ) - 1| \lesssim e^{-\beta }. \end{aligned}$$
(D.34)

Applying estimates (D.34) and (D.25) to the r.h.s. of (D.26) and recalling the definition (D.8) of \(d\equiv d(z)\ge \frac{1}{4}\),

we arrive at the inequality

$$\begin{aligned}&\left| \oint \mathrm {Tr}[f R P_q g R P_q h P_q]\right| \nonumber \\&\quad \lesssim e^{-\beta }\Vert f\Vert _{L^2} \Vert \nabla \psi \Vert _{L^2}^{4/3} \Vert \psi \Vert _{L^2}^{2/3}. \end{aligned}$$
(D.35)

This inequality, together with the relation (D.24), gives

$$\begin{aligned} \left\| {\text {den}}[N^{(qqq)}_2(\phi )]\right\| _{L^2} \lesssim&e^{-\beta } \Vert \nabla \phi \Vert _{L^2}^{4/3} \Vert \phi \Vert _{L^2}^{2/3}. \end{aligned}$$
(D.36)

Inequalities (D.18), (D.23) and (D.36) imply estimate (D.3) for \(k=2\).

Now we estimate \(N_k\) for \(k>2\). By (3.6) and (3.10), it suffices to estimate \({\text {den}}[R(\phi R)^k]\) where \(R = r_{\mathrm{per}}(z)\) is given in (3.8). Using Lemma 2.1, we see that

$$\begin{aligned} \Vert {\text {den}}[R(\phi R)^k]\Vert _{L^2} \lesssim&\Vert (1-\Delta )^{4/3+\epsilon }R(\phi R)^k\Vert _{S^2} \end{aligned}$$
(D.37)
$$\begin{aligned} \lesssim&\Vert (\phi R)^k\Vert _{S^2}. \end{aligned}$$
(D.38)

Using Hölder’s inequality with \(\frac{1}{2} = \frac{1}{6} + \frac{1}{6} + \frac{1}{6} + \) another k terms of \(\frac{1}{\infty }\), (D.38) becomes

$$\begin{aligned} \Vert {\text {den}}[R(\phi R)^k]\Vert _{L^2} \le&\Vert \phi R\Vert _{S^6}^3 \Vert \phi R\Vert ^{k-3} \end{aligned}$$
(D.39)
$$\begin{aligned} \le&\Vert \phi R\Vert _{S^6}^k , \end{aligned}$$
(D.40)

where the last line follows since \(\Vert \cdot \Vert \le \Vert \cdot \Vert _{S^p}\) for \(p < \infty \). Combining with Kato–Seiler–Simon’s inequality (2.14) and Hardy-Littlewood’s inequality (2.18), (D.40) implies (D.3) for \(k\ge 3\). \(\square \)

The rest of the proof of Proposition D.1 proceeds as the proof of in Proposition 5.3. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chenn, I., Sigal, I.M. On Derivation of the Poisson–Boltzmann Equation. J Stat Phys 180, 954–1001 (2020). https://doi.org/10.1007/s10955-020-02562-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-020-02562-8

Keywords

Navigation