Skip to main content
Log in

On Abrikosov Lattice Solutions of the Ginzburg-Landau Equations

  • Published:
Mathematical Physics, Analysis and Geometry Aims and scope Submit manuscript

Abstract

We prove existence of Abrikosov vortex lattice solutions of the Ginzburg-Landau equations of superconductivity, with multiple magnetic flux quanta per fundamental cell. We also revisit the existence proof for the Abrikosov vortex lattices, streamlining some arguments and providing some essential details missing in earlier proofs for a single magnetic flux quantum per a fundamental cell.

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

References

  1. Abrikosov, A.A.: On the magnetic properties of superconductors of the second group. J. Explt. Theoret. Phys. 32, 1147–1182 (1957)

    Google ScholarΒ 

  2. Aftalion, A., Serfaty, S.: Lowest Landau level approach in superconductivity for the Abrikosov lattice close to H c2. Selecta Math. (N.S.) 13, 183–202 (2007)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  3. Almog, Y.: On the bifurcation and stability of periodic solutions of the Ginzburg-Landau equations in the plane. SIAM J. Appl. Math. 61, 149–171 (2000)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  4. Almog, Y.: Abrikosov lattices in finite domains. Commun. Math. Phys. 262, 677–702 (2006)

    ArticleΒ  ADSΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  5. Barany, E., Golubitsky, M., Turski, J.: Bifurcations with local gauge symmetries in the Ginzburg-Landau equations. Phys. D 56, 36–56 (1992)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  6. Chapman, S.J.: Nucleation of superconductivity in decreasing fields. European J. Appl. Math. 5, 449–468 (1994)

    MathSciNetΒ  MATHΒ  Google ScholarΒ 

  7. Chapman, S.J., Howison, S.D., Ockedon, J.R.: Macroscopic models of superconductivity. SIAM Rev. 34, 529–560 (1992)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  8. Chouchkov, D., Ercolani, N.M., Rayan, S., Sigal, I.M.: Ginzburg-Landau equations on Riemann surfaces of higher genus. arXiv:1704.03422 (2017)

  9. Du, Q., Gunzburger, M.D., Peterson, J.S.: Analysis and approximation of the Ginzburg-Landau model of superconductivity. SIAM Rev. 34, 54–81 (1992)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  10. Dubrovin, D.A., Fomenko, A.T., Novikov, S.P.: Modern geometry – methods and applications. Part I. The geometry of sufraes, transformation groups, and fields. 2nd Edition. Springer-Verlag, Berlin (1984)

    MATHΒ  Google ScholarΒ 

  11. Dutour, M.: Phase diagram for Abrikosov lattice. J. Math. Phys. 42, 4915–4926 (2001)

    ArticleΒ  ADSΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  12. Dutour, M.: Bifurcation vers l β€² Γ©tat dAbrikosov et diagramme des phases. Thesis Orsay . arXiv:math-ph/9912011

  13. Eilenberger, G., Zu, A.: Theorie der periodischen LΓΆsungen der GL-Gleichungen fΓΌr Supraleiter 2. Z. Physik 180, 32–42 (1964)

    ArticleΒ  MathSciNetΒ  Google ScholarΒ 

  14. Fournais, S., Helffer, B.: Spectral methods in surface superconductivity. progress in nonlinear differential equations and their applications, Vol 77. BirkhΓ€user, Boston (2010)

  15. Gustafson, S.J., Sigal, I.M.: Mathematical concepts of quantum mechanics. Springer, Berlin (2006)

    MATHΒ  Google ScholarΒ 

  16. Gustafson, S.J., Sigal, I.M., Tzaneteas, T.: Statics and dynamics of magnetic vortices and of Nielsen-Olesen (Nambu) strings. J. Math. Phys. 51, 015217 (2010)

    ArticleΒ  ADSΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  17. Jaffe, A., Taubes, C.: Vortices and monopoles: structure of static gauge theories. Progress in Physics 2. BirkhΓ€user, Boston (1980)

    MATHΒ  Google ScholarΒ 

  18. Kleiner, W.H., Roth, L.M., Autler, S.H.: Bulk solution of Ginzburg-Landau equations for type II superconductors: upper critical field region. Phys. Rev. 133, A1226β€”A1227 (1964)

    ArticleΒ  ADSΒ  MATHΒ  Google ScholarΒ 

  19. Lasher, G.: Series solution of the Ginzburg-Landau equations for the Abrikosov mixed state. Phys. Rev. 140, A523β€”A528 (1965)

    ArticleΒ  ADSΒ  MathSciNetΒ  Google ScholarΒ 

  20. Odeh, F.: Existence and bifurcation theorems for the Ginzburg-Landau equations. J. Math. Phys. 8, 2351–2356 (1967)

    ArticleΒ  ADSΒ  Google ScholarΒ 

  21. Ovchinnikov, Y.N.: Structure of the supercponducting state near the critical fiel H c2 for values of the Ginzburg-Landau parameter ΞΊ close to unity. JETP 85(4), 818–823 (1997)

    ArticleΒ  ADSΒ  MathSciNetΒ  Google ScholarΒ 

  22. Rubinstein, J.: Six Lectures on Superconductivity. Boundaries, interfaces, and transitions (Banff, AB, 1995), 163–184, CRM Proc. Lecture Notes, 13, Amer. Math. Soc., Providence, RI (1998)

  23. Sandier, E., Serfaty, S.: Vortices in the magnetic ginzburg-landau model. Progress in nonlinear differential equations and their applications, vol. 70. BirkhΓ€user, Boston (2007)

    MATHΒ  Google ScholarΒ 

  24. Sigal, I.M.: Magnetic Vortices, Abrikosov Lattices and Automorphic Functions, in Mathematical and Computational Modelling (With Applications in Natural and Social Sciences, Engineering, and the Arts). Wiley, New York (2014)

    Google ScholarΒ 

  25. TakÑč, P.: Bifurcations and vortex formation in the Ginzburg-Landau equations. Z. Angew. Math. Mech. 81, 523–539 (2001)

    MathSciNetΒ  MATHΒ  Google ScholarΒ 

  26. Tzaneteas, T., Sigal, I.M.: Abrikosov lattice solutions of the Ginzburg-Landau equations. Contem. Math. 535, 195–213 (2011)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

  27. Tzaneteas, T., Sigal, I.M.: On Abrikosov lattice solutions of the Ginzburg-Landau equations. Math. Model. Nat. Phenom. 8(5), 190–205 (2013)

    ArticleΒ  MathSciNetΒ  MATHΒ  Google ScholarΒ 

Download references

Acknowledgments

It is a pleasure to thank Max Lein for useful discussions and the anonymous referees for reading carefully the manuscript and many useful remarks and suggestions. The first author would like to thank Dmitri Chouchkov for useful discussions. The first and third authors’ research is supported in part by NSERC Grant No. NA7901. During the work on the paper, they enjoyed the support of the NCCR SwissMAP. The first author was also supported by the NSERC CGS program. The second author (P. S.) was partially supported by Fondo Basal CMM-Chile and Fondecyt postdoctoral grant 3160055.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Israel Michael Sigal.

Appendices

Appendix

Recall that x and z are related as in (5.11) and identify x = (x1,x2) with x1 + ix2.

A On Solutions of the Linearised Problem

Lemma A.1

There is no linear solution ψ,as in Section 5, such that

$$ \psi(\bar{x}) = e^{ig_{r}(z)} \psi(x) $$
(A.1)

for some real valued g.

Proof

Assume for the sake of contradiction that suchψ exists. Then

$$ \psi(x) = e^{\frac{n}{4}(z^{2}-|z|^{2})} \theta(z) $$
(A.2)

for some holomorphic πœƒ. EquationΒ (A.1) becomes

$$ \theta(\bar{z}) = e^{\frac{n}{4}(z^{2}-\bar{z}^{2})} e^{ig_{r}(z)} \theta(z) $$
(A.3)

Taking βˆ‚ z on both sides, we see that

$$ 0 = e^{\frac{n}{4}(z^{2}-\bar{z}^{2})} e^{ig_{r}(z)} \left( -\frac{n}{2}\bar{z} \theta+i\theta \partial_{z} g_{r} + \theta^{\prime}\right) $$
(A.4)

This shows that the term in the bracket vanishes identically. In particular,

$$ \left( -\frac{n}{2}{z} +i \partial_{z} g_{r}\right)\theta = -\theta^{\prime} $$
(A.5)

Taking \(\partial _{\bar {z}}\)again, we see that

$$ \left( + i \partial_{\bar{z}}\partial_{z} g_{r}\right) \theta = 0 $$
(A.6)

Since πœƒ has at most finitely many zeros, we conclude

$$ -{\Delta} g_{r} = 0 $$
(A.7)

This shows that g r is harmonic. Since it is periodic, it is a constant. EquationΒ (A.3) then shows that such solution isimpossible. β–‘

B Theta Functions

In this appendix we review basic properties of theta functions, which are likely to be known but which we could not find in the literature. From now on, we fix a lattice shape Ο„ and a lattice

$$ \mathcal{L}_{\tau} = \mathbb{Z} + \tau \mathbb{Z} $$
(B.1)

throughout this appendix (unless otherwise stated).

3.1 B.1 Basic Properties

In this section, we prove some basic properties of the theta functions. Let n be fixed. Define for 0 ≀ m ≀ n βˆ’β€‰1,

$$\begin{array}{@{}rcl@{}} \theta_{n,m}(z) = \sum\limits_{l \in [m]_{n}} \gamma^{l^{2}} e^{2\pi i l z} \end{array} $$
(B.2)

where Ξ³ := eΟ€iΟ„/n and \([m]_{n} = \{ a \in \mathbb {Z} \ : \ a = m \bmod n \}\).

Theorem B.1

The πœƒ n, m ’s form a basis for V n that satisfy

  1. 1.

    \(\theta _{n,m}(z+\frac {1}{n}) = e^{2\pi i m/n} \theta _{n,m}(z)\)

  2. 2.

    πœƒn,m(βˆ’z) = πœƒn,nβˆ’m(z)

  3. 3.

    πœƒn,m(z + Ο„/n) = Ξ³βˆ’β€‰1e2Ο€izπœƒn, m+ 1(z)

Theorem B.2

Any n-theta function has exactly n zeros modulo translation by lattice elements. Moreover, any two theta functions that share the same zeros (counting multiplicity) are linearly dependent.

Theorem B.3

πœƒ 1,0 has a simple zero at \(\frac {1}{2}(1+\tau )\)

Proof

See Proposition 11.1.β–‘

Theorem B.4

Suppose that πœƒ ∈ V n and Οƒ ∈ V m , then πœƒΟƒ ∈ Vn + m.

Proof

Inspection.β–‘

The proof of the theorems consists of the following lemmas:

Proof

of Theorem A.2 Expanding in e2Ο€ikz for \(k \in \mathbb {Z}\), the coefficients of any element of V n satisfies the recurssion cm + n = c m ei(2m + n)πτ. This recursion implies that for 0 ≀ m ≀ n βˆ’β€‰1, we have that

$$ c_{m+ln} = c_{m} e^{i\pi\tau(l^{2}n + 2lm)} $$
(B.3)

where l is an integer. So the functions

$$ \sum\limits_{k \in \mathbb{Z}} e^{i\pi\tau(k^{2}n + 2km)} e^{2\pi i(nk +m)z} ,\ m = 0,...,n-1 $$
(B.4)

form a basis for the eigenspace. If we let l = kn + m, then we can rewrite the above as

$$ \sum\limits_{l \in [m]_{n}} e^{i\pi\tau \frac{l^{2}-m^{2}}{n}} e^{2\pi ilz} = e^{-i\pi m^{2}/n} \theta_{m} $$
(B.5)

Now we prove the three bullet points. We note that

$$ \theta_{m}\left( z+\frac{1}{n}\right) =\sum\limits_{l \in [m]_{n}} \gamma^{l^{2}} e^{2\pi i l z} e^{2\pi i l/n} $$
(B.6)

Since l ∈ [m] n , we have that \(l/n - m/n \in \mathbb {Z}\). Hence

$$ \theta_{m}\left( z+\frac{1}{n}\right) = e^{2\pi i m/n} \theta_{m}(z) $$
(B.7)

Now for the second item, we note

$$\begin{array}{@{}rcl@{}} \theta_{m}(-z) &=& \sum\limits_{l \in [m]_{n}} \gamma^{l^{2}} e^{-2\pi i l z} \end{array} $$
(B.8)
$$\begin{array}{@{}rcl@{}} &=& \sum\limits_{l \in [m]_{n}} \gamma^{(-l)^{2}} e^{2\pi i (-l) z} \end{array} $$
(B.9)
$$\begin{array}{@{}rcl@{}} &=& \sum\limits_{l \in [n-m]_{n}} \gamma^{l^{2}} e^{2\pi i l z} \end{array} $$
(B.10)
$$\begin{array}{@{}rcl@{}} &=& \theta_{n-m \bmod n}(z) \end{array} $$
(B.11)

Finally, recalling that Ξ³ = eΟ€iΟ„/n, we note that

$$\begin{array}{@{}rcl@{}} \theta_{m}(z+\tau/n) &=& \sum\limits_{k \in \mathbb{Z}} \gamma^{(kn+m)^{2}} e^{2\pi i(kn+m)z + 2\pi i(kn+m)\tau/n} \end{array} $$
(B.12)
$$\begin{array}{@{}rcl@{}} &=& \gamma^{-1} \sum\limits_{k \in \mathbb{Z}} \gamma^{k^{2}n^{2}+ 2knm+m^{2}+ 2kn+ 2m + 1} e^{2\pi i(kn+m)z} \end{array} $$
(B.13)
$$\begin{array}{@{}rcl@{}} &=& \gamma^{-1} \sum\limits_{k \in \mathbb{Z}} \gamma^{(kn+m + 1)^{2}} e^{2\pi i(kn+m)z} \end{array} $$
(B.14)
$$\begin{array}{@{}rcl@{}} &=& \gamma^{-1} e^{-2\pi i z} \theta_{m + 1 \bmod n}(z) \end{array} $$
(B.15)

β–‘

Proof

of Theorem A.3 First we prove that elements of V n has exactly n zeros modulo translation by lattice elements. We compute the winding number ofπœƒ. First, since πœƒ is holomorphic, its zeros are discrete. Hence we may assume WLOG that all the zeros are in the interior of the fundamental domain. Let Ξ© denote the fundamental domain. Then the total number of zeros of πœƒ is given by

$$ \frac{1}{2\pi i} {\int}_{\partial {\Omega}} \frac{\theta^{\prime}}{\theta} dz $$
(B.16)

Since πœƒ(z) = πœƒ(z + 1), the integral along the tΟ„ and t Ο„ + 1 for t ∈ [0, 1] is zero. Let y(z) = eβˆ’iπτ eΞ±z where Ξ± = βˆ’β€‰2Ο€i. Since πœƒ(z + Ο„) = ynπœƒ(z) and yβ€² = Ξ±y, we see that πœƒβ€²(z + Ο„) = yn(z)πœƒβ€²(z) + nΞ±yn(z)πœƒ(z). Hence, only the horizonal segment of the line integral contribute:

$$\begin{array}{@{}rcl@{}} \frac{1}{2\pi i} {\int}_{\partial {\Omega}} \frac{\theta^{\prime}}{\theta} dz &\,=\,& \frac{1}{2\pi i}{{\int}_{0}^{1}} \frac{\theta^{\prime}(t)}{\theta(t)} - \frac{\theta^{\prime}(1\,+\,\tau \,-\, t)}{\theta(1\,+\,\tau\,-\,t)} dt \end{array} $$
(B.17)
$$\begin{array}{@{}rcl@{}} &\,=\,& \frac{1}{2\pi i} {{\int}_{0}^{1}} \frac{\theta^{\prime}(t)}{\theta(t)} - \frac{y^{n}(1\,-\,t) \theta^{\prime}(1\,-\, t) \,+\, n \alpha y^{n}(1\,-\,t) \theta(1\,-\,t)}{y^{n}(1\,-\,t) \theta(1\,-\,t)} dt \end{array} $$
(B.18)
$$\begin{array}{@{}rcl@{}} &\,=\,& \frac{1}{2\pi i} {{\int}_{0}^{1}} \frac{\theta^{\prime}(t)}{\theta(t)} - \frac{\theta^{\prime}(1\,-\, t) \,+\, n \alpha \theta(1\,-\,t)}{\theta(1\,-\,t)} dt \end{array} $$
(B.19)
$$\begin{array}{@{}rcl@{}} &\,=\,& \frac{1}{2\pi i} {{\int}_{0}^{1}} \frac{\theta^{\prime}(t)}{\theta(t)} - \frac{\theta^{\prime}(1\,-\, t)}{\theta(1\,-\,t)} dt \,+\, n \end{array} $$
(B.20)
$$\begin{array}{@{}rcl@{}} &\,=\,& n \end{array} $$
(B.21)

Next, we show that any two theta functions that share the same zeros (counting multiplicity) are linearly dependent. Let πœƒ and Ο† be the two nonzero zeta functions that shares the same zeros. Set f(z) = πœƒ(a)/Ο†(z). Wes how that

  1. 1.

    f(z) can be extended analytically to all of \(\mathbb {C}\) and

  2. 2.

    f(z) is doubly periodic.

Certainly f is holomorphic away from zeros of Ο†. We only need to show that f can be extended analytically to zeros of Ο†. But this is precisely the requirement that πœƒ and Ο† share the same zeros (counting multiplicity).

For the second item, we note that

$$ f(z + 1) =\theta(z + 1)/\varphi(z + 1) = \theta(z)/\varphi(z) = f(z) $$
(B.22)

and

$$ f(z+\tau) = \frac{\theta(z+\tau)}{\varphi(z+\tau)} = \frac{e^{-2\pi i nz} e^{-\pi i n \tau} \theta(z)}{e^{-2\pi i nz} e^{-\pi i n \tau} \varphi(z)} = \frac{\theta(z)}{\varphi(z)} = f(z) $$
(B.23)

This shows that f is doubly periodic.

Now, Liouville’s theorem shows that f must be constant. It follows that πœƒ and Ο† are collinear. β–‘

3.2 B.2 Classification of Singular n-theta Functions

Theorem B.5

Let X n be the set of singular n-theta functions mod scaling. Then

$$ X_{n} = \left\{ {\theta_{0}^{n}}\left( z + \frac{1}{n}(a+b\tau) \right) e^{2\pi i b z} : a,b \in \mathbb{Z} \right\} $$
(B.24)

where πœƒ0 is a basis for V1. Moreover, |X n | = n2. The location of zeros of elements in X n form the set

$$ \frac{1}{2}(1+\tau) + \frac{1}{n}(\mathbb{Z} +\tau\mathbb{Z}) $$
(B.25)

As before, we establish the theorem through various lemmas. The idea of the proof is as follows: by Theorem A.3, we may identify elements of X n with the location of their zeros. We attempt to locate the zeros of singular n-theta function first and show that there are only n2 possible locations in a fundamental cell. So |X n | = n2. Then we explicitly construct n2 singular n-theta functions to complete the proof.

To locate the zeros of singular n-theta functions, we study the Wronskian of a particular set of nice basis element: \({\Theta }(z) := \det (\theta ^{(i)}_{j} )\) for i,j ∈{0,...,n βˆ’β€‰1}, where \(\theta ^{(i)}_{j}\) means the i-th derivative of πœƒ j (see (B.2) for definition πœƒ j ).

Proposition B.6

The function Θ is holomorphic and

  1. 1.

    The locations of the zeros of Θ are exactly the locations where a singular n-theta function can have zero.

  2. 2.

    Θ(βˆ’z) = (βˆ’β€‰1)n+ 1Θ(z),

  3. 3.

    Θ(z + 1/n) = (βˆ’β€‰1)n+ 1Θ(z),

  4. 4.

    Θ(z + Ο„/n) = (βˆ’β€‰1)n+ 1Ξ³n(nβˆ’β€‰1)ynΘ(z) where y = eβˆ’iπτeΞ±z and Ξ± = βˆ’β€‰2Ο€i.

Proof

We recall that the πœƒ m ’s form a basis for V n . If \(\theta (z) = {\sum }_{m} a^{m} \theta _{m}(z)\) has n zeros at z0, then

$$ 0 = \theta^{(i)}(z_{0}) = \sum\limits_{m} a^{m} \theta^{(i)}_{m}(z) $$
(B.26)

for i = 0,...,n βˆ’β€‰1. So the matrix \((\theta ^{(i)}_{j}(z_{0}))\) has a nonzero vector (a0,...,anβˆ’β€‰1) in its kernel. Hence Θ(z0) = 0. Conversely, if Θ(z0) = 0, then we can find a nonzero vector (a0,....,anβˆ’β€‰1) in the kernel of the matrix \((\theta ^{(i)}_{j}(z_{0}))\). Then πœƒ = amπœƒ m has n-zeros at z0.

Recall from Theorem A.2 that πœƒ m (βˆ’z) = πœƒnβˆ’m mod n(z). It follows that \(\theta ^{(k)}_{m}(-z) = (-1)^{k}\theta ^{(k)}_{n-m \bmod n}(z)\). If n is even, then after z↦ βˆ’ z, every even row in the matrix \((\theta ^{(i)}_{j})\)picks up a minus sign, and moreover, we need to interchange the m-th collumn with the(n βˆ’ m mod n)-th collumn for0 < m < n/2. Together wepick up n/2 + n/2 βˆ’ 1 minus signs for Θ. So Θ(βˆ’z) = βˆ’Ξ˜(z). If n is odd, we pick up(n βˆ’β€‰1)/2 minus signs from the even rows and need to interchange (n βˆ’β€‰1)/2 columns. So Θ(βˆ’z) = Θ(z).

Recall from Theorem A.2 that πœƒ m (z + 1/n) = ΞΆmπœƒ m (z) where ΞΆ = e2Ο€i/n. It follows after z↦z + 1/n, the m-th column of \((\theta ^{(i)}_{j})\)picks up a factor of ΞΆmβˆ’β€‰1. Hence \({\Theta }(z + 1/n) = \zeta ^{{\sum }_{k = 0}^{n-1} k}{\Theta }(z) = (-1)^{n + 1}{\Theta }(z)\).

Finally, we recall from Theorem A.2 and the definition y = eβˆ’iπτ e2Ο€iz = Ξ³βˆ’n e2Ο€iz that

$$\begin{array}{@{}rcl@{}} \theta_{m}(z+\tau/n) &=& \gamma^{-1} e^{2\pi i z} \theta_{m + 1 \bmod n}(z) \end{array} $$
(B.27)
$$\begin{array}{@{}rcl@{}} &=& \gamma^{-1} \gamma^{n} y\theta_{m + 1 \bmod n}(z) \end{array} $$
(B.28)
$$\begin{array}{@{}rcl@{}} &=& \gamma^{n-1} \theta_{m + 1 \bmod n}(z) \end{array} $$
(B.29)

Repeated differentiation shows that

$$ \theta_{m}^{(k)}(z+\tau) = \gamma^{n-1} \sum\limits_{i = 0}^{k} {k \choose i} (y)^{(i)} \theta_{m + 1}^{(k-i)}(z) $$
(B.30)

Hence

$$ \left( \theta^{(i)}_{j}(z+\tau/n)\right) = \gamma^{n-1} E\left( \begin{array}{cccccc} y \\ (y)^{\prime} & y \\ (y)^{\prime\prime} & 2(y)\prime & y \\ {\vdots} &&& {\ddots} \\ (y)^{n} && {\cdots} && y \end{array} \right) \left( \theta^{(i)}_{j}(z)\right) $$
(B.31)

where E is the matrix that corresponds to a permutation of columns(1, 2,...,n)↦(2, 3,...,n, 1). It follows that

$$\begin{array}{@{}rcl@{}} &&\det \left( \theta^{(i)}_{j}(z+\tau/n)\right) \\ &=& (-1)^{n + 1}\det \left[ \gamma^{n-1} \left( \begin{array}{cccccc} y \\ (y)^{\prime} & y \\ (y)^{\prime\prime} & 2(y)^{\prime} & y \\ &&& {\ddots} \\ (y)^{n} &&&& y \end{array} \right) \left( \theta^{(i)}_{j}(z)\right) \right] \end{array} $$
(B.32)

(where (βˆ’β€‰1)n+ 1 = det E). Hence Θ(z + Ο„) = (βˆ’β€‰1)n+ 1Ξ³n(nβˆ’β€‰1)ynΘ(z).β–‘

Corollary B.7

\({\Theta } \in V_{n^{2}}\)

Proof

The lemma above shows that

$$ {\Theta}(z + 1) = {\Theta}\left( z+ \sum\limits_{i = 1}^{n} 1/n \right) = (-1)^{(n + 1)n} {\Theta}(z) = {\Theta}(z) $$
(B.33)

We repeat the above proof with Ο„/n replaced by Ο„. Note first that πœƒ(z + Ο„) = eβˆ’β€‰2Ο€inzβˆ’Ο€inΟ„πœƒ(z) for all πœƒ ∈ V n . Set Y = eβˆ’2Ο€inzβˆ’Ο€inΟ„, then we see that

$$ \left( \theta^{(i)}_{j}(z+\tau)\right) = \left( \begin{array}{cccccc} Y \\ (Y)^{\prime} & Y \\ (Y)^{\prime\prime} & 2(Y)^{\prime} & Y \\ {\vdots} &&& {\ddots} \\ (Y)^{n} && {\cdots} && Y \end{array} \right) \left( \theta^{(i)}_{j}(z)\right) $$
(B.34)

Taking det of both sides, we see that \({\Theta }(z+\tau ) = Y^{n} {\Theta }(z) = e^{-2\pi i n^{2} z-\pi i n^{2}} {\Theta }(z)\), which is precisely the defining conditions of elements of \(V_{n^{2}}\).β–‘

Corollary B.8

|X n | = n2.

Proof

The uniqueness Theorem A.3 shows us that |X n | is equal to the number of possible locations of zeros of singular n-theta functions. Proposition A.7 shows that that this is equal to the size of the zero set of Θ mod L Ο„ . Since \({\Theta } \in V_{n^{2}}\). We conclude by Theorem A.3, again, that |X n | = n2.β–‘

Now, we obtain explicit formulae for elements of X n . To do this, we need the following lemma

Lemma B.9

If πœƒ ∈ V n , so is

$$ \gamma(z) = \theta\left( z + \frac{1}{n}(a+b\tau) \right) e^{2\pi i b z} $$
(B.35)

for \(a,b \in \mathbb {Z}\).

Proof

We check that

$$\begin{array}{@{}rcl@{}} \gamma(z + 1) &=& \theta\left( z + \frac{1}{n}(a+b\tau) + 1 \right) e^{2\pi i b z + 2\pi ib} \end{array} $$
(B.36)
$$\begin{array}{@{}rcl@{}} &=& \gamma(z) \end{array} $$
(B.37)

since \(b \in \mathbb {Z}\). Similarly,

$$\begin{array}{@{}rcl@{}} \gamma(z+\tau) &=& \theta\left( z + \frac{1}{n}(a+b\tau) +\tau \right) e^{2\pi i b z + 2\pi i b \tau} \end{array} $$
(B.38)
$$\begin{array}{@{}rcl@{}} &=& e^{-\pi i n \tau - 2\pi i n z - 2\pi i (a+b\tau)} \theta\left( z + \frac{1}{n}(a+b\tau) \right) e^{2\pi i b z + 2\pi i b \tau} \end{array} $$
(B.39)
$$\begin{array}{@{}rcl@{}} &=& e^{-\pi i n \tau - 2\pi i n z} \gamma(z) \end{array} $$
(B.40)

since \(a,b \in \mathbb {Z}\).β–‘

Now, let πœƒ0 be a basis for V1. From Theorems A.4 and A.5, we see that that \({\theta _{0}^{n}} \in X_{n}\), it follows by Lemma A.10 that

$$ \theta_{a,b}(z):={\theta_{0}^{n}}\left( z + \frac{1}{n}(a+b\tau) \right) e^{2\pi i b z} $$
(B.41)

are all in X n for \(a,b \in \mathbb {Z}\). But there are exactly n2 = |X n | number of distinct such functions (mod scaling). So X n is contains exactly these elements. Moreover, by Proposition 11.1, the zero of πœƒ0 is at \(\frac {1}{2}(1+\tau )\). So the zeros of πœƒa,b are located at \(\frac {1}{2}(1+\tau )-\frac {1}{n}(a+b\tau )\).

C Choice of Ο‡ g

The action of point groups is given by

$$ \psi(g x) = e^{i\chi_{g}} \psi(x). $$
(C.1)

for some Ο‡ g , which we determine below.

Proposition C.1

Let \(g \in SH(\mathcal {L})\) and ψ is a linear solution satisfying (C.1), then Ο‡ g are constant.

Proof

We identify \(SH(\mathcal {L})\) as a subset of \(\mathbb {C}\) so that gx is the multiplication of the two complex numbers g and x. Assume that Ο‡ g satisfies (C.1). Since ψ is a linear solution, by (5.11), we can find a holomorphic theta function πœƒ such that πœƒ(x) = h(x)ψ(x) for some smooth, nonvanishing, h with the property \((\bar \partial h)(x) = \frac {b}{2} xh(x)\) where \(\bar {\partial }:=\frac {1}{2}(\partial _{x_1}+i \partial _{x_2})\). Then (C.1) is equivalent to the fact that

$$ H_{g}(x) := h(g x) e^{i\chi_{g}} h(x)^{-1} $$
(C.2)

is holomorphic. Taking \(\overline {\partial }\), this requirement is equivalent to

$$\begin{array}{@{}rcl@{}} 0&=& \bar{\partial}(h(g x) e^{i\chi_{g}} h(x )^{-1}) \end{array} $$
(C.3)
$$\begin{array}{@{}rcl@{}} &=& \;\left( i\bar{\partial} \chi_{g} + \frac{b}{2}\bar{g}gx - \frac{b}{2}x\right)h(g x) e^{i\chi_{g}} h(x)^{-1} . \end{array} $$
(C.4)

Since |g| = 1 and \(h(g x) e^{i\chi _{g}} h(x)^{-1}\)is invertible, we see that

$$ \bar{\partial} \chi_{g} = 0 $$
(C.5)

Since Ο‡ g are real valued, it is a constant. β–‘

As a result of the the proposition, it suffices for us to look for gauge invariant (ψ,A) under actions of \(H(\mathcal {L})\) whose gauge factor \(h_{g}(x)=e^{i \chi _{g}}\) is a constant. Hence we consider spaces of the form

$$ \left\{ \psi\left( R_{\xi}^{-1} i x\right) = \eta \psi(x),\ R_{\xi} A\left( R_{\xi}^{-1} x\right) = \eta^{\prime} A(x)\right \} $$
(C.6)

where \(\eta , \eta ^{\prime } \in \mathbb {C}\). One realises that such space corresponds to irreducible representations of \(H(\mathcal {L})\).

D Table of C 6-Equivariant Theta Functions

Vortex number

Value of r

Theta functions that span Vn,6,r

n = 2

0

πœƒ 0

Β 

2

πœƒ 2

n = 4

0

\({\theta _{0}^{2}}\)

Β 

1

πœƒ 4

Β 

2

πœƒ 0 πœƒ 4

Β 

4

\({\theta _{2}^{2}}\)

n = 6

0

\({\theta _{0}^{3}}, \ {\theta _{2}^{3}}, {\theta _{4}^{2}} \theta _{2}^{-1}\)

Β 

1

πœƒ 0 πœƒ 4

Β 

2

\({\theta _{0}^{2}} \theta _{2}\)

Β 

3

πœƒ 4 πœƒ 2

Β 

4

\(\theta _{0} {\theta _{2}^{2}}\)

n = 8

0

\({\theta _{0}^{4}}, \ \theta _{0} {\theta _{2}^{3}}\)

Β 

1

\({\theta _{0}^{2}} \theta _{4}\)

Β 

2

\({\theta _{2}^{4}}, \ {\theta _{4}^{2}}, \ {\theta _{0}^{3}} \theta _{2}\)

Β 

3

πœƒ 0 πœƒ 4 πœƒ 2

Β 

4

\({\theta _{0}^{2}} {\theta _{2}^{2}}\)

Β 

5

\(\theta _{4} {\theta _{2}^{2}}\)

n = 10

0

\({\theta _{0}^{5}},\ {\theta _{0}^{2}} {\theta _{2}^{3}}\)

Β 

1

\({\theta _{0}^{3}} \theta _{4}, \ \theta _{4} {\theta _{2}^{3}}\)

Β 

2

\({\theta _{0}^{4}} \theta _{2},\ \theta _{0} {\theta _{2}^{4}}, \ \theta _{0} {\theta _{4}^{2}}\)

Β 

3

\({\theta _{0}^{2}} \theta _{4} \theta _{2}\)

Β 

4

\({\theta _{0}^{3}} {\theta _{2}^{2}} , \ {\theta _{2}^{5}}, \ {\theta _{4}^{2}} \theta _{2}\)

Β 

5

\(\theta _{0} \theta _{4} {\theta _{2}^{2}}\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chenn, I., Smyrnelis, P. & Sigal, I.M. On Abrikosov Lattice Solutions of the Ginzburg-Landau Equations. Math Phys Anal Geom 21, 7 (2018). https://doi.org/10.1007/s11040-017-9257-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11040-017-9257-x

Keywords

Mathematics Subject Classification (2010)

Navigation